Psychology and Neuroscience Faculty Database
Psychology and Neuroscience
Arts & Sciences
Duke University

 HOME > Arts & Sciences > pn > Faculty    Search Help Login pdf version printable version 

Publications of Kathleen A. Welsh-Bohmer    :chronological  alphabetical  combined listing:

%% Journal Articles   
@article{fds374980,
   Author = {Farrer, TJ and Bigler, ED and Tsui-Caldwell, YHW and Abildskov, TJ and Tschanz, JT and Welsh-Bohmer, KA},
   Title = {Scheltens ratings, clinical white matter hyperintensities
             and executive functioning in the Cache County Memory
             Study.},
   Journal = {Appl Neuropsychol Adult},
   Pages = {1-7},
   Year = {2023},
   Month = {December},
   url = {http://dx.doi.org/10.1080/23279095.2023.2287140},
   Abstract = {OBJECTIVE: Examine the association between
             neuropsychologically assessed executive function and
             clinically identifiable white matter burden from magnetic
             resonance imaging, using a visual rating system (Scheltens
             Rating System) applied to the Cache County Memory Study
             (CCMS) archival database. METHOD: We used the Scheltens
             Ratings Scale to quantify white matter lesion burden in the
             CCMS sample and used this metric as a predictor of executive
             function. The sample included 60 individuals with dementia
             and 13 healthy controls. RESULTS: Higher Scheltens ratings
             were associated with poorer task performance on an Executive
             Function composite score of common neuropsychological tests.
             This association held true for both controls and dementing
             cases. CONCLUSIONS: The current findings support extensive
             prior literature demonstrating the association between brain
             vascular health determined by white matter burden and
             clinical outcomes based on neuropsychological assessment of
             cognitive performance.},
   Doi = {10.1080/23279095.2023.2287140},
   Key = {fds374980}
}

@article{fds370292,
   Author = {Blair, EM and Reale, BK and Zahuranec, DB and Forman, J and Langa, KM and Giordani, BJ and Plassman, BL and Welsh-Bohmer, KA and Wang, J and Kollman, CD and Levine, DA},
   Title = {Influence of mild cognitive impairment on patient and care
             partner decision-making for acute ischemic
             stroke.},
   Journal = {J Stroke Cerebrovasc Dis},
   Volume = {32},
   Number = {6},
   Pages = {107068},
   Year = {2023},
   Month = {June},
   url = {http://dx.doi.org/10.1016/j.jstrokecerebrovasdis.2023.107068},
   Abstract = {GOALS: Evidence suggests that patients with mild cognitive
             impairment (MCI) receive fewer treatments for acute ischemic
             stroke and other cardiovascular diseases than patients with
             normal cognition. Little is known about how patient and care
             partner preferences for ischemic stroke treatment differ
             between the patient population with MCI and the population
             with normal cognition. This study aimed to understand how
             patient MCI diagnosis influences patient and care partner
             decision-making for acute ischemic stroke treatments.
             METHODS: Multi-center qualitative study using in-person
             semi-structured interviews with 20 MCI and normal cognition
             patient-care partner dyads using a standard guide. The
             present study reports results on patient and care partner
             preferences for a clinical vignette patient to receive three
             non-invasive treatments (intravenous tissue plasminogen
             activator, inpatient rehabilitation, and secondary
             preventive medications) and two invasive treatments (feeding
             tube and carotid endarterectomy) after acute ischemic
             stroke. We used qualitative content analysis to identify
             themes. FINDINGS: We identified three major themes: (1)
             Patients with MCI desired non-invasive treatments after
             stroke, similar to patients with normal cognition and for
             similar reasons; (2) Patients with MCI expressed different
             preferences than patients with normal cognition for two
             invasive treatments after stroke: carotid endarterectomy and
             feeding tube placement; and (3) Patients with MCI expressed
             more skepticism of the stroke treatment options and less
             decisiveness in decision-making than patients with normal
             cognition. CONCLUSIONS: These results suggest that patient
             MCI diagnosis may contribute to differences in patient and
             care partner preferences for invasive treatments after
             stroke, but not for non-invasive treatments.},
   Doi = {10.1016/j.jstrokecerebrovasdis.2023.107068},
   Key = {fds370292}
}

@article{fds371156,
   Author = {Grazia, A and Altomare, D and Preis, L and Monsch, AU and Cappa, SF and Gauthier, S and Frölich, L and Winblad, B and Welsh-Bohmer, KA and Teipel, SJ and Boccardi, M and Consortium for the Harmonization of
             Neuropsychological Assessment},
   Title = {Feasibility of a standard cognitive assessment in European
             academic memory clinics.},
   Journal = {Alzheimers Dement},
   Volume = {19},
   Number = {6},
   Pages = {2276-2286},
   Year = {2023},
   Month = {June},
   url = {http://dx.doi.org/10.1002/alz.12830},
   Abstract = {INTRODUCTION: Standardized cognitive assessment would
             enhance diagnostic reliability across memory clinics. An
             expert consensus adapted the Uniform Dataset (UDS)-3 for
             European centers, the clinician's UDS (cUDS). This study
             assessed its implementation acceptability and feasibility.
             METHODS: We developed a survey investigating barriers,
             facilitators, and willingness to implement the cUDS. With a
             mixed-methods design, we analyzed data from academic memory
             clinics. RESULTS: Seventy-eight percent of responding
             clinicians were experienced neuropsychologists/psychologists
             and 22% were medical specialists coming from 18 European
             countries. Sixty-five percent clinicians were willing to
             implement cUDS. General barriers related to implementation
             (43%) and clinical-methodological domains (21%). Favorable
             clinicians reported finances (15%) and digitalization (9%)
             as facilitating, but unavailability of local norms (23%) as
             hindering. Unfavorable clinicians reported logistical (23%)
             and time issues (18%). DISCUSSION: Despite challenges, data
             showed moderate clinicians' acceptability and requirements
             to improve feasibility. Nonetheless, these results come from
             academic clinicians. The next steps will require feasibility
             evaluation in non-academic contexts.},
   Doi = {10.1002/alz.12830},
   Key = {fds371156}
}

@article{fds371673,
   Author = {Huggins, LKL and Min, SH and Kaplan, S and Wei, J and Welsh-Bohmer, K and Xu, H},
   Title = {Meta-Analysis of Variations in Association between APOE ɛ4
             and Alzheimer's Disease and Related Dementias Across
             Hispanic Regions of Origin.},
   Journal = {J Alzheimers Dis},
   Volume = {93},
   Number = {3},
   Pages = {1095-1109},
   Year = {2023},
   url = {http://dx.doi.org/10.3233/JAD-221167},
   Abstract = {BACKGROUND: Emerging research has shown racial and ethnic
             variations in the magnitude of association between the
             apolipoprotein ɛ4 (APOE ɛ4) allele and the risk of
             developing Alzheimer's disease and related dementias (ADRD).
             Studies researching this association among Hispanic groups
             within and outside of the United States have produced
             inconsistent results. OBJECTIVE: To examine the association
             between the APOE ɛ4 allele and the risk of developing ADRD
             in global Hispanic populations from different ethnic regions
             of origin. METHODS: PubMed, Embase, Scopus, and PsycInfo
             were searched for studies relating to Hispanic/Latin
             American origin, APOE ɛ4, and ADRD. Odds ratios (OR) of
             ADRD risk for individuals with APOE ɛ4 versus those without
             APOE ɛ4 were extracted and calculated using random effects
             analysis. RESULTS: 20 eligible studies represented Caribbean
             Hispanic, Mexican, South American, Spanish, and Cuban
             groups. Overall, APOE ɛ4 was significantly associated with
             increased risk of ADRD (Odds Ratio [OR] 3.80, 95% CI:
             2.38-6.07). The association was only significant in the
             South American (OR: 4.61, 95% CI: 2.74-7.75) subgroup.
             CONCLUSION: There was an association between APOE ɛ4 and
             increased ADRD risk for the South American subgroup. The
             strength of this association varied across Hispanic
             subgroups. Data is limited with more studies especially
             needed for adjusted analysis on Spanish, Central American,
             Cuban Hispanic, and Caribbean Hispanic groups. Results
             suggest additional environmental or genetic risk factors are
             associated with ethnic variations.},
   Doi = {10.3233/JAD-221167},
   Key = {fds371673}
}

@article{fds372749,
   Author = {Zou, H and Luo, S and Liu, H and Lutz, MW and Bennett, DA and Plassman, BL and Welsh-Bohmer, KA},
   Title = {Genotypic Effects of the TOMM40'523 Variant and APOE on
             Longitudinal Cognitive Change over 4 Years: The TOMMORROW
             Study.},
   Journal = {J Prev Alzheimers Dis},
   Volume = {10},
   Number = {4},
   Pages = {886-894},
   Year = {2023},
   url = {http://dx.doi.org/10.14283/jpad.2023.115},
   Abstract = {BACKGROUND: The 523 poly-T length polymorphism (rs10524523)
             in TOMM40 has been reported to influence longitudinal
             cognitive test performance within APOE ε3/3 carriers. The
             results from prior studies are inconsistent. It is also
             unclear whether specific APOE and TOMM40 genotypes
             contribute to heterogeneity in longitudinal cognitive
             performance during the preclinical stages of AD. OBJECTIVES:
             To determine the effects of these genes on longitudinal
             cognitive change in early preclinical stages of AD, we used
             the clinical trial data from the recently concluded
             TOMMORROW study to examine the effects of APOE and TOMM40
             genotypes on neuropsychological test performance. DESIGN: A
             phase 3, double-blind, placebo-controlled, randomized
             clinical trial. SETTING: Academic affiliated and private
             research clinics in Australia, Germany, Switzerland, the UK,
             and the USA. PARTICIPANTS: Cognitively normal older adults
             aged 65 to 83. INTERVENTION: Pioglitazone tablet.
             MEASUREMENTS: Participants from the TOMMORROW trial were
             stratified based on APOE genotype (APOE ε3/3, APOE ε3/4,
             APOE ε4/4). APOE ε3/3 carriers were further stratified by
             TOMM40'523 genotype. The final analysis dataset consists of
             1,330 APOE ε3/3 carriers and 7,001 visits. Linear mixed
             models were used to compare the rates of decline in
             cognition across APOE groups and the APOE ε3/3 carriers
             with different TOMM40'523 genotypes. RESULTS: APOE ε3/4 and
             APOE ε4/4 genotypes compared with the APOE ε3/3 genotype
             were associated with worse performance on measures of global
             cognition, episodic memory, and expressive language.
             Further, over the four years of observation, the APOE ε3/3
             carriers with the TOMM40'523-S/S genotype showed better
             global cognition and accelerated rates of cognitive decline
             on tests of global cognition, executive function, and
             attentional processing compared to APOE ε3/3 carriers with
             TOMM40'523-S/VL and VL/VL genotypes and compared to the APOE
             ε3/4 and APOE ε4/4 carriers. CONCLUSIONS: We suggest that
             both APOE and TOMM40 genotypes may independently contribute
             to cognitive heterogeneity in the pre-MCI stages of AD.
             Controlling for this genetic variability will be important
             in clinical trials designed to slow the rate of cognitive
             decline and/or prevent symptom onset in preclinical
             AD.},
   Doi = {10.14283/jpad.2023.115},
   Key = {fds372749}
}

@article{fds373574,
   Author = {Watts, A and Haneline, S and Welsh-Bohmer, KA and Wu, J and Alexander,
             R and Swerdlow, RH and Burns, DK and Saunders, AM},
   Title = {TOMM40 '523 Genotype Distinguishes Patterns of Cognitive
             Improvement for Executive Function in APOEɛ3
             Homozygotes.},
   Journal = {J Alzheimers Dis},
   Volume = {95},
   Number = {4},
   Pages = {1697-1707},
   Year = {2023},
   url = {http://dx.doi.org/10.3233/JAD-230066},
   Abstract = {BACKGROUND: TOMM40 '523 has been associated with cognitive
             performance and risk for developing Alzheimer's disease
             independent of the effect of APOE genotype. Few studies have
             considered the longitudinal effect of this genotype on
             change in cognition over time. OBJECTIVE: Our objective was
             to evaluate the relationship between TOMM40 genotype status
             and change in cognitive performance in the TOMMORROW study,
             which was designed to prospectively evaluate an algorithm
             that includes TOMM40 '523 for genetic risk for conversion to
             mild cognitive impairment. METHODS: We used latent growth
             curve models to estimate the effect of TOMM40 allele carrier
             (short, very long) status on the intercept and slope of
             change in cognitive performance in four broad cognitive
             domains (attention, memory, executive function, and
             language) and a combined overall cognitive score over 30
             months. RESULTS: TOMM40 very long allele carriers had
             significantly lower baseline performance for the combined
             overall cognitive function score (B = -0.088,
             p = 0.034) and for the executive function domain score
             (B = -0.143, p = 0.013). Slopes for TOMM40 very long
             carriers had significantly greater increases over time for
             the executive function domain score only. In sensitivity
             analyses, the results for executive function were observed
             in participants who remained clinically stable, but not in
             those who progressed clinically over the study duration.
             CONCLUSIONS: Our results add to the growing body of evidence
             that TOMM40, in the absence of APOEɛ4, may contribute to
             cognitive changes with aging and dementia and support the
             view that mitochondrial function is an important contributor
             to Alzheimer's disease risk.},
   Doi = {10.3233/JAD-230066},
   Key = {fds373574}
}

@article{fds373575,
   Author = {Welsh-Bohmer, KA and Kerchner, GA and Dhadda, S and Garcia, M and Miller, DS and Natanegara, F and Raket, LL and Robieson, W and Siemers,
             ER and Carrillo, MC and Weber, CJ},
   Title = {Decision making in clinical trials: Interim analyses,
             innovative design, and biomarkers.},
   Journal = {Alzheimers Dement (N Y)},
   Volume = {9},
   Number = {4},
   Pages = {e12421},
   Year = {2023},
   url = {http://dx.doi.org/10.1002/trc2.12421},
   Abstract = {The efficient and accurate execution of clinical trials
             testing novel treatments for Alzheimer's disease (AD) is a
             critical component of the field's collective efforts to
             develop effective disease-modifying treatments for AD. The
             lengthy and heterogeneous nature of clinical progression in
             AD contributes to the challenges inherent in demonstrating a
             clinically meaningful benefit of any potential new AD
             therapy. The failure of many large and expensive clinical
             trials to date has prompted a focus on optimizing all
             aspects of decision making, to not only expedite the
             development of new treatments, but also maximize the value
             of the information that each clinical trial yields, so that
             all future clinical trials (including those that are
             negative) will contribute toward advancing the field. To
             address this important topic the Alzheimer's Association
             Research Roundtable convened December 1-2, 2020. The goals
             focused around identifying new directions and actionable
             steps to enhance clinical trial decision making in planned
             future studies.},
   Doi = {10.1002/trc2.12421},
   Key = {fds373575}
}

@article{fds365210,
   Author = {Largent, EA and Walter, S and Childs, N and Dacks, PA and Dodge, S and Florian, H and Jackson, J and Llibre Guerra and JJ and Iturriaga, E and Miller, DS and Moreno, M and Nosheny, RL and Obisesan, TO and Portacolone, E and Siddiqi, B and Silverberg, N and Warren, RC and Welsh-Bohmer, KA and Edelmayer, RM and Participant FIRST Work
             Group},
   Title = {Putting participants and study partners FIRST when clinical
             trials end early.},
   Journal = {Alzheimers Dement},
   Volume = {18},
   Number = {12},
   Pages = {2736-2746},
   Year = {2022},
   Month = {December},
   url = {http://dx.doi.org/10.1002/alz.12732},
   Abstract = {Between 2018 and 2019, multiple clinical trials ended
             earlier than planned, resulting in calls to improve
             communication with and support for participants and their
             study partners ("dyads"). The multidisciplinary Participant
             Follow-Up Improvement in Research Studies and Trials
             (Participant FIRST) Work Group met throughout 2021. Its
             goals were to identify best practices for communicating with
             and supporting dyads affected by early trial stoppage. The
             Participant FIRST Work Group identified 17 key
             recommendations spanning the pre-trial, mid-trial, and
             post-trial periods. These focus on prospectively allocating
             sufficient resources for orderly closeout; developing
             dyad-centered communication plans; helping dyads build and
             maintain support networks; and, if a trial stops, informing
             dyads rapidly. Participants and study partners invest time,
             effort, and hope in their research participation. The
             research community should take intentional steps toward
             better communicating with and supporting participants when
             clinical trials end early. The Participant FIRST
             recommendations are a practical guide for embarking on that
             journey.},
   Doi = {10.1002/alz.12732},
   Key = {fds365210}
}

@article{fds374098,
   Author = {Baloni, P and Arnold, M and Buitrago, L and Nho, K and Moreno, H and Huynh,
             K and Brauner, B and Louie, G and Kueider-Paisley, A and Suhre, K and Saykin, AJ and Ekroos, K and Meikle, PJ and Hood, L and Price, ND and Alzheimer’s Disease Metabolomics Consortium, and Doraiswamy,
             PM and Funk, CC and Hernández, AI and Kastenmüller, G and Baillie, R and Han, X and Kaddurah-Daouk, R},
   Title = {Multi-Omic analyses characterize the ceramide/sphingomyelin
             pathway as a therapeutic target in Alzheimer's
             disease.},
   Journal = {Commun Biol},
   Volume = {5},
   Number = {1},
   Pages = {1074},
   Year = {2022},
   Month = {October},
   url = {http://dx.doi.org/10.1038/s42003-022-04011-6},
   Abstract = {Dysregulation of sphingomyelin and ceramide metabolism have
             been implicated in Alzheimer's disease. Genome-wide and
             transcriptome-wide association studies have identified
             various genes and genetic variants in lipid metabolism that
             are associated with Alzheimer's disease. However, the
             molecular mechanisms of sphingomyelin and ceramide
             disruption remain to be determined. We focus on the
             sphingolipid pathway and carry out multi-omics analyses to
             identify central and peripheral metabolic changes in
             Alzheimer's patients, correlating them to imaging features.
             Our multi-omics approach is based on (a) 2114 human
             post-mortem brain transcriptomics to identify differentially
             expressed genes; (b) in silico metabolic flux analysis on
             context-specific metabolic networks identified differential
             reaction fluxes; (c) multimodal neuroimaging analysis on
             1576 participants to associate genetic variants in
             sphingomyelin pathway with Alzheimer's disease pathogenesis;
             (d) plasma metabolomic and lipidomic analysis to identify
             associations of lipid species with dysregulation in
             Alzheimer's; and (e) metabolite genome-wide association
             studies to define receptors within the pathway as a
             potential drug target. We validate our hypothesis in
             amyloidogenic APP/PS1 mice and show prolonged exposure to
             fingolimod alleviated synaptic plasticity and cognitive
             impairment in mice. Our integrative multi-omics approach
             identifies potential targets in the sphingomyelin pathway
             and suggests modulators of S1P metabolism as possible
             candidates for Alzheimer's disease treatment.},
   Doi = {10.1038/s42003-022-04011-6},
   Key = {fds374098}
}

@article{fds356500,
   Author = {Levine, DA and Galecki, AT and Plassman, BL and Fagerlin, A and Wallner,
             LP and Langa, KM and Whitney, RT and Nallamothu, BK and Morgenstern, LB and Reale, BK and Blair, EM and Giordani, B and Welsh-Bohmer, KA and Kabeto,
             MU and Zahuranec, DB},
   Title = {The Association Between Mild Cognitive Impairment Diagnosis
             and Patient Treatment Preferences: a Survey of Older
             Adults.},
   Journal = {J Gen Intern Med},
   Volume = {37},
   Number = {8},
   Pages = {1925-1934},
   Year = {2022},
   Month = {June},
   url = {http://dx.doi.org/10.1007/s11606-021-06839-w},
   Abstract = {BACKGROUND: Older patients (65+) with mild cognitive
             impairment (MCI) receive less guideline-concordant care for
             cardiovascular disease (CVD) and other conditions than
             patients with normal cognition (NC). One potential
             explanation is that patients with MCI want less treatment
             than patients with NC; however, the treatment preferences of
             patients with MCI have not been studied. OBJECTIVE: To
             determine whether patients with MCI have different treatment
             preferences than patients with NC. DESIGN: Cross-sectional
             survey conducted at two academic medical centers from
             February to December 2019 PARTICIPANTS: Dyads of older
             outpatients with MCI and NC and patient-designated
             surrogates. MAIN MEASURES: The modified Life-Support
             Preferences-Predictions Questionnaire score measured
             patients' preferences for life-sustaining treatment
             decisions in six health scenarios including stroke and acute
             myocardial infarction (range, 0-24 treatments rejected with
             greater scores indicating lower desire for treatment). KEY
             RESULTS: The survey response rate was 73.4%. Of 136
             recruited dyads, 127 (93.4%) completed the survey (66 MCI
             and 61 NC). The median number of life-sustaining treatments
             rejected across health scenarios did not differ
             significantly between patients with MCI and patients with NC
             (4.5 vs 6.0; P=0.55). Most patients with MCI (80%) and NC
             (80%) desired life-sustaining treatments in their current
             health (P=0.99). After adjusting for patient and surrogate
             factors, the difference in mean counts of rejected
             treatments between patients with MCI and patients with NC
             was not statistically significant (adjusted ratio, 1.08, 95%
             CI, 0.80-1.44; P=0.63). CONCLUSION: We did not find evidence
             that patients with MCI want less treatment than patients
             with NC. These findings suggest that other provider and
             system factors might contribute to patients with MCI getting
             less guideline-concordant care.},
   Doi = {10.1007/s11606-021-06839-w},
   Key = {fds356500}
}

@article{fds359831,
   Author = {Hao, N and Wang, Z and Liu, P and Becker, R and Yang, S and Yang, K and Pei,
             Z and Zhang, P and Xia, J and Shen, L and Wang, L and Welsh-Bohmer, KA and Sanders, L and Lee, LP and Huang, TJ},
   Title = {Acoustofluidic multimodal diagnostic system for Alzheimer's
             disease.},
   Journal = {Biosens Bioelectron},
   Volume = {196},
   Pages = {113730},
   Year = {2022},
   Month = {January},
   url = {http://dx.doi.org/10.1016/j.bios.2021.113730},
   Abstract = {Alzheimer's disease (AD) is a progressive and irreversible
             neurodegenerative brain disorder that affects tens of
             millions of older adults worldwide and has significant
             economic and societal impacts. Despite its prevalence and
             severity, early diagnosis of AD remains a considerable
             challenge. Here we report an integrated acoustofluidics-based
             diagnostic system (ADx), which combines triple functions of
             acoustics, microfluidics, and orthogonal biosensors for
             clinically accurate, sensitive, and rapid detection of AD
             biomarkers from human plasma. We design and fabricate a
             surface acoustic wave-based acoustofluidic separation device
             to isolate and purify AD biomarkers to increase the
             signal-to-noise ratio. Multimodal biosensors within the
             integrated ADx are fabricated by in-situ patterning of the
             ZnO nanorod array and deposition of Ag nanoparticles onto
             the ZnO nanorods for surface-enhanced Raman scattering
             (SERS) and electrochemical immunosensors. We obtain the
             label-free detections of SERS and electrochemical
             immunoassay of clinical plasma samples from AD patients and
             healthy controls with high sensitivity and specificity. We
             believe that this efficient integration provides promising
             solutions for the early diagnosis of AD.},
   Doi = {10.1016/j.bios.2021.113730},
   Key = {fds359831}
}

@article{fds362114,
   Author = {Kimmel, HJ and Levine, DA and Whitney, RT and Forman, J and Plassman,
             BL and Fagerlin, A and Welsh-Bohmer, KA and Reale, BK and Galecki, AT and Blair, E and Langa, KM and Giordani, B and Kollman, C and Wang, J and Zahuranec, DB},
   Title = {A Mixed-Methods Study of the Impact of Mild Cognitive
             Impairment Diagnosis on Patient and Care Partner Perception
             of Health Risks.},
   Journal = {J Alzheimers Dis},
   Volume = {85},
   Number = {3},
   Pages = {1175-1187},
   Year = {2022},
   url = {http://dx.doi.org/10.3233/JAD-215155},
   Abstract = {BACKGROUND: Older patients (≥65 years) with mild cognitive
             impairment (MCI) are undertreated for cardiovascular disease
             (CVD). One reason for this disparity could be that patients
             with MCI might underestimate the chances of CVD and
             overestimate dementia. OBJECTIVE: To compare conceptions of
             health risk between older patients with MCI and normal
             cognition (NC) and their care partners. METHODS: We
             conducted a multi-center mixed-methods study of patient-care
             partner dyads completing written quantitative surveys (73%
             response rate; 127 dyads: 66 MCI and 61 NC) or
             semi-structured interviews (20 dyads: 11 MCI, and 9 NC).
             Surveys assessed two-year patient risks of dementia, heart
             attack, stroke, and fall. Interviews assessed similar health
             risks and reasons for risk perceptions. RESULTS: On surveys,
             a similarly low proportion of MCI and NC patients felt they
             were at risk of stroke (5% versus 2%; p = 0.62) and
             heart attack (2% versus 0%; p = 0.99). More MCI than NC
             patients perceived dementia risk (26% versus 2%;
             p < 0.001). Care partners' survey findings were similar.
             Interviews generally confirmed these patterns and also
             identified reasons for future health concerns. For both MCI
             and NC dyads, personal experience with cognitive decline or
             CVD (personal or family history) increased concerns about
             each disease. Additionally, perceptions of irreversibility
             and lack of treatment for cognitive decline increased
             concern about dementia. CONCLUSION: Less use of CVD
             treatments in MCI seems unlikely to be driven by
             differential perceptions of CVD risk. Future work to improve
             awareness of CVD risks in older patients and dementia risk
             in patients with MCI are warranted.},
   Doi = {10.3233/JAD-215155},
   Key = {fds362114}
}

@article{fds366215,
   Author = {Schneider, LS and Bennett, DA and Farlow, MR and Peskind, ER and Raskind, MA and Sano, M and Stern, Y and Haneline, S and Welsh-Bohmer,
             KA and O'Neil, J and Walter, R and Maresca, S and Culp, M and Alexander, R and Saunders, AM and Burns, DK and Chiang, C},
   Title = {Adjudicating Mild Cognitive Impairment Due to Alzheimer's
             Disease as a Novel Endpoint Event in the TOMMORROW
             Prevention Clinical Trial.},
   Journal = {J Prev Alzheimers Dis},
   Volume = {9},
   Number = {4},
   Pages = {625-634},
   Year = {2022},
   url = {http://dx.doi.org/10.14283/jpad.2022.72},
   Abstract = {BACKGROUND: The onset of mild cognitive impairment (MCI) is
             an essential outcome in Alzheimer's disease (AD) prevention
             trials and a compelling milestone for clinically meaningful
             change. Determining MCI, however, may be variable and
             subject to disagreement. Adjudication procedures may improve
             the reliability of these determinations. We report the
             performance of an adjudication committee for an AD
             prevention trial. METHODS: The TOMMORROW prevention trial
             selected cognitively normal participants at increased
             genetic risk for AD and randomized them to low-dose
             pioglitazone or placebo treatment. When adjudication
             criteria were triggered, a participant's clinical
             information was randomly assigned to a three-member panel of
             a six-member independent adjudication committee.
             Determination of whether or not a participant reached MCI
             due to AD or AD dementia proceeded through up to three
             review stages - independent review, collaborative review,
             and full committee review - requiring a unanimous decision
             and ratification by the chair. RESULTS: Of 3494 participants
             randomized, the committee adjudicated on 648 cases from 386
             participants, resulting in 96 primary endpoint events. Most
             participants had cases that were adjudicated once (n = 235,
             60.9%); the rest had cases that were adjudicated multiple
             times. Cases were evenly distributed among the eight
             possible three-member panels. Most adjudicated cases
             (485/648, 74.8%) were decided within the independent review
             (stage 1); 14.0% required broader collaborative review
             (stage 2), and 11.1% needed full committee discussion (stage
             3). The primary endpoint event decision rate was 39/485
             (8.0%) for stage 1, 29/91 (31.9%) for stage 2, and 28/72
             (38.9%) for stage 3. Agreement between the primary event
             outcomes supported by investigators' clinical diagnoses and
             the decisions of the adjudication committee increased from
             50% to approximately 93% (after around 100 cases) before
             settling at 80-90% for the remainder of the study.
             CONCLUSIONS: The adjudication process was designed to
             provide independent, consistent determinations of the trial
             endpoints. These outcomes demonstrated the extent of
             uncertainty among trial investigators and agreement between
             adjudicators when the transition to MCI due to AD was
             prospectively assessed. These methods may inform clinical
             endpoint determination in future AD secondary prevention
             studies. Reliable, accurate assessment of clinical events is
             critical for prevention trials and may mean the difference
             between success and failure.},
   Doi = {10.14283/jpad.2022.72},
   Key = {fds366215}
}

@article{fds371157,
   Author = {Atkins, AS and Kraus, MS and Welch, M and Yuan, Z and Stevens, H and Welsh-Bohmer, KA and Keefe, RSE},
   Title = {Remote self-administration of digital cognitive tests using
             the Brief Assessment of Cognition: Feasibility, reliability,
             and sensitivity to subjective cognitive decline.},
   Journal = {Front Psychiatry},
   Volume = {13},
   Pages = {910896},
   Year = {2022},
   url = {http://dx.doi.org/10.3389/fpsyt.2022.910896},
   Abstract = {Cognitive impairment is a common and pervasive feature of
             etiologically diverse disorders of the central nervous
             system, and a target indication for a growing number of
             symptomatic and disease modifying drugs. Remotely acquired
             digital endpoints have been recognized for their potential
             in providing frequent, real-time monitoring of cognition,
             but their ultimate value will be determined by the
             reliability and sensitivity of measurement in the
             populations of interest. To this end, we describe initial
             validation of remote self-administration of cognitive tests
             within a regulatorily compliant tablet-based platform.
             Participants were 61 older adults (age 55+), including 20
             individuals with subjective cognitive decline (SCD). To
             allow comparison between remote (in-home) and site-based
             testing, participants completed 2 testing sessions 1 week
             apart. Results for three of four cognitive domains assessed
             demonstrated equivalence between remote and site-based
             tests, with high cross-modality ICCs (absolute agreement)
             for Symbol Coding (ICC = 0.75), Visuospatial Working Memory
             (ICC = 0.70) and Verbal Fluency (ICC > 0.73). Group
             differences in these domains were significant and reflected
             sensitivity to objective cognitive impairment in the SCD
             group for both remote and site-based testing (p < 0.05). In
             contrast, performance on tests of verbal episodic memory
             suggested inflated performance during unmonitored testing
             and indicate reliable use of remote cognitive assessments
             may depend on the construct, as well as the population being
             tested.},
   Doi = {10.3389/fpsyt.2022.910896},
   Key = {fds371157}
}

@article{fds372272,
   Author = {Welsh-Bohmer, KA and Byrd, GS and Dewees, R and Bozoki, AC and Martin,
             PM and Plassman, B and Price, SR},
   Title = {Raising Awareness of Alzheimer's Disease and Dementia in
             Native Americans in North Carolina.},
   Journal = {N C Med J},
   Volume = {83},
   Number = {1},
   Pages = {77-78},
   Year = {2022},
   url = {http://dx.doi.org/10.18043/ncm.83.1.77},
   Doi = {10.18043/ncm.83.1.77},
   Key = {fds372272}
}

@article{fds358313,
   Author = {Shadyab, AH and LaCroix, AZ and Feldman, HH and van Dyck, CH and Okonkwo, OC and Tam, SP and Fairchild, JK and Welsh-Bohmer, KA and Matthews, G and Bennett, D and Shadyab, AA and Schafer, KA and Morrison,
             RH and Kipperman, SA and Mason, J and Tan, D and Thomas, RG and Cotman, CW and Baker, LD and ADCS EXERT Study Group},
   Title = {Recruitment of a multi-site randomized controlled trial of
             aerobic exercise for older adults with amnestic mild
             cognitive impairment: The EXERT trial.},
   Journal = {Alzheimers Dement},
   Volume = {17},
   Number = {11},
   Pages = {1808-1817},
   Year = {2021},
   Month = {November},
   url = {http://dx.doi.org/10.1002/alz.12401},
   Abstract = {INTRODUCTION: Effective strategies to recruit older adults
             with mild cognitive impairment (MCI) into nonpharmacological
             intervention trials are lacking. METHODS: Recruitment for
             EXERT, a multisite randomized controlled 18-month trial
             examining the effects of aerobic exercise on cognitive
             trajectory in adults with amnestic MCI, involved a diverse
             portfolio of strategies to enroll 296 participants. RESULTS:
             Recruitment occurred September 2016 through March 2020 and
             was initially slow. After mass mailings of 490,323 age- and
             geo-targeted infographic postcards and brochures,
             recruitment rates increased substantially, peaking at 16
             randomizations/month in early 2020. Mass mailings accounted
             for 52% of randomized participants, whereas 25% were
             recruited from memory clinic rosters, electronic health
             records, and national and local registries. Other sources
             included news broadcasts, public service announcements
             (PSA), local advertising, and community presentations.
             DISCUSSION: Age- and geo-targeted mass mailing of
             infographic materials was the most effective approach in
             recruiting older adults with amnestic MCI into an 18-month
             exercise trial.},
   Doi = {10.1002/alz.12401},
   Key = {fds358313}
}

@article{fds371158,
   Author = {Burns, DK and Alexander, RC and Welsh-Bohmer, KA and Culp, M and Chiang,
             C and O'Neil, J and Evans, RM and Harrigan, P and Plassman, BL and Burke,
             JR and Wu, J and Lutz, MW and Haneline, S and Schwarz, AJ and Schneider,
             LS and Yaffe, K and Saunders, AM and Ratti, E and TOMMORROW study
             investigators},
   Title = {Safety and efficacy of pioglitazone for the delay of
             cognitive impairment in people at risk of Alzheimer's
             disease (TOMMORROW): a prognostic biomarker study and a
             phase 3, randomised, double-blind, placebo-controlled
             trial.},
   Journal = {Lancet Neurol},
   Volume = {20},
   Number = {7},
   Pages = {537-547},
   Year = {2021},
   Month = {July},
   url = {http://dx.doi.org/10.1016/S1474-4422(21)00043-0},
   Abstract = {BACKGROUND: The identification of people at risk of
             cognitive impairment is essential for improving recruitment
             in secondary prevention trials of Alzheimer's disease. We
             aimed to test and qualify a biomarker risk assignment
             algorithm (BRAA) to identify participants at risk of
             developing mild cognitive impairment due to Alzheimer's
             disease within 5 years, and to evaluate the safety and
             efficacy of low-dose pioglitazone to delay onset of mild
             cognitive impairment in these at-risk participants. METHODS:
             In this phase 3, multicentre, randomised, double-blind,
             placebo-controlled, parallel-group study, we enrolled
             cognitively healthy, community living participants aged
             65-83 years from 57 academic affiliated and private research
             clinics in Australia, Germany, Switzerland, the UK, and the
             USA. By use of the BRAA, participants were grouped as high
             risk or low risk. Participants at high risk were randomly
             assigned 1:1 to receive oral pioglitazone (0·8 mg/day
             sustained release) or placebo, and all low-risk participants
             received placebo. Study investigators, site staff, sponsor
             personnel, and study participants were masked to genotype,
             risk assignment, and treatment assignment. The planned study
             duration was the time to accumulate 202 events of mild
             cognitive impairment due to Alzheimer's disease in White
             participants who were at high risk (the population on whom
             the genetic analyses that informed the BRAA development was
             done). Primary endpoints were time-to-event comparisons
             between participants at high risk and low risk given placebo
             (for the BRAA objective), and between participants at high
             risk given pioglitazone or placebo (for the efficacy
             objective). The primary analysis included all participants
             who were randomly assigned, received at least one dose of
             study drug, and had at least one valid post-baseline visit,
             with significance set at p=0·01. The safety analysis
             included all participants who were randomly assigned and
             received at least one dose of study medication. An efficacy
             futility analysis was planned for when approximately 33% of
             the anticipated events occurred in the high-risk, White,
             non-Hispanic or Latino group. This trial is registered with
             ClinicalTrials.gov, NCT01931566. FINDINGS: Between Aug 28,
             2013, and Dec 21, 2015, we enrolled 3494 participants (3061
             at high risk and 433 at low risk). Of those participants,
             1545 were randomly assigned to pioglitazone and 1516 to
             placebo. 1104 participants discontinued treatment (464
             assigned to the pioglitazone group, 501 in the placebo high
             risk group, and 139 in the placebo low risk group). 3399
             participants had at least one dose of study drug or placebo
             and at least one post-baseline follow-up visit, and were
             included in the efficacy analysis. 3465 participants were
             included in the safety analysis (1531 assigned to the
             pioglitazone group, 1507 in the placebo high risk group, and
             427 in the placebo low risk group). In the full analysis
             set, 46 (3·3%) of 1406 participants at high risk given
             placebo had mild cognitive impairment due to Alzheimer's
             disease, versus four (1·0%) of 402 participants at low risk
             given placebo (hazard ratio 3·26, 99% CI 0·85-12·45;
             p=0·023). 39 (2·7%) of 1430 participants at high risk
             given pioglitazone had mild cognitive impairment, versus 46
             (3·3%) of 1406 participants at high risk given placebo
             (hazard ratio 0·80, 99% CI 0·45-1·40; p=0·307). In the
             safety analysis set, seven (0·5%) of 1531 participants at
             high risk given pioglitazone died versus 21 (1·4%) of 1507
             participants at high risk given placebo. There were no other
             notable differences in adverse events between groups. The
             study was terminated in January, 2018, after failing to meet
             the non-futility threshold. INTERPRETATION: Pioglitazone did
             not delay the onset of mild cognitive impairment. The
             biomarker algorithm demonstrated a 3 times enrichment of
             events in the high risk placebo group compared with the low
             risk placebo group, but did not reach the pre-specified
             significance threshold. Because we did not complete the
             study as planned, findings can only be considered
             exploratory. The conduct of this study could prove useful to
             future clinical development strategies for Alzheimer's
             disease prevention studies. FUNDING: Takeda and
             Zinfandel.},
   Doi = {10.1016/S1474-4422(21)00043-0},
   Key = {fds371158}
}

@article{fds358002,
   Author = {Clausen, AN and Bouchard, HC and VA Mid-Atlantic MIRECC Workgroup, and Welsh-Bohmer, KA and Morey, RA},
   Title = {Assessment of Neuropsychological Function in Veterans With
             Blast-Related Mild Traumatic Brain Injury and Subconcussive
             Blast Exposure.},
   Journal = {Front Psychol},
   Volume = {12},
   Pages = {686330},
   Year = {2021},
   url = {http://dx.doi.org/10.3389/fpsyg.2021.686330},
   Abstract = {Objective: The majority of combat-related head injuries are
             associated with blast exposure. While Veterans with mild
             traumatic brain injury (mTBI) report cognitive complaints
             and exhibit poorer neuropsychological performance, there is
             little evidence examining the effects of subconcussive blast
             exposure, which does not meet clinical symptom criteria for
             mTBI during the acute period following exposure. We compared
             chronic effects of combat-related blast mTBI and
             combat-related subconcussive blast exposure on
             neuropsychological performance in Veterans. Methods:
             Post-9/11 Veterans with combat-related subconcussive blast
             exposure (n = 33), combat-related blast mTBI (n = 26), and
             controls (n = 33) without combat-related blast exposure,
             completed neuropsychological assessments of intellectual and
             executive functioning, processing speed, and working memory
             via NIH toolbox, assessment of clinical psychopathology, a
             retrospective account of blast exposures and
             non-blast-related head injuries, and self-reported current
             medication. Huber Robust Regressions were employed to
             compare neuropsychological performance across groups.
             Results: Veterans with combat-related blast mTBI and
             subconcussive blast exposure displayed significantly slower
             processing speed compared with controls. After adjusting for
             post-traumatic stress disorder and depressive symptoms,
             those with combat-related mTBI exhibited slower processing
             speed than controls. Conclusion: Veterans in the
             combat-related blast mTBI group exhibited slower processing
             speed relative to controls even when controlling for PTSD
             and depression. Cognition did not significantly differ
             between subconcussive and control groups or subconcussive
             and combat-related blast mTBI groups. Results suggest
             neurocognitive assessment may not be sensitive enough to
             detect long-term effects of subconcussive blast exposure, or
             that psychiatric symptoms may better account for cognitive
             sequelae following combat-related subconcussive blast
             exposure or combat-related blast mTBI.},
   Doi = {10.3389/fpsyg.2021.686330},
   Key = {fds358002}
}

@article{fds351411,
   Author = {Cocroft, S and Welsh-Bohmer, KA and Plassman, BL and Chanti-Ketterl,
             M and Edmonds, H and Gwyther, L and McCart, M and MacDonald, H and Potter,
             G and Burke, JR},
   Title = {Racially diverse participant registries to facilitate the
             recruitment of African Americans into presymptomatic
             Alzheimer's disease studies.},
   Journal = {Alzheimers Dement},
   Volume = {16},
   Number = {8},
   Pages = {1107-1114},
   Year = {2020},
   Month = {August},
   url = {http://dx.doi.org/10.1002/alz.12048},
   Abstract = {INTRODUCTION: The Alzheimer's Disease Prevention Registry
             (ADPR) of the Joseph and Kathleen Bryan Alzheimer's Disease
             Research Center at Duke University has been successful in
             achieving a racially diverse and "research ready" cohort of
             cognitively healthy volunteers. METHODS: The ADPR is based
             on an infrastructure that includes: (1) an administrative
             leadership team; (2) a coordinating center; (3) an IT
             management team; (4) a community engagement team; and (5)
             collaborations with study partners across disciplines.
             RESULTS: The ADPR currently has more than 4677 members, 26%
             of whom are African American. The ADPR has supported 21
             studies including 8 biomarker studies, 7 clinical trials, 4
             cognitive neuroscience studies, and 2 studies assessing
             novel computerized measures. DISCUSSION: We describe our
             experiences establishing and maintaining a diverse ADPR as
             well as insights on recruitment strategies to increase the
             representation of African Americans in Alzheimer's disease
             studies.},
   Doi = {10.1002/alz.12048},
   Key = {fds351411}
}

@article{fds347140,
   Author = {Blumenthal, JA and Smith, PJ and Mabe, S and Hinderliter, A and Welsh-Bohmer, K and Browndyke, JN and Doraiswamy, PM and Lin, P-H and Kraus, WE and Burke, JR and Sherwood, A},
   Title = {Longer Term Effects of Diet and Exercise on Neurocognition:
             1-Year Follow-up of the ENLIGHTEN Trial.},
   Journal = {J Am Geriatr Soc},
   Volume = {68},
   Number = {3},
   Pages = {559-568},
   Year = {2020},
   Month = {March},
   url = {http://dx.doi.org/10.1111/jgs.16252},
   Abstract = {OBJECTIVES: To evaluate the longer term changes in executive
             functioning among participants with cardiovascular disease
             (CVD) risk factors and cognitive impairments with no
             dementia (CIND) randomized to a diet and exercise
             intervention. DESIGN: A 2 (Exercise) × 2 (Dietary
             Approaches to Stop Hypertension [DASH] eating plan)
             factorial randomized clinical trial. SETTING: Academic
             tertiary care medical center. PARTICIPANTS: Volunteer sample
             of 160 older sedentary adults with CIND and at least one
             additional CVD risk factor enrolled in the ENLIGHTEN trial
             between December 2011 and March 2016. INTERVENTIONS: Six
             months of aerobic exercise (AE), DASH diet counseling,
             combined AE + DASH, or health education (HE) controls.
             MEASUREMENTS: Neurocognitive battery recommended by the
             Neuropsychological Working Group for Vascular Cognitive
             Disorders including measures of executive function, memory,
             and language/verbal fluency. Secondary outcomes included the
             Clinical Dementia Rating-Sum of Boxes (CDR-SB), Six-Minute
             Walk Distance (6MWD), and CVD risk including blood pressure,
             body weight, and CVD medication burden. RESULTS: Despite
             discontinuation of lifestyle changes, participants in the
             exercise groups retained better executive function 1 year
             post-intervention (P = .041) compared with non-exercise
             groups, with a similar, albeit weaker, pattern in the DASH
             groups (P = .054), without variation over time
             (P's > .867). Participants in the exercise groups also
             achieved greater sustained improvements in 6MWD compared
             with non-Exercise participants (P < .001). Participants
             in the DASH groups exhibited lower CVD risk relative to
             non-DASH participants (P = .032); no differences in CVD risk
             were observed for participants in the Exercise groups
             compared with non-Exercise groups (P = .711). In post hoc
             analyses, the AE + DASH group had better performance on
             executive functioning (P < .001) and CDR-SB (P = .011)
             compared with HE controls. CONCLUSION: For participants with
             CIND and CVD risk factors, exercise for 6 months promoted
             better executive functioning compared with non-exercisers
             through 1-year post-intervention, although its clinical
             significance is uncertain. J Am Geriatr Soc 68:559-568,
             2020.},
   Doi = {10.1111/jgs.16252},
   Key = {fds347140}
}

@article{fds352912,
   Author = {Smith, PJ and Mabe, SM and Sherwood, A and Doraiswamy, PM and Welsh-Bohmer, KA and Burke, JR and Kraus, WE and Lin, P-H and Browndyke,
             JN and Babyak, MA and Hinderliter, AL and Blumenthal,
             JA},
   Title = {Metabolic and Neurocognitive Changes Following Lifestyle
             Modification: Examination of Biomarkers from the ENLIGHTEN
             Randomized Clinical Trial.},
   Journal = {J Alzheimers Dis},
   Volume = {77},
   Number = {4},
   Pages = {1793-1803},
   Year = {2020},
   url = {http://dx.doi.org/10.3233/JAD-200374},
   Abstract = {BACKGROUND: Previous studies have demonstrated that aerobic
             exercise (AE) and the Dietary Approaches to Stop
             Hypertension (DASH) diet can improve neurocognition.
             However, the mechanisms by which lifestyle improves
             neurocognition have not been widely studied. We examined the
             associations between changes in metabolic, neurotrophic, and
             inflammatory biomarkers with executive functioning among
             participants from the Exercise and Nutritional Interventions
             for Neurocognitive Health Enhancement (ENLIGHTEN) trial.
             OBJECTIVE: To examine the association between changes in
             metabolic function and neurocognition among older adults
             with cognitive impairment, but without dementia (CIND)
             participating in a comprehensive lifestyle intervention.
             METHODS: ENLIGHTEN participants were randomized using a 2×2
             factorial design to receive AE, DASH, both AE+DASH, or a
             health education control condition (HE) for six months.
             Metabolic biomarkers included insulin resistance
             (homeostatic model assessment [HOMA-IR]), leptin, and
             insulin-like growth factor (IGF-1); neurotrophic biomarkers
             included brain derived neurotrophic factor (BDNF) and
             vascular endothelial growth factor (VEGF); and inflammatory
             biomarkers included interleukin-6 (IL-6) and C-Reactive
             Protein (CRP). RESULTS: Participants included 132 sedentary
             older adults (mean age = 65 [SD = 7]) with CIND.
             Results demonstrated that both AE (d = 0.48,
             p = 0.015) and DASH improved metabolic function
             (d = 0.37, p = 0.039), without comparable
             improvements in neurotrophic or inflammatory biomarkers.
             Greater improvements in metabolic function, including
             reduced HOMA-IR (B = -2.3 [-4.3, -0.2], p = 0.033)
             and increased IGF-1 (B = 3.4 [1.2, 5.7], p = 0.004),
             associated with increases in Executive Function. CONCLUSION:
             Changes in neurocognition after lifestyle modification are
             associated with improved metabolic function.},
   Doi = {10.3233/JAD-200374},
   Key = {fds352912}
}

@article{fds365162,
   Author = {Dodge, HH and Goldstein, FC and Wakim, NI and Gefen, T and Teylan, M and Chan, KCG and Kukull, WA and Barnes, LL and Giordani, B and Hughes, TM and Kramer, JH and Loewenstein, DA and Marson, DC and Mungas, DM and Mattek,
             N and Sachs, BC and Salmon, DP and Willis-Parker, M and Welsh-Bohmer,
             KA and Wild, KV and Morris, JC and Weintraub, S and National Alzheimer's
             Coordinating Center (NACC)},
   Title = {Differentiating among stages of cognitive impairment in
             aging: Version 3 of the Uniform Data Set (UDS)
             neuropsychological test battery and MoCA index
             scores.},
   Journal = {Alzheimers Dement (N Y)},
   Volume = {6},
   Number = {1},
   Pages = {e12103},
   Year = {2020},
   url = {http://dx.doi.org/10.1002/trc2.12103},
   Abstract = {INTRODUCTION: Federally funded Alzheimer's Disease Centers
             in the United States have been using a standardized
             neuropsychological test battery as part of the National
             Alzheimer's Coordinating Center Uniform Data Set (UDS) since
             2005. Version 3 (V3) of the UDS replaced the previous
             version (V2) in 2015. We compared V2 and V3
             neuropsychological tests with respect to their ability to
             distinguish among the Clinical Dementia Rating (CDR) global
             scores of 0, 0.5, and 1. METHODS: First, we matched
             participants receiving V2 tests (V2 cohort) and V3 tests (V3
             cohort) in their cognitive functions using tests common to
             both versions. Then, we compared receiver-operating
             characteristic (ROC) area under the curve in differentiating
             CDRs for the remaining tests. RESULTS: Some V3 tests
             performed better than V2 tests in differentiating between
             CDR 0.5 and 0, but the improvement was limited to Caucasian
             participants. DISCUSSION: Further efforts to improve the
             ability for early identification of cognitive decline among
             diverse racial groups are required.},
   Doi = {10.1002/trc2.12103},
   Key = {fds365162}
}

@article{fds371159,
   Author = {Ma, Y and Jun, GR and Zhang, X and Chung, J and Naj, AC and Chen, Y and Bellenguez, C and Hamilton-Nelson, K and Martin, ER and Kunkle, BW and Bis, JC and Debette, S and DeStefano, AL and Fornage, M and Nicolas, G and van Duijn, C and Bennett, DA and De Jager and PL and Mayeux, R and Haines,
             JL and Pericak-Vance, MA and Seshadri, S and Lambert, J-C and Schellenberg, GD and Lunetta, KL and Farrer, LA and Alzheimer’s
             Disease Sequencing Project and Alzheimer’s Disease Exome
             Sequencing–France Project},
   Title = {Analysis of Whole-Exome Sequencing Data for Alzheimer
             Disease Stratified by APOE Genotype.},
   Journal = {JAMA Neurol},
   Volume = {76},
   Number = {9},
   Pages = {1099-1108},
   Year = {2019},
   Month = {September},
   url = {http://dx.doi.org/10.1001/jamaneurol.2019.1456},
   Abstract = {IMPORTANCE: Previous genome-wide association studies of
             common variants identified associations for Alzheimer
             disease (AD) loci evident only among individuals with
             particular APOE alleles. OBJECTIVE: To identify APOE
             genotype-dependent associations with infrequent and rare
             variants using whole-exome sequencing. DESIGN, SETTING, AND
             PARTICIPANTS: The discovery stage included 10 441
             non-Hispanic white participants in the Alzheimer Disease
             Sequencing Project. Replication was sought in 2 independent,
             whole-exome sequencing data sets (1766 patients with AD,
             2906 without AD [controls]) and a chip-based genotype
             imputation data set (8728 patients with AD, 9808 controls).
             Bioinformatics and functional analyses were conducted using
             clinical, cognitive, neuropathologic, whole-exome
             sequencing, and gene expression data obtained from a
             longitudinal cohort sample including 402 patients with AD
             and 647 controls. Data were analyzed between March 2017 and
             September 2018. MAIN OUTCOMES AND MEASURES: Score, Firth,
             and sequence kernel association tests were used to test the
             association of AD risk with individual variants and genes in
             subgroups of APOE ε4 carriers and noncarriers. Results with
             P ≤ 1 × 10-5 were further evaluated in the
             replication data sets and combined by meta-analysis.
             RESULTS: Among 3145 patients with AD and 4213 controls
             lacking ε4 (mean [SD] age, 83.4 [7.6] years; 4363 [59.3.%]
             women), novel genome-wide significant associations were
             obtained in the discovery sample with rs536940594 in
             AC099552 (odds ratio [OR], 88.0; 95% CI, 9.08-852.0;
             P = 2.22 × 10-7) and rs138412600 in GPAA1 (OR,
             1.78; 95% CI, 1.44-2.2; meta-P = 7.81 × 10-8).
             GPAA1 was also associated with expression in the brain of
             GPAA1 (β = -0.08; P = .03) and its repressive
             transcription factor, FOXG1 (β = 0.13; P = .003),
             and global cognition function (β = -0.53;
             P = .009). Significant gene-wide associations (threshold
             P ≤ 6.35 × 10-7) were observed for OR8G5
             (P = 4.67 × 10-7), IGHV3-7 (P = 9.75 × 10-16),
             and SLC24A3 (P = 2.67 × 10-12) in 2377 patients
             with AD and 706 controls with ε4 (mean [SD] age, 75.2 [9.6]
             years; 1668 [54.1%] women). CONCLUSIONS AND RELEVANCE: The
             study identified multiple possible novel associations for AD
             with individual and aggregated rare variants in groups of
             individuals with and without APOE ε4 alleles that reinforce
             known and suggest additional pathways leading to
             AD.},
   Doi = {10.1001/jamaneurol.2019.1456},
   Key = {fds371159}
}

@article{fds371160,
   Author = {Kunkle, BW and Grenier-Boley, B and Sims, R and Bis, JC and Damotte, V and Naj, AC and Boland, A and Vronskaya, M and van der Lee, SJ and Amlie-Wolf, A and Bellenguez, C and Frizatti, A and Chouraki, V and Martin, ER and Sleegers, K and Badarinarayan, N and Jakobsdottir, J and Hamilton-Nelson, KL and Moreno-Grau, S and Olaso, R and Raybould, R and Chen, Y and Kuzma, AB and Hiltunen, M and Morgan, T and Ahmad, S and Vardarajan, BN and Epelbaum, J and Hoffmann, P and Boada, M and Beecham,
             GW and Garnier, J-G and Harold, D and Fitzpatrick, AL and Valladares, O and Moutet, M-L and Gerrish, A and Smith, AV and Qu, L and Bacq, D and Denning,
             N and Jian, X and Zhao, Y and Del Zompo and M and Fox, NC and Choi, S-H and Mateo, I and Hughes, JT and Adams, HH and Malamon, J and Sanchez-Garcia,
             F and Patel, Y and Brody, JA and Dombroski, BA and Naranjo, MCD and Daniilidou, M and Eiriksdottir, G and Mukherjee, S and Wallon, D and Uphill, J and Aspelund, T and Cantwell, LB and Garzia, F and Galimberti,
             D and Hofer, E and Butkiewicz, M and Fin, B and Scarpini, E and Sarnowski,
             C and Bush, WS and Meslage, S and Kornhuber, J and White, CC and Song, Y and Barber, RC and Engelborghs, S and Sordon, S and Voijnovic, D and Adams,
             PM and Vandenberghe, R and Mayhaus, M and Cupples, LA and Albert, MS and De
             Deyn, PP and Gu, W and Himali, JJ and Beekly, D and Squassina, A and Hartmann, AM and Orellana, A and Blacker, D and Rodriguez-Rodriguez,
             E and Lovestone, S and Garcia, ME and Doody, RS and Munoz-Fernadez, C and Sussams, R and Lin, H and Fairchild, TJ and Benito, YA and Holmes, C and Karamujić-Čomić, H and Frosch, MP and Thonberg, H and Maier, W and Roshchupkin, G and Ghetti, B and Giedraitis, V and Kawalia, A and Li, S and Huebinger, RM and Kilander, L and Moebus, S and Hernández, I and Kamboh, MI and Brundin, R and Turton, J and Yang, Q and Katz, MJ and Concari, L and Lord, J and Beiser, AS and Keene, CD and Helisalmi, S and Kloszewska, I and Kukull, WA and Koivisto, AM and Lynch, A and Tarraga,
             L and Larson, EB and Haapasalo, A and Lawlor, B and Mosley, TH and Lipton,
             RB and Solfrizzi, V and Gill, M and Longstreth, WT and Montine, TJ and Frisardi, V and Diez-Fairen, M and Rivadeneira, F and Petersen, RC and Deramecourt, V and Alvarez, I and Salani, F and Ciaramella, A and Boerwinkle, E and Reiman, EM and Fievet, N and Rotter, JI and Reisch,
             JS and Hanon, O and Cupidi, C and Uitterlinden, AGA and Royall, DR and Dufouil, C and Maletta, RG and de Rojas, I and Sano, M and Brice, A and Cecchetti, R and George-Hyslop, PS and Ritchie, K and Tsolaki, M and Tsuang, DW and Dubois, B and Craig, D and Wu, C-K and Soininen, H and Avramidou, D and Albin, RL and Fratiglioni, L and Germanou, A and Apostolova, LG and Keller, L and Koutroumani, M and Arnold, SE and Panza, F and Gkatzima, O and Asthana, S and Hannequin, D and Whitehead,
             P and Atwood, CS and Caffarra, P and Hampel, H and Quintela, I and Carracedo, Á and Lannfelt, L and Rubinsztein, DC and Barnes, LL and Pasquier, F and Frölich, L and Barral, S and McGuinness, B and Beach,
             TG and Johnston, JA and Becker, JT and Passmore, P and Bigio, EH and Schott, JM and Bird, TD and Warren, JD and Boeve, BF and Lupton, MK and Bowen, JD and Proitsi, P and Boxer, A and Powell, JF and Burke, JR and Kauwe, JSK and Burns, JM and Mancuso, M and Buxbaum, JD and Bonuccelli,
             U and Cairns, NJ and McQuillin, A and Cao, C and Livingston, G and Carlson,
             CS and Bass, NJ and Carlsson, CM and Hardy, J and Carney, RM and Bras, J and Carrasquillo, MM and Guerreiro, R and Allen, M and Chui, HC and Fisher,
             E and Masullo, C and Crocco, EA and DeCarli, C and Bisceglio, G and Dick,
             M and Ma, L and Duara, R and Graff-Radford, NR and Evans, DA and Hodges, A and Faber, KM and Scherer, M and Fallon, KB and Riemenschneider, M and Fardo, DW and Heun, R and Farlow, MR and Kölsch, H and Ferris, S and Leber, M and Foroud, TM and Heuser, I and Galasko, DR and Giegling, I and Gearing, M and Hüll, M and Geschwind, DH and Gilbert, JR and Morris, J and Green, RC and Mayo, K and Growdon, JH and Feulner, T and Hamilton, RL and Harrell, LE and Drichel, D and Honig, LS and Cushion, TD and Huentelman,
             MJ and Hollingworth, P and Hulette, CM and Hyman, BT and Marshall, R and Jarvik, GP and Meggy, A and Abner, E and Menzies, GE and Jin, L-W and Leonenko, G and Real, LM and Jun, GR and Baldwin, CT and Grozeva, D and Karydas, A and Russo, G and Kaye, JA and Kim, R and Jessen, F and Kowall,
             NW and Vellas, B and Kramer, JH and Vardy, E and LaFerla, FM and Jöckel,
             K-H and Lah, JJ and Dichgans, M and Leverenz, JB and Mann, D and Levey, AI and Pickering-Brown, S and Lieberman, AP and Klopp, N and Lunetta, KL and Wichmann, H-E and Lyketsos, CG and Morgan, K and Marson, DC and Brown,
             K and Martiniuk, F and Medway, C and Mash, DC and Nöthen, MM and Masliah,
             E and Hooper, NM and McCormick, WC and Daniele, A and McCurry, SM and Bayer, A and McDavid, AN and Gallacher, J and McKee, AC and van den
             Bussche, H and Mesulam, M and Brayne, C and Miller, BL and Riedel-Heller, S and Miller, CA and Miller, JW and Al-Chalabi, A and Morris, JC and Shaw, CE and Myers, AJ and Wiltfang, J and O'Bryant, S and Olichney, JM and Alvarez, V and Parisi, JE and Singleton, AB and Paulson, HL and Collinge, J and Perry, WR and Mead, S and Peskind, E and Cribbs, DH and Rossor, M and Pierce, A and Ryan, NS and Poon, WW and Nacmias, B and Potter, H and Sorbi, S and Quinn, JF and Sacchinelli, E and Raj, A and Spalletta, G and Raskind, M and Caltagirone, C and Bossù, P and Orfei, MD and Reisberg, B and Clarke, R and Reitz, C and Smith, AD and Ringman, JM and Warden, D and Roberson, ED and Wilcock, G and Rogaeva,
             E and Bruni, AC and Rosen, HJ and Gallo, M and Rosenberg, RN and Ben-Shlomo, Y and Sager, MA and Mecocci, P and Saykin, AJ and Pastor, P and Cuccaro, ML and Vance, JM and Schneider, JA and Schneider, LS and Slifer, S and Seeley, WW and Smith, AG and Sonnen, JA and Spina, S and Stern, RA and Swerdlow, RH and Tang, M and Tanzi, RE and Trojanowski,
             JQ and Troncoso, JC and Van Deerlin and VM and Van Eldik and LJ and Vinters,
             HV and Vonsattel, JP and Weintraub, S and Welsh-Bohmer, KA and Wilhelmsen, KC and Williamson, J and Wingo, TS and Woltjer, RL and Wright, CB and Yu, C-E and Yu, L and Saba, Y and Alzheimer Disease
             Genetics Consortium (ADGC), and European Alzheimer’s Disease
             Initiative (EADI), and Cohorts for Heart and Aging Research in
             Genomic Epidemiology Consortium (CHARGE), and Genetic and Environmental Risk in AD/Defining Genetic and Polygenic and Environmental Risk for Alzheimer’s Disease Consortium
             (GERAD/PERADES), and Pilotto, A and Bullido, MJ and Peters, O and Crane, PK and Bennett, D and Bosco, P and Coto, E and Boccardi, V and De
             Jager, PL and Lleo, A and Warner, N and Lopez, OL and Ingelsson, M and Deloukas, P and Cruchaga, C and Graff, C and Gwilliam, R and Fornage, M and Goate, AM and Sanchez-Juan, P and Kehoe, PG and Amin, N and Ertekin-Taner, N and Berr, C and Debette, S and Love, S and Launer, LJ and Younkin, SG and Dartigues, J-F and Corcoran, C and Ikram, MA and Dickson, DW and Nicolas, G and Campion, D and Tschanz, J and Schmidt, H and Hakonarson, H and Clarimon, J and Munger, R and Schmidt, R and Farrer,
             LA and Van Broeckhoven and C and O'Donovan, MC and DeStefano, AL and Jones,
             L and Haines, JL and Deleuze, J-F and Owen, MJ and Gudnason, V and Mayeux,
             R and Escott-Price, V and Psaty, BM and Ramirez, A and Wang, L-S and Ruiz,
             A and van Duijn, CM and Holmans, PA and Seshadri, S and Williams, J and Amouyel, P and Schellenberg, GD and Lambert, J-C and Pericak-Vance,
             MA},
   Title = {Author Correction: Genetic meta-analysis of diagnosed
             Alzheimer's disease identifies new risk loci and implicates
             Aβ, tau, immunity and lipid processing.},
   Journal = {Nat Genet},
   Volume = {51},
   Number = {9},
   Pages = {1423-1424},
   Year = {2019},
   Month = {September},
   url = {http://dx.doi.org/10.1038/s41588-019-0495-7},
   Abstract = {An amendment to this paper has been published and can be
             accessed via a link at the top of the paper.},
   Doi = {10.1038/s41588-019-0495-7},
   Key = {fds371160}
}

@article{fds345607,
   Author = {Mathew, JP and Welsh-Bohmer, KA and Newman, MF},
   Title = {Nomenclature for Perioperative Cognitive Disorders:
             Comment.},
   Journal = {Anesthesiology},
   Volume = {131},
   Number = {2},
   Pages = {443-444},
   Year = {2019},
   Month = {August},
   url = {http://dx.doi.org/10.1097/ALN.0000000000002831},
   Doi = {10.1097/ALN.0000000000002831},
   Key = {fds345607}
}

@article{fds371161,
   Author = {Ma, Y and Jun, GR and Chung, J and Zhang, X and Kunkle, BW and Naj, AC and White, CC and Bennett, DA and De Jager and PL and Alzheimer’s Disease
             Genetics Consortium, and Mayeux, R and Haines, JL and Pericak-Vance,
             MA and Schellenberg, GD and Farrer, LA and Lunetta,
             KL},
   Title = {CpG-related SNPs in the MS4A region have a dose-dependent
             effect on risk of late-onset Alzheimer disease.},
   Journal = {Aging Cell},
   Volume = {18},
   Number = {4},
   Pages = {e12964},
   Year = {2019},
   Month = {August},
   url = {http://dx.doi.org/10.1111/acel.12964},
   Abstract = {CpG-related single nucleotide polymorphisms (CGS) have the
             potential to perturb DNA methylation; however, their effects
             on Alzheimer disease (AD) risk have not been evaluated
             systematically. We conducted a genome-wide association study
             using a sliding-window approach to measure the combined
             effects of CGSes on AD risk in a discovery sample of 24
             European ancestry cohorts (12,181 cases, 12,601 controls)
             from the Alzheimer's Disease Genetics Consortium (ADGC) and
             replication sample of seven European ancestry cohorts (7,554
             cases, 27,382 controls) from the International Genomics of
             Alzheimer's Project (IGAP). The potential functional
             relevance of significant associations was evaluated by
             analysis of methylation and expression levels in brain
             tissue of the Religious Orders Study and the Rush Memory and
             Aging Project (ROSMAP), and in whole blood of Framingham
             Heart Study participants (FHS). Genome-wide significant
             (p < 5 × 10-8 ) associations were identified with 171
             1.0 kb-length windows spanning 932 kb in the APOE region
             (top p < 2.2 × 10-308 ), five windows at BIN1 (top
             p = 1.3 × 10-13 ), two windows at MS4A6A (top
             p = 2.7 × 10-10 ), two windows near MS4A4A (top
             p = 6.4 × 10-10 ), and one window at PICALM
             (p = 6.3 × 10-9 ). The total number of CGS-derived CpG
             dinucleotides in the window near MS4A4A was associated with
             AD risk (p = 2.67 × 10-10 ), brain DNA methylation
             (p = 2.15 × 10-10 ), and gene expression in brain
             (p = 0.03) and blood (p = 2.53 × 10-4 ). Pathway
             analysis of the genes responsive to changes in the
             methylation quantitative trait locus signal at MS4A4A
             (cg14750746) showed an enrichment of methyltransferase
             functions. We confirm the importance of CGS in AD and the
             potential for creating a functional CpG dosage-derived
             genetic score to predict AD risk.},
   Doi = {10.1111/acel.12964},
   Key = {fds371161}
}

@article{fds343521,
   Author = {Klinger, RY and Cooter, M and Bisanar, T and Terrando, N and Berger, M and Podgoreanu, MV and Stafford-Smith, M and Newman, MF and Mathew, JP and Neurologic Outcomes Research Group of the Duke Heart
             Center},
   Title = {Intravenous Lidocaine Does Not Improve Neurologic Outcomes
             after Cardiac Surgery: A Randomized Controlled
             Trial.},
   Journal = {Anesthesiology},
   Volume = {130},
   Number = {6},
   Pages = {958-970},
   Year = {2019},
   Month = {June},
   url = {http://dx.doi.org/10.1097/ALN.0000000000002668},
   Abstract = {BACKGROUND: Cognitive decline after cardiac surgery occurs
             frequently and persists in a significant proportion of
             patients. Preclinical studies and human trials suggest that
             intravenous lidocaine may confer protection in the setting
             of neurologic injury. It was hypothesized that lidocaine
             administration would reduce cognitive decline after cardiac
             surgery compared to placebo. METHODS: After institutional
             review board approval, 478 patients undergoing cardiac
             surgery were enrolled into this multicenter, prospective,
             randomized, double-blinded, placebo-controlled, parallel
             group trial. Subjects were randomized to lidocaine 1 mg/kg
             bolus after the induction of anesthesia followed by a
             continuous infusion (48 μg · kg · min for the first hour,
             24 μg · kg · min for the second hour, and 10 μg · kg ·
             min for the next 46 h) or saline with identical volume and
             rate changes to preserve blinding. Cognitive function was
             assessed preoperatively and at 6 weeks and 1 yr
             postoperatively using a standard neurocognitive test
             battery. The primary outcome was change in cognitive
             function between baseline and 6 weeks postoperatively,
             adjusting for age, years of education, baseline cognition,
             race, and procedure type. RESULTS: Among the 420 allocated
             subjects who returned for 6-week follow-up (lidocaine: N =
             211; placebo: N = 209), there was no difference in the
             continuous cognitive score change (adjusted mean difference
             [95% CI], 0.02 (-0.05, 0.08); P = 0.626). Cognitive deficit
             (greater than 1 SD decline in at least one cognitive domain)
             at 6 weeks occurred in 41% (87 of 211) in the lidocaine
             group versus 40% (83 of 209) in the placebo group (adjusted
             odds ratio [95% CI], 0.94 [0.63, 1.41]; P = 0.766). There
             were no differences in any quality of life outcomes between
             treatment groups. At the 1-yr follow-up, there continued to
             be no difference in cognitive score change, cognitive
             deficit, or quality of life. CONCLUSIONS: Intravenous
             lidocaine administered during and after cardiac surgery did
             not reduce postoperative cognitive decline at 6
             weeks.},
   Doi = {10.1097/ALN.0000000000002668},
   Key = {fds343521}
}

@article{fds371162,
   Author = {Goudey, B and Fung, BJ and Schieber, C and Faux, NG and Alzheimer’s
             Disease Metabolomics Consortium, and Alzheimer’s Disease
             Neuroimaging Initiative},
   Title = {A blood-based signature of cerebrospinal fluid Aβ1-42
             status.},
   Journal = {Sci Rep},
   Volume = {9},
   Number = {1},
   Pages = {4163},
   Year = {2019},
   Month = {March},
   url = {http://dx.doi.org/10.1038/s41598-018-37149-7},
   Abstract = {It is increasingly recognized that Alzheimer's disease (AD)
             exists before dementia is present and that shifts in amyloid
             beta occur long before clinical symptoms can be detected.
             Early detection of these molecular changes is a key aspect
             for the success of interventions aimed at slowing down rates
             of cognitive decline. Recent evidence indicates that of the
             two established methods for measuring amyloid, a decrease in
             cerebrospinal fluid (CSF) amyloid β1-42 (Aβ1-42) may be an
             earlier indicator of Alzheimer's disease risk than measures
             of amyloid obtained from Positron Emission Tomography (PET).
             However, CSF collection is highly invasive and expensive. In
             contrast, blood collection is routinely performed, minimally
             invasive and cheap. In this work, we develop a blood-based
             signature that can provide a cheap and minimally invasive
             estimation of an individual's CSF amyloid status using a
             machine learning approach. We show that a Random Forest
             model derived from plasma analytes can accurately predict
             subjects as having abnormal (low) CSF Aβ1-42 levels
             indicative of AD risk (0.84 AUC, 0.78 sensitivity, and 0.73
             specificity). Refinement of the modeling indicates that only
             APOEε4 carrier status and four plasma analytes (CGA,
             Aβ1-42, Eotaxin 3, APOE) are required to achieve a high
             level of accuracy. Furthermore, we show across an
             independent validation cohort that individuals with
             predicted abnormal CSF Aβ1-42 levels transitioned to an AD
             diagnosis over 120 months significantly faster than those
             with predicted normal CSF Aβ1-42 levels and that the
             resulting model also validates reasonably across PET Aβ1-42
             status (0.78 AUC). This is the first study to show that a
             machine learning approach, using plasma protein levels, age
             and APOEε4 carrier status, is able to predict CSF Aβ1-42
             status, the earliest risk indicator for AD, with high
             accuracy.},
   Doi = {10.1038/s41598-018-37149-7},
   Key = {fds371162}
}

@article{fds371163,
   Author = {Kunkle, BW and Grenier-Boley, B and Sims, R and Bis, JC and Damotte, V and Naj, AC and Boland, A and Vronskaya, M and van der Lee, SJ and Amlie-Wolf, A and Bellenguez, C and Frizatti, A and Chouraki, V and Martin, ER and Sleegers, K and Badarinarayan, N and Jakobsdottir, J and Hamilton-Nelson, KL and Moreno-Grau, S and Olaso, R and Raybould, R and Chen, Y and Kuzma, AB and Hiltunen, M and Morgan, T and Ahmad, S and Vardarajan, BN and Epelbaum, J and Hoffmann, P and Boada, M and Beecham,
             GW and Garnier, J-G and Harold, D and Fitzpatrick, AL and Valladares, O and Moutet, M-L and Gerrish, A and Smith, AV and Qu, L and Bacq, D and Denning,
             N and Jian, X and Zhao, Y and Del Zompo and M and Fox, NC and Choi, S-H and Mateo, I and Hughes, JT and Adams, HH and Malamon, J and Sanchez-Garcia,
             F and Patel, Y and Brody, JA and Dombroski, BA and Naranjo, MCD and Daniilidou, M and Eiriksdottir, G and Mukherjee, S and Wallon, D and Uphill, J and Aspelund, T and Cantwell, LB and Garzia, F and Galimberti,
             D and Hofer, E and Butkiewicz, M and Fin, B and Scarpini, E and Sarnowski,
             C and Bush, WS and Meslage, S and Kornhuber, J and White, CC and Song, Y and Barber, RC and Engelborghs, S and Sordon, S and Voijnovic, D and Adams,
             PM and Vandenberghe, R and Mayhaus, M and Cupples, LA and Albert, MS and De
             Deyn, PP and Gu, W and Himali, JJ and Beekly, D and Squassina, A and Hartmann, AM and Orellana, A and Blacker, D and Rodriguez-Rodriguez,
             E and Lovestone, S and Garcia, ME and Doody, RS and Munoz-Fernadez, C and Sussams, R and Lin, H and Fairchild, TJ and Benito, YA and Holmes, C and Karamujić-Čomić, H and Frosch, MP and Thonberg, H and Maier, W and Roshchupkin, G and Ghetti, B and Giedraitis, V and Kawalia, A and Li, S and Huebinger, RM and Kilander, L and Moebus, S and Hernández, I and Kamboh, MI and Brundin, R and Turton, J and Yang, Q and Katz, MJ and Concari, L and Lord, J and Beiser, AS and Keene, CD and Helisalmi, S and Kloszewska, I and Kukull, WA and Koivisto, AM and Lynch, A and Tarraga,
             L and Larson, EB and Haapasalo, A and Lawlor, B and Mosley, TH and Lipton,
             RB and Solfrizzi, V and Gill, M and Longstreth, WT and Montine, TJ and Frisardi, V and Diez-Fairen, M and Rivadeneira, F and Petersen, RC and Deramecourt, V and Alvarez, I and Salani, F and Ciaramella, A and Boerwinkle, E and Reiman, EM and Fievet, N and Rotter, JI and Reisch,
             JS and Hanon, O and Cupidi, C and Andre Uitterlinden and AG and Royall, DR and Dufouil, C and Maletta, RG and de Rojas, I and Sano, M and Brice, A and Cecchetti, R and George-Hyslop, PS and Ritchie, K and Tsolaki, M and Tsuang, DW and Dubois, B and Craig, D and Wu, C-K and Soininen, H and Avramidou, D and Albin, RL and Fratiglioni, L and Germanou, A and Apostolova, LG and Keller, L and Koutroumani, M and Arnold, SE and Panza, F and Gkatzima, O and Asthana, S and Hannequin, D and Whitehead,
             P and Atwood, CS and Caffarra, P and Hampel, H and Quintela, I and Carracedo, Á and Lannfelt, L and Rubinsztein, DC and Barnes, LL and Pasquier, F and Frölich, L and Barral, S and McGuinness, B and Beach,
             TG and Johnston, JA and Becker, JT and Passmore, P and Bigio, EH and Schott, JM and Bird, TD and Warren, JD and Boeve, BF and Lupton, MK and Bowen, JD and Proitsi, P and Boxer, A and Powell, JF and Burke, JR and Kauwe, JSK and Burns, JM and Mancuso, M and Buxbaum, JD and Bonuccelli,
             U and Cairns, NJ and McQuillin, A and Cao, C and Livingston, G and Carlson,
             CS and Bass, NJ and Carlsson, CM and Hardy, J and Carney, RM and Bras, J and Carrasquillo, MM and Guerreiro, R and Allen, M and Chui, HC and Fisher,
             E and Masullo, C and Crocco, EA and DeCarli, C and Bisceglio, G and Dick,
             M and Ma, L and Duara, R and Graff-Radford, NR and Evans, DA and Hodges, A and Faber, KM and Scherer, M and Fallon, KB and Riemenschneider, M and Fardo, DW and Heun, R and Farlow, MR and Kölsch, H and Ferris, S and Leber, M and Foroud, TM and Heuser, I and Galasko, DR and Giegling, I and Gearing, M and Hüll, M and Geschwind, DH and Gilbert, JR and Morris, J and Green, RC and Mayo, K and Growdon, JH and Feulner, T and Hamilton, RL and Harrell, LE and Drichel, D and Honig, LS and Cushion, TD and Huentelman,
             MJ and Hollingworth, P and Hulette, CM and Hyman, BT and Marshall, R and Jarvik, GP and Meggy, A and Abner, E and Menzies, GE and Jin, L-W and Leonenko, G and Real, LM and Jun, GR and Baldwin, CT and Grozeva, D and Karydas, A and Russo, G and Kaye, JA and Kim, R and Jessen, F and Kowall,
             NW and Vellas, B and Kramer, JH and Vardy, E and LaFerla, FM and Jöckel,
             K-H and Lah, JJ and Dichgans, M and Leverenz, JB and Mann, D and Levey, AI and Pickering-Brown, S and Lieberman, AP and Klopp, N and Lunetta, KL and Wichmann, H-E and Lyketsos, CG and Morgan, K and Marson, DC and Brown,
             K and Martiniuk, F and Medway, C and Mash, DC and Nöthen, MM and Masliah,
             E and Hooper, NM and McCormick, WC and Daniele, A and McCurry, SM and Bayer, A and McDavid, AN and Gallacher, J and McKee, AC and van den
             Bussche, H and Mesulam, M and Brayne, C and Miller, BL and Riedel-Heller, S and Miller, CA and Miller, JW and Al-Chalabi, A and Morris, JC and Shaw, CE and Myers, AJ and Wiltfang, J and O'Bryant, S and Olichney, JM and Alvarez, V and Parisi, JE and Singleton, AB and Paulson, HL and Collinge, J and Perry, WR and Mead, S and Peskind, E and Cribbs, DH and Rossor, M and Pierce, A and Ryan, NS and Poon, WW and Nacmias, B and Potter, H and Sorbi, S and Quinn, JF and Sacchinelli, E and Raj, A and Spalletta, G and Raskind, M and Caltagirone, C and Bossù, P and Orfei, MD and Reisberg, B and Clarke, R and Reitz, C and Smith, AD and Ringman, JM and Warden, D and Roberson, ED and Wilcock, G and Rogaeva,
             E and Bruni, AC and Rosen, HJ and Gallo, M and Rosenberg, RN and Ben-Shlomo, Y and Sager, MA and Mecocci, P and Saykin, AJ and Pastor, P and Cuccaro, ML and Vance, JM and Schneider, JA and Schneider, LS and Slifer, S and Seeley, WW and Smith, AG and Sonnen, JA and Spina, S and Stern, RA and Swerdlow, RH and Tang, M and Tanzi, RE and Trojanowski,
             JQ and Troncoso, JC and Van Deerlin and VM and Van Eldik and LJ and Vinters,
             HV and Vonsattel, JP and Weintraub, S and Welsh-Bohmer, KA and Wilhelmsen, KC and Williamson, J and Wingo, TS and Woltjer, RL and Wright, CB and Yu, C-E and Yu, L and Saba, Y and Pilotto, A and Bullido,
             MJ and Peters, O and Crane, PK and Bennett, D and Bosco, P and Coto, E and Boccardi, V and De Jager and PL and Lleo, A and Warner, N and Lopez, OL and Ingelsson, M and Deloukas, P and Cruchaga, C and Graff, C and Gwilliam,
             R and Fornage, M and Goate, AM and Sanchez-Juan, P and Kehoe, PG and Amin,
             N and Ertekin-Taner, N and Berr, C and Debette, S and Love, S and Launer,
             LJ and Younkin, SG and Dartigues, J-F and Corcoran, C and Ikram, MA and Dickson, DW and Nicolas, G and Campion, D and Tschanz, J and Schmidt, H and Hakonarson, H and Clarimon, J and Munger, R and Schmidt, R and Farrer,
             LA and Van Broeckhoven and C and C O'Donovan and M and DeStefano, AL and Jones, L and Haines, JL and Deleuze, J-F and Owen, MJ and Gudnason, V and Mayeux, R and Escott-Price, V and Psaty, BM and Ramirez, A and Wang,
             L-S and Ruiz, A and van Duijn, CM and Holmans, PA and Seshadri, S and Williams, J and Amouyel, P and Schellenberg, GD and Lambert, J-C and Pericak-Vance, MA and Alzheimer Disease Genetics Consortium
             (ADGC),, and European Alzheimer’s Disease Initiative
             (EADI),, and Cohorts for Heart and Aging Research in Genomic
             Epidemiology Consortium (CHARGE),, and Genetic and Environmental Risk in AD/Defining Genetic and Polygenic and Environmental Risk for Alzheimer’s Disease Consortium
             (GERAD/PERADES)},
   Title = {Genetic meta-analysis of diagnosed Alzheimer's disease
             identifies new risk loci and implicates Aβ, tau, immunity
             and lipid processing.},
   Journal = {Nat Genet},
   Volume = {51},
   Number = {3},
   Pages = {414-430},
   Year = {2019},
   Month = {March},
   url = {http://dx.doi.org/10.1038/s41588-019-0358-2},
   Abstract = {Risk for late-onset Alzheimer's disease (LOAD), the most
             prevalent dementia, is partially driven by genetics. To
             identify LOAD risk loci, we performed a large genome-wide
             association meta-analysis of clinically diagnosed LOAD
             (94,437 individuals). We confirm 20 previous LOAD risk loci
             and identify five new genome-wide loci (IQCK, ACE, ADAM10,
             ADAMTS1, and WWOX), two of which (ADAM10, ACE) were
             identified in a recent genome-wide association
             (GWAS)-by-familial-proxy of Alzheimer's or dementia.
             Fine-mapping of the human leukocyte antigen (HLA) region
             confirms the neurological and immune-mediated disease
             haplotype HLA-DR15 as a risk factor for LOAD. Pathway
             analysis implicates immunity, lipid metabolism, tau binding
             proteins, and amyloid precursor protein (APP) metabolism,
             showing that genetic variants affecting APP and Aβ
             processing are associated not only with early-onset
             autosomal dominant Alzheimer's disease but also with LOAD.
             Analyses of risk genes and pathways show enrichment for rare
             variants (P = 1.32 × 10-7), indicating that
             additional rare variants remain to be identified. We also
             identify important genetic correlations between LOAD and
             traits such as family history of dementia and
             education.},
   Doi = {10.1038/s41588-019-0358-2},
   Key = {fds371163}
}

@article{fds371164,
   Author = {Hu, Y and Li, M and Lu, Q and Weng, H and Wang, J and Zekavat, SM and Yu, Z and Li, B and Gu, J and Muchnik, S and Shi, Y and Kunkle, BW and Mukherjee, S and Natarajan, P and Naj, A and Kuzma, A and Zhao, Y and Crane, PK and Alzheimer’s Disease Genetics Consortium,, and Lu, H and Zhao,
             H},
   Title = {A statistical framework for cross-tissue transcriptome-wide
             association analysis.},
   Journal = {Nat Genet},
   Volume = {51},
   Number = {3},
   Pages = {568-576},
   Year = {2019},
   Month = {March},
   url = {http://dx.doi.org/10.1038/s41588-019-0345-7},
   Abstract = {Transcriptome-wide association analysis is a powerful
             approach to studying the genetic architecture of complex
             traits. A key component of this approach is to build a model
             to impute gene expression levels from genotypes by using
             samples with matched genotypes and gene expression data in a
             given tissue. However, it is challenging to develop robust
             and accurate imputation models with a limited sample size
             for any single tissue. Here, we first introduce a multi-task
             learning method to jointly impute gene expression in 44
             human tissues. Compared with single-tissue methods, our
             approach achieved an average of 39% improvement in
             imputation accuracy and generated effective imputation
             models for an average of 120% more genes. We describe a
             summary-statistic-based testing framework that combines
             multiple single-tissue associations into a powerful metric
             to quantify the overall gene-trait association. We applied
             our method, called UTMOST (unified test for molecular
             signatures), to multiple genome-wide-association results and
             demonstrate its advantages over single-tissue
             strategies.},
   Doi = {10.1038/s41588-019-0345-7},
   Key = {fds371164}
}

@article{fds341018,
   Author = {Blumenthal, JA and Smith, PJ and Mabe, S and Hinderliter, A and Lin,
             P-H and Liao, L and Welsh-Bohmer, KA and Browndyke, JN and Kraus, WE and Doraiswamy, PM and Burke, JR and Sherwood, A},
   Title = {Lifestyle and neurocognition in older adults with cognitive
             impairments: A randomized trial.},
   Journal = {Neurology},
   Volume = {92},
   Number = {3},
   Pages = {e212-e223},
   Year = {2019},
   Month = {January},
   url = {http://dx.doi.org/10.1212/WNL.0000000000006784},
   Abstract = {OBJECTIVE: To determine the independent and additive effects
             of aerobic exercise (AE) and the Dietary Approaches to Stop
             Hypertension (DASH) diet on executive functioning in adults
             with cognitive impairments with no dementia (CIND) and risk
             factors for cardiovascular disease (CVD). METHODS: A 2-by-2
             factorial (exercise/no exercise and DASH diet/no DASH diet)
             randomized clinical trial was conducted in 160 sedentary men
             and women (age >55 years) with CIND and CVD risk factors.
             Participants were randomly assigned to 6 months of AE, DASH
             diet nutritional counseling, a combination of both AE and
             DASH, or health education (HE). The primary endpoint was a
             prespecified composite measure of executive function;
             secondary outcomes included measures of language/verbal
             fluency, memory, and ratings on the modified Clinical
             Dementia Rating Scale. RESULTS: Participants who engaged in
             AE (d = 0.32, p = 0.046) but not those who consumed the DASH
             diet (d = 0.30, p = 0.059) demonstrated significant
             improvements in the executive function domain. The largest
             improvements were observed for participants randomized to
             the combined AE and DASH diet group (d = 0.40, p = 0.012)
             compared to those receiving HE. Greater aerobic fitness (b =
             2.3, p = 0.049), reduced CVD risk (b = 2.6, p = 0.042), and
             reduced sodium intake (b = 0.18, p = 0.024) were associated
             with improvements in executive function. There were no
             significant improvements in the memory or language/verbal
             fluency domains. CONCLUSIONS: These preliminary findings
             show that AE promotes improved executive functioning in
             adults at risk for cognitive decline. CLINICALTRIALSGOV
             IDENTIFIER: NCT01573546. CLASSIFICATION OF EVIDENCE: This
             study provides Class I evidence that for adults with CIND,
             AE but not the DASH diet significantly improves executive
             functioning.},
   Doi = {10.1212/WNL.0000000000006784},
   Key = {fds341018}
}

@article{fds341567,
   Author = {Tsui-Caldwell, YHW and Farrer, TJ and McDonnell, Z and Christensen,
             Z and Finuf, C and Bigler, ED and Tschanz, JT and Norton, MC and Welsh-Bohmer, KA},
   Title = {MRI Clinical Ratings and Cognitive Function in a
             Cross-Sectional Population Study of Dementia: The Cache
             County Memory Study.},
   Journal = {J Prev Alzheimers Dis},
   Volume = {6},
   Number = {2},
   Pages = {100-107},
   Year = {2019},
   url = {http://dx.doi.org/10.14283/jpad.2019.1},
   Abstract = {BACKGROUND: White matter integrity in aging populations is
             associated with increased risk of cognitive decline,
             dementia diagnosis, and mortality. Population-based data can
             elucidate this association. OBJECTIVES: To examine the
             association between white matter integrity, as measured by a
             clinical rating scale of hyperintensities, and mental status
             in older adults including advanced aging. DESIGN: Scheltens
             Ratings Scale was used to qualitatively assess white matter
             (WM) hyperintensities in participants of the Cache County
             Memory Study (CCMS), an epidemiological study of Alzheimer's
             disease in an exceptionally long-lived population. Further,
             the relation between Mini-Mental State Exam (MMSE) and WM
             hyperintensities were explored. METHOD: Participants
             consisted of 415 individuals with dementia and 22 healthy
             controls. RESULTS: CCMS participants, including healthy
             controls, had high levels of WM pathology as measured by
             Scheltens Ratings Scale score. While age did not
             significantly relate to WM pathology, higher Scheltens
             Ratings Scale scores were associated with lower MMSE
             findings (correlation between -0.14 and -0.22; p < .05).
             CONCLUSIONS: WM pathology was common in this county-wide
             population sample of those ranging in age from 65 to 106.
             Increased WM burden was found to be significantly associated
             with decreased overall MMSE performance.},
   Doi = {10.14283/jpad.2019.1},
   Key = {fds341567}
}

@article{fds344656,
   Author = {Knodt, AR and Burke, JR and Welsh-Bohmer, KA and Plassman, BL and Burns,
             DK and Brannan, SK and Kukulka, M and Wu, J and Hariri,
             AR},
   Title = {Effects of pioglitazone on mnemonic hippocampal function: A
             blood oxygen level-dependent functional magnetic resonance
             imaging study in elderly adults.},
   Journal = {Alzheimers Dement (N Y)},
   Volume = {5},
   Pages = {254-263},
   Year = {2019},
   url = {http://dx.doi.org/10.1016/j.trci.2019.05.004},
   Abstract = {INTRODUCTION: Mitochondrial dysfunction is implicated in the
             pathophysiology of Alzheimer's disease (AD). Accordingly,
             drugs that positively influence mitochondrial function are
             being evaluated in delay-of-onset clinical trials with
             at-risk individuals. Such ongoing clinical research can be
             advanced by developing a better understanding of how these
             drugs affect intermediate brain phenotypes associated with
             both AD risk and pathophysiology. METHODS: Using a
             randomized, parallel-group, placebo-controlled design in 55
             healthy elderly volunteers, we explored the effects of oral,
             low-dose pioglitazone, a thiazolidinedione with
             promitochondrial effects, on hippocampal activity measured
             with functional magnetic resonance imaging during the
             encoding of novel face-name pairs. RESULTS: Compared with
             placebo, 0.6 mg of pioglitazone (but not 2.1 mg, 3.9 mg,
             or 6.0 mg) administered daily for 14 days was associated
             with significant increases in right hippocampal activation
             during encoding of novel face-name pairs at day 7 and day
             14, relative to baseline. DISCUSSION: Our exploratory
             analyses suggest that low-dose pioglitazone has measurable
             effects on mnemonic brain function associated with AD risk
             and pathophysiology.},
   Doi = {10.1016/j.trci.2019.05.004},
   Key = {fds344656}
}

@article{fds346727,
   Author = {Smith, PJ and Mabe, S and Sherwood, A and Babyak, MA and Doraiswamy, PM and Welsh-Bohmer, KA and Kraus, W and Burke, J and Hinderliter, A and Blumenthal, JA},
   Title = {Association Between Insulin Resistance, Plasma Leptin, and
             Neurocognition in Vascular Cognitive Impairment.},
   Journal = {J Alzheimers Dis},
   Volume = {71},
   Number = {3},
   Pages = {921-929},
   Year = {2019},
   url = {http://dx.doi.org/10.3233/JAD-190569},
   Abstract = {BACKGROUND: Greater body weight has been associated
             impairments in neurocognition and greater dementia risk,
             although the mechanisms linking weight and neurocognition
             have yet to be adequately delineated. OBJECTIVE: To examine
             metabolic mechanisms underlying the association between
             obesity and neurocognition. METHODS: We conducted a
             secondary analysis of weight, neurocognition, and the
             potentially mediating role of metabolic and inflammatory
             biomarkers among 160 participants from the ENLIGHTEN trial
             of vascular cognitive impairment, no dementia (CIND).
             Neurocognition was assessed using a 45-minute assessment
             battery assessing Executive Function, Verbal and Visual
             Memory. We considered three metabolic biomarkers: insulin
             resistance (homeostatic model assessment [HOMA-IR]), plasma
             leptin, and insulin-like growth factor (IGF-1). Inflammation
             was assessed using C-reactive protein. Multiple regression
             analyses were used. RESULTS: Participants included 160
             sedentary older adults with CIND. Participants tended to be
             overweight or obese (mean BMI = 32.5 [SD = 4.8]).
             Women exhibited higher BMI (p = 0.043), CRP
             (p < 0.001), and leptin (p < 0.001) compared with
             men. Higher BMI levels were associated with worse
             performance on measures of Executive Function (β= -0.16,
             p = 0.024) and Verbal Memory (β= -0.16,
             p = 0.030), but not Visual Memory (β= 0.05,
             p = 0.500). Worse metabolic biomarker profiles also were
             associated with lower Executive Function (β= -0.12,
             p = 0.050). Mediation analyses suggested leptin was a
             plausible candidate as a mediator between BMI and Executive
             Function. CONCLUSIONS: In overweight and obese adults with
             vascular CIND, the association between greater weight and
             poorer executive function may be mediated by higher leptin
             resistance.},
   Doi = {10.3233/JAD-190569},
   Key = {fds346727}
}

@article{fds346902,
   Author = {Burns, DK and Chiang, C and Welsh-Bohmer, KA and Brannan, SK and Culp,
             M and O'Neil, J and Runyan, G and Harrigan, P and Plassman, BL and Lutz, M and Lai, E and Haneline, S and Yarnall, D and Yarbrough, D and Metz, C and Ponduru, S and Sundseth, S and Saunders, AM},
   Title = {The TOMMORROW study: Design of an Alzheimer's disease
             delay-of-onset clinical trial.},
   Journal = {Alzheimers Dement (N Y)},
   Volume = {5},
   Pages = {661-670},
   Year = {2019},
   url = {http://dx.doi.org/10.1016/j.trci.2019.09.010},
   Abstract = {INTRODUCTION: Alzheimer's disease (AD) is a continuum with
             neuropathologies manifesting years before clinical symptoms;
             thus, AD research is attempting to identify more
             disease-modifying approaches to test treatments administered
             before full disease expression. Designing such trials in
             cognitively normal elderly individuals poses unique
             challenges. METHODS: The TOMMORROW study was a phase 3
             double-blind, parallel-group study designed to support
             qualification of a novel genetic biomarker risk assignment
             algorithm (BRAA) and to assess efficacy and safety of
             low-dose pioglitazone to delay onset of mild cognitive
             impairment due to AD. Eligible participants were stratified
             based on the BRAA (using TOMM40 rs 10524523 genotype,
             Apolipoprotein E genotype, and age), with high-risk
             individuals receiving low-dose pioglitazone or placebo and
             low-risk individuals receiving placebo. The primary endpoint
             was time to the event of mild cognitive impairment due to
             AD. The primary objectives were to compare the primary
             endpoint between high- and low-risk placebo groups (for BRAA
             qualification) and between high-risk pioglitazone and
             high-risk placebo groups (for pioglitazone efficacy).
             Approximately 300 individuals were also asked to participate
             in a volumetric magnetic resonance imaging substudy at
             selected sites. RESULTS: The focus of this paper is on the
             design of the study; study results will be presented in a
             separate paper. DISCUSSION: The design of the TOMMORROW
             study addressed many key challenges to conducting a
             dual-objective phase 3 pivotal AD clinical trial in
             presymptomatic individuals. Experiences from planning and
             executing the TOMMORROW study may benefit future AD
             prevention/delay-of-onset trials.},
   Doi = {10.1016/j.trci.2019.09.010},
   Key = {fds346902}
}

@article{fds351449,
   Author = {Gauthier, S and Alam, J and Fillit, H and Iwatsubo, T and Liu-Seifert,
             H and Sabbagh, M and Salloway, S and Sampaio, C and Sims, JR and Sperling,
             B and Sperling, R and Welsh-Bohmer, KA and Touchon, J and Vellas, B and Aisen, P},
   Title = {Combination Therapy for Alzheimer's Disease: Perspectives of
             the EU/US CTAD Task Force.},
   Journal = {J Prev Alzheimers Dis},
   Volume = {6},
   Number = {3},
   Pages = {164-168},
   Year = {2019},
   url = {http://dx.doi.org/10.14283/jpad.2019.12},
   Abstract = {Combination therapy is expected to play an important role
             for the treatment of Alzheimer's disease (AD). In October
             2018, the European Union-North American Clinical Trials in
             Alzheimer's Disease Task Force (EU/US CTAD Task Force) met
             to discuss scientific, regulatory, and logistical challenges
             to the development of combination therapy for AD and current
             efforts to address these challenges. Task Force members
             unanimously agreed that successful treatment of AD will
             likely require combination therapy approaches that target
             multiple mechanisms and pathways. They further agreed on the
             need for global collaboration and sharing of data and
             resources to accelerate development of such
             approaches.},
   Doi = {10.14283/jpad.2019.12},
   Key = {fds351449}
}

@article{fds340273,
   Author = {Deming, Y and Dumitrescu, L and Barnes, LL and Thambisetty, M and Kunkle, B and Gifford, KA and Bush, WS and Chibnik, LB and Mukherjee, S and De Jager and PL and Kukull, W and Huentelman, M and Crane, PK and Resnick,
             SM and Keene, CD and Montine, TJ and Schellenberg, GD and Haines, JL and Zetterberg, H and Blennow, K and Larson, EB and Johnson, SC and Albert,
             M and Moghekar, A and Del Aguila and JL and Fernandez, MV and Budde, J and Hassenstab, J and Fagan, AM and Riemenschneider, M and Petersen, RC and Minthon, L and Chao, MJ and Van Deerlin and VM and Lee, VM-Y and Shaw, LM and Trojanowski, JQ and Peskind, ER and Li, G and Davis, LK and Sealock, JM and Cox, NJ and Alzheimer’s Disease Neuroimaging Initiative
             (ADNI), and Alzheimer Disease Genetics Consortium (ADGC), and Goate, AM and Bennett, DA and Schneider, JA and Jefferson, AL and Cruchaga, C and Hohman, TJ},
   Title = {Sex-specific genetic predictors of Alzheimer's disease
             biomarkers.},
   Journal = {Acta Neuropathol},
   Volume = {136},
   Number = {6},
   Pages = {857-872},
   Publisher = {Springer Nature},
   Year = {2018},
   Month = {December},
   url = {http://dx.doi.org/10.1007/s00401-018-1881-4},
   Abstract = {Cerebrospinal fluid (CSF) levels of amyloid-β 42 (Aβ42)
             and tau have been evaluated as endophenotypes in Alzheimer's
             disease (AD) genetic studies. Although there are sex
             differences in AD risk, sex differences have not been
             evaluated in genetic studies of AD endophenotypes. We
             performed sex-stratified and sex interaction genetic
             analyses of CSF biomarkers to identify sex-specific
             associations. Data came from a previous genome-wide
             association study (GWAS) of CSF Aβ42 and tau (1527 males,
             1509 females). We evaluated sex interactions at previous
             loci, performed sex-stratified GWAS to identify sex-specific
             associations, and evaluated sex interactions at sex-specific
             GWAS loci. We then evaluated sex-specific associations
             between prefrontal cortex (PFC) gene expression at relevant
             loci and autopsy measures of plaques and tangles using data
             from the Religious Orders Study and Rush Memory and Aging
             Project. In Aβ42, we observed sex interactions at one
             previous and one novel locus: rs316341 within SERPINB1
             (p = 0.04) and rs13115400 near LINC00290
             (p = 0.002). These loci showed stronger associations
             among females (β = - 0.03, p = 4.25 × 10-8;
             β = 0.03, p = 3.97 × 10-8) than males
             (β = - 0.02, p = 0.009; β = 0.01,
             p = 0.20). Higher levels of expression of SERPINB1,
             SERPINB6, and SERPINB9 in PFC was associated with higher
             levels of amyloidosis among females (corrected p
             values < 0.02) but not males (p > 0.38). In total
             tau, we observed a sex interaction at a previous locus,
             rs1393060 proximal to GMNC (p = 0.004), driven by a
             stronger association among females (β = 0.05,
             p = 4.57 × 10-10) compared to males (β = 0.02,
             p = 0.03). There was also a sex-specific association
             between rs1393060 and tangle density at autopsy
             (pfemale = 0.047; pmale = 0.96), and higher levels
             of expression of two genes within this locus were associated
             with lower tangle density among females (OSTN p = 0.006;
             CLDN16 p = 0.002) but not males (p ≥ 0.32).
             Results suggest a female-specific role for SERPINB1 in
             amyloidosis and for OSTN and CLDN16 in tau pathology.
             Sex-specific genetic analyses may improve understanding of
             AD's genetic architecture.},
   Doi = {10.1007/s00401-018-1881-4},
   Key = {fds340273}
}

@article{fds339515,
   Author = {Hohman, TJ and Dumitrescu, L and Barnes, LL and Thambisetty, M and Beecham, G and Kunkle, B and Gifford, KA and Bush, WS and Chibnik, LB and Mukherjee, S and De Jager and PL and Kukull, W and Crane, PK and Resnick,
             SM and Keene, CD and Montine, TJ and Schellenberg, GD and Haines, JL and Zetterberg, H and Blennow, K and Larson, EB and Johnson, SC and Albert,
             M and Bennett, DA and Schneider, JA and Jefferson, AL and Alzheimer’s
             Disease Genetics Consortium and the Alzheimer’s Disease
             Neuroimaging Initiative},
   Title = {Sex-Specific Association of Apolipoprotein E With
             Cerebrospinal Fluid Levels of Tau.},
   Journal = {JAMA Neurol},
   Volume = {75},
   Number = {8},
   Pages = {989-998},
   Year = {2018},
   Month = {August},
   url = {http://dx.doi.org/10.1001/jamaneurol.2018.0821},
   Abstract = {IMPORTANCE: The strongest genetic risk factor for Alzheimer
             disease (AD), the apolipoprotein E (APOE) gene, has a
             stronger association among women compared with men. Yet
             limited work has evaluated the association between APOE
             alleles and markers of AD neuropathology in a sex-specific
             manner. OBJECTIVE: To evaluate sex differences in the
             association between APOE and markers of AD neuropathology
             measured in cerebrospinal fluid (CSF) during life or in
             brain tissue at autopsy. DESIGN, SETTING, AND PARTICIPANTS:
             This multicohort study selected data from 10 longitudinal
             cohort studies of normal aging and AD. Cohorts had variable
             recruitment criteria and follow-up intervals and included
             population-based and clinic-based samples. Inclusion in our
             analysis required APOE genotype data and either CSF data
             available for analysis. Analyses began on November 6, 2017,
             and were completed on December 20, 2017. MAIN OUTCOMES AND
             MEASURES: Biomarker analyses included levels of β-amyloid
             42, total tau, and phosphorylated tau measured in CSF.
             Autopsy analyses included Consortium to Establish a Registry
             for Alzheimer's Disease staging for neuritic plaques and
             Braak staging for neurofibrillary tangles. RESULTS: Of the
             1798 patients in the CSF biomarker cohort, 862 were women,
             226 had AD, 1690 were white, and the mean (SD) age was 70
             [9] years. Of the 5109 patients in the autopsy cohort, 2813
             were women, 4953 were white, and the mean (SD) age was 84
             (9) years. After correcting for multiple comparisons using
             the Bonferroni procedure, we observed a statistically
             significant interaction between APOE-ε4 and sex on CSF
             total tau (β = 0.41; 95% CI, 0.27-0.55; P < .001)
             and phosphorylated tau (β = 0.24; 95% CI, 0.09-0.38;
             P = .001), whereby APOE showed a stronger association
             among women compared with men. Post hoc analyses suggested
             this sex difference was present in amyloid-positive
             individuals (β = 0.41; 95% CI, 0.20-0.62; P < .001)
             but not among amyloid-negative individuals (β = 0.06;
             95% CI, -0.18 to 0.31; P = .62). We did not observe sex
             differences in the association between APOE and β-amyloid
             42, neuritic plaque burden, or neurofibrillary tangle
             burden. CONCLUSIONS AND RELEVANCE: We provide robust
             evidence of a stronger association between APOE-ε4 and CSF
             tau levels among women compared with men across multiple
             independent data sets. Interestingly, APOE-ε4 is not
             differentially associated with autopsy measures of
             neurofibrillary tangles. Together, the sex difference in the
             association between APOE and CSF measures of tau and the
             lack of a sex difference in the association with
             neurofibrillary tangles at autopsy suggest that APOE may
             modulate risk for neurodegeneration in a sex-specific
             manner, particularly in the presence of amyloidosis.},
   Doi = {10.1001/jamaneurol.2018.0821},
   Key = {fds339515}
}

@article{fds330585,
   Author = {Browndyke, JN and Berger, M and Smith, PJ and Harshbarger, TB and Monge,
             ZA and Panchal, V and Bisanar, TL and Glower, DD and Alexander, JH and Cabeza, R and Welsh-Bohmer, K and Newman, MF and Mathew, JP and Duke
             Neurologic Outcomes Research Group (NORG)},
   Title = {Task-related changes in degree centrality and local
             coherence of the posterior cingulate cortex after major
             cardiac surgery in older adults.},
   Journal = {Hum Brain Mapp},
   Volume = {39},
   Number = {2},
   Pages = {985-1003},
   Year = {2018},
   Month = {February},
   url = {http://dx.doi.org/10.1002/hbm.23898},
   Abstract = {OBJECTIVES: Older adults often display postoperative
             cognitive decline (POCD) after surgery, yet it is unclear to
             what extent functional connectivity (FC) alterations may
             underlie these deficits. We examined for postoperative
             voxel-wise FC changes in response to increased working
             memory load demands in cardiac surgery patients and
             nonsurgical controls. EXPERIMENTAL DESIGN: Older cardiac
             surgery patients (n = 25) completed a verbal N-back
             working memory task during MRI scanning and cognitive
             testing before and 6 weeks after surgery; nonsurgical
             controls with cardiac disease (n = 26) underwent these
             assessments at identical time intervals. We measured
             postoperative changes in degree centrality, the number of
             edges attached to a brain node, and local coherence, the
             temporal homogeneity of regional functional correlations,
             using voxel-wise graph theory-based FC metrics. Group ×
             time differences were evaluated in these FC metrics
             associated with increased N-back working memory load
             (2-back > 1-back), using a two-stage partitioned
             variance, mixed ANCOVA. PRINCIPAL OBSERVATIONS: Cardiac
             surgery patients demonstrated postoperative working memory
             load-related degree centrality increases in the left dorsal
             posterior cingulate cortex (dPCC; p < .001, cluster
             p-FWE < .05). The dPCC also showed a postoperative
             increase in working memory load-associated local coherence
             (p < .001, cluster p-FWE < .05). dPCC degree
             centrality and local coherence increases were inversely
             associated with global cognitive change in surgery patients
             (p < .01), but not in controls. CONCLUSIONS: Cardiac
             surgery patients showed postoperative increases in working
             memory load-associated degree centrality and local coherence
             of the dPCC that were inversely associated with
             postoperative global cognitive outcomes and independent of
             perioperative cerebrovascular damage.},
   Doi = {10.1002/hbm.23898},
   Key = {fds330585}
}

@article{fds339675,
   Author = {Peloso, GM and van der Lee, SJ and Sims, R and Naj, AC and Bellenguez,
             C and Badarinarayan, N and Jakobsdottir, J and Kunkle, BW and Boland, A and Raybould, R and Bis, JC and Martin, ER and Grenier-Boley, B and Heilmann-Heimbach, S and Chouraki, V and Kuzma, AB and Sleegers, K and Vronskaya, M and Ruiz, A and Graham, RR and Olaso, R and Hoffmann, P and Grove, ML and Vardarajan, BN and Hiltunen, M and Nöthen, MM and White,
             CC and Hamilton-Nelson, KL and Epelbaum, J and Maier, W and Choi, SH and Beecham, GW and Dulary, C and Herms, S and Smith, AV and Funk, CC and Derbois, and Forstner, AJ and Ahmad, S and Li, H and Bacq, D and Harold,
             D and Satizabal, CL and Valladares, O and Squassina, A and Thomas, R and Brody, JA and Qu, L and Sánchez-Juan, P and Morgan, T and Wolters, FJ and Zhao, Y and Garcia, FS and Denning, N and Fornage, M and Malamon, J and Naranjo, MCD and Majounie, E and Mosley, TH and Dombroski, B and Wallon,
             D and Lupton, MK and Dupuis, J and Whitehead, P and Fratiglioni, L and Medway, C and Jian, X and Mukherjee, S and Keller, L and Brown, K and Lin,
             H and Cantwell, LB and Panza, F and McGuinness, B and Moreno-Grau, S and Burgess, JD and Solfrizzi, V and Proitsi, P and Adams, HH and Allen, M and Seripa, D and Pastor, P and Cupples, LA and Price, ND and Hannequin, D and Frank-García, A and Levy, D and Chakrabarty, P and Caffarra, P and Giegling, I and Beiser, AS and Giedraitis, V and Hampel, H and Garcia,
             ME and Wang, X and Lannfelt, L and Mecocci, P and Eiriksdottir, G and Crane, PK},
   Title = {Genetically elevated high-density lipoprotein cholesterol
             through the cholesteryl ester transfer protein gene does not
             associate with risk of Alzheimer's disease},
   Journal = {Alzheimer's and Dementia: Diagnosis, Assessment and Disease
             Monitoring},
   Volume = {10},
   Pages = {595-598},
   Publisher = {Elsevier BV},
   Year = {2018},
   Month = {January},
   url = {http://dx.doi.org/10.1016/j.dadm.2018.08.008},
   Abstract = {Introduction: There is conflicting evidence whether
             high-density lipoprotein cholesterol (HDL-C) is a risk
             factor for Alzheimer's disease (AD) and dementia. Genetic
             variation in the cholesteryl ester transfer protein (CETP)
             locus is associated with altered HDL-C. We aimed to assess
             AD risk by genetically predicted HDL-C. Methods: Ten single
             nucleotide polymorphisms within the CETP locus predicting
             HDL-C were applied to the International Genomics of
             Alzheimer's Project (IGAP) exome chip stage 1 results in up
             16,097 late onset AD cases and 18,077 cognitively normal
             elderly controls. We performed instrumental variables
             analysis using inverse variance weighting, weighted median,
             and MR-Egger. Results: Based on 10 single nucleotide
             polymorphisms distinctly predicting HDL-C in the CETP locus,
             we found that HDL-C was not associated with risk of AD (P
             >.7). Discussion: Our study does not support the role of
             HDL-C on risk of AD through HDL-C altered by CETP. This
             study does not rule out other mechanisms by which HDL-C
             affects risk of AD.},
   Doi = {10.1016/j.dadm.2018.08.008},
   Key = {fds339675}
}

@article{fds339320,
   Author = {Atkins, AS and Khan, A and Ulshen, D and Vaughan, A and Balentin, D and Dickerson, H and Liharska, LE and Plassman, B and Welsh-Bohmer, K and Keefe, RSE},
   Title = {Assessment of Instrumental Activities of Daily Living in
             Older Adults with Subjective Cognitive Decline Using the
             Virtual Reality Functional Capacity Assessment Tool
             (VRFCAT).},
   Journal = {J Prev Alzheimers Dis},
   Volume = {5},
   Number = {4},
   Pages = {216-234},
   Year = {2018},
   url = {http://dx.doi.org/10.14283/jpad.2018.28},
   Abstract = {BACKGROUND: Continuing advances in the understanding of
             Alzheimer's disease progression have inspired development of
             disease-modifying therapeutics intended for use in
             preclinical populations. However, identification of
             clinically meaningful cognitive and functional outcomes for
             individuals who are, by definition, asymptomatic remains a
             significant challenge. Clinical trials for prevention and
             early intervention require measures with increased
             sensitivity to subtle deficits in instrumental activities of
             daily living (IADL) that comprise the first functional
             declines in prodromal disease. Validation of potential
             endpoints is required to ensure measure sensitivity and
             reliability in the populations of interest. OBJECTIVES: The
             present research validates use of the Virtual Reality
             Functional Capacity Assessment Tool (VRFCAT) for
             performance-based assessment of IADL functioning in older
             adults (age 55+) with subjective cognitive decline. DESIGN:
             Cross-sectional validation study. SETTING: All participants
             were evaluated on-site at NeuroCog Trials, Durham, NC, USA.
             PARTICIPANTS: Participants included 245 healthy younger
             adults ages 20-54 (131 female), 247 healthy older adults
             ages 55-91 (151 female) and 61 older adults with subjective
             cognitive decline (SCD) ages 56-97 (45 female). MEASURES:
             Virtual Reality Functional Capacity Assessment Tool; Brief
             Assessment of Cognition App; Alzheimer's Disease Cooperative
             Study Prevention Instrument Project - Mail-In Cognitive
             Function Screening Instrument; Alzheimer's Disease
             Cooperative Study Instrumental Activities of Daily Living -
             Prevention Instrument, University of California, San Diego
             Performance-Based Skills Assessment - Validation of
             Intermediate Measures; Montreal Cognitive Assessment; Trail
             Making Test- Part B. RESULTS: Participants with SCD
             performed significantly worse than age-matched normative
             controls on all VRFCAT endpoints, including total completion
             time, errors and forced progressions (p≤0001 for all,
             after Bonferonni correction). Consistent with prior
             findings, both groups performed significantly worse than
             healthy younger adults (age 20-54). Participants with SCD
             also performed significantly worse than controls on
             objective cognitive measures. VRFCAT performance was
             strongly correlated with cognitive performance. In the SCD
             group, VRFCAT performance was strongly correlated with
             cognitive performance across nearly all tests with
             significant correlation coefficients ranging from 0.3 to
             0.7; VRFCAT summary measures all had correlations greater
             than r=0.5 with MoCA performance and BAC App Verbal Memory
             (p<0.01 for all). CONCLUSIONS: Findings suggest the VRFCAT
             provides a sensitive tool for evaluation of IADL functioning
             in individuals with subjective cognitive decline. Strong
             correlations with cognition across groups suggest the VRFCAT
             may be uniquely suited for clinical trials in preclinical
             AD, as well as longitudinal investigations of the
             relationship between cognition and function.},
   Doi = {10.14283/jpad.2018.28},
   Key = {fds339320}
}

@article{fds337115,
   Author = {Romero, HR and Monsch, AU and Hayden, KM and Plassman, BL and Atkins,
             AS and Keefe, RSE and Brewster, S and Chiang, C and O'Neil, J and Runyan,
             G and Atkinson, MJ and Crawford, S and Budur, K and Burns, DK and Welsh-Bohmer, KA},
   Title = {TOMMORROW neuropsychological battery: German language
             validation and normative study.},
   Journal = {Alzheimers Dement (N Y)},
   Volume = {4},
   Pages = {314-323},
   Year = {2018},
   url = {http://dx.doi.org/10.1016/j.trci.2018.06.009},
   Abstract = {INTRODUCTION: Assessment of preclinical Alzheimer's disease
             (AD) requires reliable and validated methods to detect
             subtle cognitive changes. The battery of standardized
             cognitive assessments that is used for diagnostic criteria
             for mild cognitive impairment due to AD in the TOMMORROW
             study have only been fully validated in English-speaking
             countries. We conducted a validation and normative study of
             the German language version of the TOMMORROW
             neuropsychological test battery, which tests episodic
             memory, language, visuospatial ability, executive function,
             and attention. METHODS: German-speaking cognitively healthy
             controls (NCs) and subjects with AD were recruited from a
             memory clinic at a Swiss medical center. Construct validity,
             test-retest, and alternate form reliability were assessed in
             NCs. Criterion and discriminant validities of the cognitive
             measures were tested using logistic regression and
             discriminant analysis. Cross-cultural equivalency of
             performance of the German language tests was compared with
             English language tests. RESULTS: A total of 198 NCs and 25
             subjects with AD (aged 65-88 years) were analyzed. All
             German language tests discriminated NCs from persons with
             AD. Episodic memory tests had the highest potential to
             discriminate with almost twice the predictive power of any
             other domain. Test-retest reliability of the test battery
             was adequate, and alternate form reliability for episodic
             memory tests was supported. For most tests, age was a
             significant predictor of group effect sizes; therefore,
             normative data were stratified by age. Validity and
             reliability results were similar to those in the published
             US cognitive testing literature. DISCUSSION: This study
             establishes the reliability and validity of the German
             language TOMMORROW test battery, which performed similarly
             to the English language tests. Some variations in test
             performance underscore the importance of regional normative
             values. The German language battery and normative data will
             improve the precision of measuring cognition and diagnosing
             incident mild cognitive impairment due to AD in clinical
             settings in German-speaking countries.},
   Doi = {10.1016/j.trci.2018.06.009},
   Key = {fds337115}
}

@article{fds333554,
   Author = {Weintraub, S and Carrillo, MC and Farias, ST and Goldberg, TE and Hendrix, JA and Jaeger, J and Knopman, DS and Langbaum, JB and Park, DC and Ropacki, MT and Sikkes, SAM and Welsh-Bohmer, KA and Bain, LJ and Brashear, R and Budur, K and Graf, A and Martenyi, F and Storck, MS and Randolph, C},
   Title = {Measuring cognition and function in the preclinical stage of
             Alzheimer's disease.},
   Journal = {Alzheimers Dement (N Y)},
   Volume = {4},
   Pages = {64-75},
   Year = {2018},
   url = {http://dx.doi.org/10.1016/j.trci.2018.01.003},
   Abstract = {The Alzheimer's Association's Research Roundtable met in
             November 2016 to explore how best to measure changes in
             cognition and function in the preclinical stage of
             Alzheimer's disease. This review will cover the tools and
             instruments currently available to identify populations for
             prevention trials, and measure subtle disease progression in
             the earliest stages of Alzheimer's disease, and will include
             discussions of suitable cognitive, behavioral, functional,
             composite, and biological endpoints for prevention trials.
             Current prevention trials are reviewed including TOMMOROW,
             Alzheimer's Prevention Initiative Autosomal Dominant
             Alzheimer's Disease Trial, the Alzheimer's Prevention
             Initiative Generation Study, and the Anti-Amyloid Treatment
             in Asymptomatic Alzheimer's to compare current approaches
             and tools that are being developed.},
   Doi = {10.1016/j.trci.2018.01.003},
   Key = {fds333554}
}

@article{fds332791,
   Author = {Weintraub, S and Besser, L and Dodge, HH and Teylan, M and Ferris, S and Goldstein, FC and Giordani, B and Kramer, J and Loewenstein, D and Marson, D and Mungas, D and Salmon, D and Welsh-Bohmer, K and Zhou, X-H and Shirk, SD and Atri, A and Kukull, WA and Phelps, C and Morris,
             JC},
   Title = {Version 3 of the Alzheimer Disease Centers'
             Neuropsychological Test Battery in the Uniform Data Set
             (UDS).},
   Journal = {Alzheimer Dis Assoc Disord},
   Volume = {32},
   Number = {1},
   Pages = {10-17},
   Year = {2018},
   url = {http://dx.doi.org/10.1097/WAD.0000000000000223},
   Abstract = {INTRODUCTION: The neuropsychological battery of the Uniform
             Data Set (UDSNB) was implemented in 2005 by the National
             Institute on Aging (NIA) Alzheimer Disease Centers program
             to measure cognitive performance in dementia and mild
             cognitive impairment due to Alzheimer Disease. This paper
             describes a revision, the UDSNB 3.0. METHODS: The
             Neuropsychology Work Group of the NIA Clinical Task Force
             recommended revisions through a process of due diligence to
             address shortcomings of the original battery. The UDSNB 3.0
             covers episodic memory, processing speed, executive
             function, language, and constructional ability. Data from
             3602 cognitively normal participants in the National
             Alzheimer Coordinating Center database were analyzed.
             RESULTS: Descriptive statistics are presented. Multivariable
             linear regression analyses demonstrated score differences by
             age, sex, and education and were also used to create a
             normative calculator available online. DISCUSSION: The UDSNB
             3.0 neuropsychological battery provides a valuable non
             proprietary resource for conducting research on cognitive
             aging and dementia.},
   Doi = {10.1097/WAD.0000000000000223},
   Key = {fds332791}
}

@article{fds342566,
   Author = {Whitson, HE and Potter, GG and Feld, JA and Plassman, BL and Reynolds,
             K and Sloane, R and Welsh-Bohmer, KA},
   Title = {Dual-Task Gait and Alzheimer's Disease Genetic Risk in
             Cognitively Normal Adults: A Pilot Study.},
   Journal = {J Alzheimers Dis},
   Volume = {64},
   Number = {4},
   Pages = {1137-1148},
   Year = {2018},
   url = {http://dx.doi.org/10.3233/JAD-180016},
   Abstract = {BACKGROUND: Dual-task paradigms, in which an individual
             performs tasks separately and then concurrently, often
             demonstrate that people with neurodegenerative disorders
             experience more dual-task interference, defined as worse
             performance in the dual-task condition compared to the
             single-task condition. OBJECTIVE: To examine how
             gait-cognition dual-task performance differs between
             cognitively normal older adults with and without an APOE ɛ4
             allele. METHODS: Twenty-nine individuals ages 60 to 72 with
             normal cognition completed a dual-task protocol in which
             walking and cognitive tasks (executive function, memory)
             were performed separately and concurrently. Fourteen
             participants carried APOE ɛ4 alleles (ɛ3/ɛ4 or ɛ2/ɛ4);
             fifteen had APOE genotypes (ɛ2/ɛ2, ɛ2/ɛ3, or ɛ3/ɛ3)
             associated with lower risk of Alzheimer's disease (AD).
             RESULTS: The two risk groups did not differ by age, sex,
             race, education, or gait or cognitive measures under
             single-task conditions. Compared to low risk participants,
             APOE ɛ4 carriers tended to exhibit greater dual-task
             interference. Both the memory and executive function tasks
             resulted in dual-task interference on gait, but effect sizes
             for a group difference were larger when the cognitive task
             was executive function. In the dual-task protocol that
             combined walking and the executive function task, effect
             sizes for group difference in gait interference were larger
             (0.62- 0.70) than for cognitive interference (0.45- 0.47).
             DISCUSSION: Dual-task paradigms may reveal subtle changes in
             brain function in asymptomatic individuals at heightened
             risk of AD.},
   Doi = {10.3233/JAD-180016},
   Key = {fds342566}
}

@article{fds339676,
   Author = {Morrison, RL and Pei, H and Novak, G and Kaufer, DI and Welsh-Bohmer,
             KA and Ruhmel, S and Narayan, VA},
   Title = {A computerized, self-administered test of verbal episodic
             memory in elderly patients with mild cognitive impairment
             and healthy participants: A randomized, crossover,
             validation study.},
   Journal = {Alzheimers Dement (Amst)},
   Volume = {10},
   Pages = {647-656},
   Year = {2018},
   url = {http://dx.doi.org/10.1016/j.dadm.2018.08.010},
   Abstract = {INTRODUCTION: Performance of "Revere", a novel
             iPad-administered word-list recall (WLR) test, in
             quantifying deficits in verbal episodic memory, was
             evaluated versus examiner-administered Rey Auditory Verbal
             Learning Test (RAVLT) in patients with mild cognitive
             impairment and cognitively normal participants. METHODS:
             Elderly patients with clinically diagnosed mild cognitive
             impairment (Montreal Cognitive Assessment score 24-27) and
             cognitively normal (Montreal Cognitive Assessment score
             ≥28) were administered RAVLT or Revere in a randomized
             crossover design. RESULTS: A total of 153/161 participants
             (Revere/RAVLT n = 75; RAVLT/Revere n = 78) were
             randomized; 148 (97%) completed study; 121 patients (mean
             [standard deviation] age: 70.4 [7.84] years) were included
             for analysis. Word-list recall scores (8 trials) were
             comparable between Revere and RAVLT (Pearson's correlation
             coefficients: 0.12-0.70; least square mean difference
             [Revere-RAVLT]: -0.84 [90% CI, -1.15; -0.54]). Model factor
             estimates indicated trial (P < .001), period (P < .001)
             and evaluation sequence (P = .038) as significant factors.
             Learning over trials index and serial position effects were
             comparable. DISCUSSION: Participants' verbal recall
             performance on Revere and RAVLT were equivalent.},
   Doi = {10.1016/j.dadm.2018.08.010},
   Key = {fds339676}
}

@article{fds328886,
   Author = {Sims, R and van der Lee, SJ and Naj, AC and Bellenguez, C and Badarinarayan, N and Jakobsdottir, J and Kunkle, BW and Boland, A and Raybould, R and Bis, JC and Martin, ER and Grenier-Boley, B and Heilmann-Heimbach, S and Chouraki, V and Kuzma, AB and Sleegers, K and Vronskaya, M and Ruiz, A and Graham, RR and Olaso, R and Hoffmann, P and Grove, ML and Vardarajan, BN and Hiltunen, M and Nöthen, MM and White,
             CC and Hamilton-Nelson, KL and Epelbaum, J and Maier, W and Choi, S-H and Beecham, GW and Dulary, C and Herms, S and Smith, AV and Funk, CC and Derbois, C and Forstner, AJ and Ahmad, S and Li, H and Bacq, D and Harold,
             D and Satizabal, CL and Valladares, O and Squassina, A and Thomas, R and Brody, JA and Qu, L and Sánchez-Juan, P and Morgan, T and Wolters, FJ and Zhao, Y and Garcia, FS and Denning, N and Fornage, M and Malamon, J and Naranjo, MCD and Majounie, E and Mosley, TH and Dombroski, B and Wallon,
             D and Lupton, MK and Dupuis, J and Whitehead, P and Fratiglioni, L and Medway, C and Jian, X and Mukherjee, S and Keller, L and Brown, K and Lin,
             H and Cantwell, LB and Panza, F and McGuinness, B and Moreno-Grau, S and Burgess, JD and Solfrizzi, V and Proitsi, P and Adams, HH and Allen, M and Seripa, D and Pastor, P and Cupples, LA and Price, ND and Hannequin, D and Frank-García, A and Levy, D and Chakrabarty, P and Caffarra, P and Giegling, I and Beiser, AS and Giedraitis, V and Hampel, H and Garcia,
             ME and Wang, X and Lannfelt, L and Mecocci, P and Eiriksdottir, G and Crane, PK and Pasquier, F and Boccardi, V and Henández, I and Barber,
             RC and Scherer, M and Tarraga, L and Adams, PM and Leber, M and Chen, Y and Albert, MS and Riedel-Heller, S and Emilsson, V and Beekly, D and Braae,
             A and Schmidt, R and Blacker, D and Masullo, C and Schmidt, H and Doody,
             RS and Spalletta, G and Longstreth, WT and Fairchild, TJ and Bossù, P and Lopez, OL and Frosch, MP and Sacchinelli, E and Ghetti, B and Yang, Q and Huebinger, RM and Jessen, F and Li, S and Kamboh, MI and Morris, J and Sotolongo-Grau, O and Katz, MJ and Corcoran, C and Dunstan, M and Braddel, A and Thomas, C and Meggy, A and Marshall, R and Gerrish, A and Chapman, J and Aguilar, M and Taylor, S and Hill, M and Fairén, MD and Hodges, A and Vellas, B and Soininen, H and Kloszewska, I and Daniilidou, M and Uphill, J and Patel, Y and Hughes, JT and Lord, J and Turton, J and Hartmann, AM and Cecchetti, R and Fenoglio, C and Serpente, M and Arcaro, M and Caltagirone, C and Orfei, MD and Ciaramella, A and Pichler, S and Mayhaus, M and Gu, W and Lleó, A and Fortea, J and Blesa, R and Barber, IS and Brookes, K and Cupidi, C and Maletta, RG and Carrell, D and Sorbi, S and Moebus, S and Urbano, M and Pilotto, A and Kornhuber, J and Bosco, P and Todd, S and Craig, D and Johnston, J and Gill, M and Lawlor, B and Lynch, A and Fox, NC and Hardy,
             J and ARUK Consortium, and Albin, RL and Apostolova, LG and Arnold, SE and Asthana, S and Atwood, CS and Baldwin, CT and Barnes, LL and Barral, S and Beach, TG and Becker, JT and Bigio, EH and Bird, TD and Boeve, BF and Bowen, JD and Boxer, A and Burke, JR and Burns, JM and Buxbaum, JD and Cairns, NJ and Cao, C and Carlson, CS and Carlsson, CM and Carney, RM and Carrasquillo, MM and Carroll, SL and Diaz, CC and Chui, HC and Clark,
             DG and Cribbs, DH and Crocco, EA and DeCarli, C and Dick, M and Duara, R and Evans, DA and Faber, KM and Fallon, KB and Fardo, DW and Farlow, MR and Ferris, S and Foroud, TM and Galasko, DR and Gearing, M and Geschwind,
             DH and Gilbert, JR and Graff-Radford, NR and Green, RC and Growdon, JH and Hamilton, RL and Harrell, LE and Honig, LS and Huentelman, MJ and Hulette, CM and Hyman, BT and Jarvik, GP and Abner, E and Jin, L-W and Jun,
             G and Karydas, A and Kaye, JA and Kim, R and Kowall, NW and Kramer, JH and LaFerla, FM and Lah, JJ and Leverenz, JB and Levey, AI and Li, G and Lieberman, AP and Lunetta, KL and Lyketsos, CG and Marson, DC and Martiniuk, F and Mash, DC and Masliah, E and McCormick, WC and McCurry,
             SM and McDavid, AN and McKee, AC and Mesulam, M and Miller, BL and Miller,
             CA and Miller, JW and Morris, JC and Murrell, JR and Myers, AJ and O'Bryant, S and Olichney, JM and Pankratz, VS and Parisi, JE and Paulson, HL and Perry, W and Peskind, E and Pierce, A and Poon, WW and Potter, H and Quinn, JF and Raj, A and Raskind, M and Reisberg, B and Reitz, C and Ringman, JM and Roberson, ED and Rogaeva, E and Rosen, HJ and Rosenberg, RN and Sager, MA and Saykin, AJ and Schneider, JA and Schneider, LS and Seeley, WW and Smith, AG and Sonnen, JA and Spina, S and Stern, RA and Swerdlow, RH and Tanzi, RE and Thornton-Wells, TA and Trojanowski, JQ and Troncoso, JC and Van Deerlin and VM and Van Eldik,
             LJ and Vinters, HV and Vonsattel, JP and Weintraub, S and Welsh-Bohmer,
             KA and Wilhelmsen, KC and Williamson, J and Wingo, TS and Woltjer, RL and Wright, CB and Yu, C-E and Yu, L and Garzia, F and Golamaully, F and Septier, G and Engelborghs, S and Vandenberghe, R and De Deyn and PP and Fernadez, CM and Benito, YA and Thonberg, H and Forsell, C and Lilius,
             L and Kinhult-Stählbom, A and Kilander, L and Brundin, R and Concari,
             L and Helisalmi, S and Koivisto, AM and Haapasalo, A and Dermecourt, V and Fievet, N and Hanon, O and Dufouil, C and Brice, A and Ritchie, K and Dubois, B and Himali, JJ and Keene, CD and Tschanz, J and Fitzpatrick,
             AL and Kukull, WA and Norton, M and Aspelund, T and Larson, EB and Munger,
             R and Rotter, JI and Lipton, RB and Bullido, MJ and Hofman, A and Montine,
             TJ and Coto, E and Boerwinkle, E and Petersen, RC and Alvarez, V and Rivadeneira, F and Reiman, EM and Gallo, M and O'Donnell, CJ and Reisch,
             JS and Bruni, AC and Royall, DR and Dichgans, M and Sano, M and Galimberti,
             D and St George-Hyslop and P and Scarpini, E and Tsuang, DW and Mancuso, M and Bonuccelli, U and Winslow, AR and Daniele, A and Wu, C-K and GERAD/PERADES, CHARGE and ADGC, EADI and Peters, O and Nacmias, B and Riemenschneider, M and Heun, R and Brayne, C and Rubinsztein, DC and Bras, J and Guerreiro, R and Al-Chalabi, A and Shaw, CE and Collinge, J and Mann, D and Tsolaki, M and Clarimón, J and Sussams, R and Lovestone, S and O'Donovan, MC and Owen, MJ and Behrens, TW and Mead, S and Goate, AM and Uitterlinden, AG and Holmes, C and Cruchaga, C and Ingelsson, M and Bennett, DA and Powell, J and Golde, TE and Graff, C and De Jager and PL and Morgan, K and Ertekin-Taner, N and Combarros, O and Psaty, BM and Passmore, P and Younkin, SG and Berr, C and Gudnason, V and Rujescu, D and Dickson, DW and Dartigues, J-F and DeStefano, AL and Ortega-Cubero,
             S and Hakonarson, H and Campion, D and Boada, M and Kauwe, JK and Farrer,
             LA and Van Broeckhoven and C and Ikram, MA and Jones, L and Haines, JL and Tzourio, C and Launer, LJ and Escott-Price, V and Mayeux, R and Deleuze,
             J-F and Amin, N and Holmans, PA and Pericak-Vance, MA and Amouyel, P and van Duijn, CM and Ramirez, A and Wang, L-S and Lambert, J-C and Seshadri, S and Williams, J and Schellenberg, GD},
   Title = {Rare coding variants in PLCG2, ABI3, and TREM2 implicate
             microglial-mediated innate immunity in Alzheimer's
             disease.},
   Journal = {Nat Genet},
   Volume = {49},
   Number = {9},
   Pages = {1373-1384},
   Year = {2017},
   Month = {September},
   url = {http://dx.doi.org/10.1038/ng.3916},
   Abstract = {We identified rare coding variants associated with
             Alzheimer's disease in a three-stage case-control study of
             85,133 subjects. In stage 1, we genotyped 34,174 samples
             using a whole-exome microarray. In stage 2, we tested
             associated variants (P < 1 × 10-4) in 35,962 independent
             samples using de novo genotyping and imputed genotypes. In
             stage 3, we used an additional 14,997 samples to test the
             most significant stage 2 associations (P < 5 × 10-8) using
             imputed genotypes. We observed three new genome-wide
             significant nonsynonymous variants associated with
             Alzheimer's disease: a protective variant in PLCG2
             (rs72824905: p.Pro522Arg, P = 5.38 × 10-10, odds ratio (OR)
             = 0.68, minor allele frequency (MAF)cases = 0.0059,
             MAFcontrols = 0.0093), a risk variant in ABI3 (rs616338:
             p.Ser209Phe, P = 4.56 × 10-10, OR = 1.43, MAFcases = 0.011,
             MAFcontrols = 0.008), and a new genome-wide significant
             variant in TREM2 (rs143332484: p.Arg62His, P = 1.55 ×
             10-14, OR = 1.67, MAFcases = 0.0143, MAFcontrols = 0.0089),
             a known susceptibility gene for Alzheimer's disease. These
             protein-altering changes are in genes highly expressed in
             microglia and highlight an immune-related protein-protein
             interaction network enriched for previously identified risk
             genes in Alzheimer's disease. These genetic findings provide
             additional evidence that the microglia-mediated innate
             immune response contributes directly to the development of
             Alzheimer's disease.},
   Doi = {10.1038/ng.3916},
   Key = {fds328886}
}

@article{fds327422,
   Author = {Jun, GR and Chung, J and Mez, J and Barber, R and Beecham, GW and Bennett,
             DA and Buxbaum, JD and Byrd, GS and Carrasquillo, MM and Crane, PK and Cruchaga, C and De Jager and P and Ertekin-Taner, N and Evans, D and Fallin, MD and Foroud, TM and Friedland, RP and Goate, AM and Graff-Radford, NR and Hendrie, H and Hall, KS and Hamilton-Nelson,
             KL and Inzelberg, R and Kamboh, MI and Kauwe, JSK and Kukull, WA and Kunkle, BW and Kuwano, R and Larson, EB and Logue, MW and Manly, JJ and Martin, ER and Montine, TJ and Mukherjee, S and Naj, A and Reiman, EM and Reitz, C and Sherva, R and St George-Hyslop and PH and Thornton, T and Younkin, SG and Vardarajan, BN and Wang, L-S and Wendlund, JR and Winslow, AR and Alzheimer's Disease Genetics Consortium, and Haines, J and Mayeux, R and Pericak-Vance, MA and Schellenberg, G and Lunetta, KL and Farrer, LA},
   Title = {Transethnic genome-wide scan identifies novel Alzheimer's
             disease loci.},
   Journal = {Alzheimers Dement},
   Volume = {13},
   Number = {7},
   Pages = {727-738},
   Year = {2017},
   Month = {July},
   url = {http://dx.doi.org/10.1016/j.jalz.2016.12.012},
   Abstract = {INTRODUCTION: Genetic loci for Alzheimer's disease (AD) have
             been identified in whites of European ancestry, but the
             genetic architecture of AD among other populations is less
             understood. METHODS: We conducted a transethnic genome-wide
             association study (GWAS) for late-onset AD in Stage 1 sample
             including whites of European Ancestry, African-Americans,
             Japanese, and Israeli-Arabs assembled by the Alzheimer's
             Disease Genetics Consortium. Suggestive results from Stage 1
             from novel loci were followed up using summarized results in
             the International Genomics Alzheimer's Project GWAS dataset.
             RESULTS: Genome-wide significant (GWS) associations in
             single-nucleotide polymorphism (SNP)-based tests (P < 5 ×
             10-8) were identified for SNPs in PFDN1/HBEGF,
             USP6NL/ECHDC3, and BZRAP1-AS1 and for the interaction of the
             (apolipoprotein E) APOE ε4 allele with NFIC SNP. We also
             obtained GWS evidence (P < 2.7 × 10-6) for gene-based
             association in the total sample with a novel locus, TPBG
             (P = 1.8 × 10-6). DISCUSSION: Our findings highlight the
             value of transethnic studies for identifying novel AD
             susceptibility loci.},
   Doi = {10.1016/j.jalz.2016.12.012},
   Key = {fds327422}
}

@article{fds322219,
   Author = {Browndyke, JN and Berger, M and Harshbarger, TB and Smith, PJ and White,
             W and Bisanar, TL and Alexander, JH and Gaca, JG and Welsh-Bohmer, K and Newman, MF and Mathew, JP},
   Title = {Resting-State Functional Connectivity and Cognition After
             Major Cardiac Surgery in Older Adults without Preoperative
             Cognitive Impairment: Preliminary Findings.},
   Journal = {J Am Geriatr Soc},
   Volume = {65},
   Number = {1},
   Pages = {e6-e12},
   Year = {2017},
   Month = {January},
   url = {http://dx.doi.org/10.1111/jgs.14534},
   Abstract = {OBJECTIVES: To look for changes in intrinsic functional
             brain connectivity associated with postoperative changes in
             cognition, a common complication in seniors undergoing major
             surgery, using resting-state functional magnetic resonance
             imaging. DESIGN: Objective cognitive testing and functional
             brain imaging were prospectively performed at preoperative
             baseline and 6 weeks after surgery and at the same time
             intervals in nonsurgical controls. SETTING: Academic medical
             center. PARTICIPANTS: Older adults undergoing cardiac
             surgery (n = 12) and nonsurgical older adult controls with a
             history of coronary artery disease (n = 12); no participants
             had cognitive impairment at preoperative baseline
             (Mini-Mental State Examination score >27). MEASUREMENTS:
             Differences in resting-state functional connectivity (RSFC)
             and global cognitive change relationships were assessed
             using a voxel-wise intrinsic connectivity method,
             controlling for demographic factors and pre- and
             perioperative cerebral white matter disease volume. Analyses
             were corrected for multiple comparisons (false discovery
             rate P < .01). RESULTS: Global cognitive change after
             cardiac surgery was significantly associated with intrinsic
             RSFC changes in regions of the posterior cingulate cortex
             and right superior frontal gyrus-anatomical and functional
             locations of the brain's default mode network (DMN). No
             statistically significant relationships were found between
             global cognitive change and RSFC change in nonsurgical
             controls. CONCLUSION: Clinicians have long known that some
             older adults develop postoperative cognitive dysfunction
             (POCD) after anesthesia and surgery, yet the neurobiological
             correlates of POCD are not well defined. The current results
             suggest that altered RSFC in specific DMN regions is
             positively correlated with global cognitive change 6 weeks
             after cardiac surgery, suggesting that DMN activity and
             connectivity could be important diagnostic markers of POCD
             or intervention targets for potential POCD treatment
             efforts.},
   Doi = {10.1111/jgs.14534},
   Key = {fds322219}
}

@article{fds339373,
   Author = {Aisen, P and Touchon, J and Amariglio, R and Andrieu, S and Bateman, R and Breitner, J and Donohue, M and Dunn, B and Doody, R and Fox, N and Gauthier, S and Grundman, M and Hendrix, S and Ho, C and Isaac, M and Raman, R and Rosenberg, P and Schindler, R and Schneider, L and Sperling, R and Tariot, P and Welsh-Bohmer, K and Weiner, M and Vellas,
             B},
   Title = {EU/US/CTAD Task Force: Lessons Learned from Recent and
             Current Alzheimer's Prevention Trials.},
   Journal = {J Prev Alzheimers Dis},
   Volume = {4},
   Number = {2},
   Pages = {116-124},
   Year = {2017},
   url = {http://dx.doi.org/10.14283/jpad.2017.13},
   Abstract = {At a meeting of the EU/US/Clinical Trials in Alzheimer's
             Disease (CTAD) Task Force in December 2016, an international
             group of investigators from industry, academia, and
             regulatory agencies reviewed lessons learned from ongoing
             and planned prevention trials, which will help guide future
             clinical trials of AD treatments, particularly in the
             pre-clinical space. The Task Force discussed challenges that
             need to be addressed across all aspects of clinical trials,
             calling for innovation in recruitment and retention,
             infrastructure development, and the selection of outcome
             measures. While cognitive change provides a marker of
             disease progression across the disease continuum, there
             remains a need to identify the optimal assessment tools that
             provide clinically meaningful endpoints. Patient- and
             informant-reported assessments of cognition and function may
             be useful but present additional challenges. Imaging and
             other biomarkers are also essential to maximize the
             efficiency of and the information learned from clinical
             trials.},
   Doi = {10.14283/jpad.2017.13},
   Key = {fds339373}
}

@article{fds326771,
   Author = {Blumenthal, JA and Smith, PJ and Mabe, S and Hinderliter, A and Welsh-Bohmer, K and Browndyke, JN and Lin, P-H and Kraus, W and Doraiswamy, PM and Burke, J and Sherwood, A},
   Title = {Lifestyle and Neurocognition in Older Adults With
             Cardiovascular Risk Factors and Cognitive
             Impairment.},
   Journal = {Psychosom Med},
   Volume = {79},
   Number = {6},
   Pages = {719-727},
   Year = {2017},
   url = {http://dx.doi.org/10.1097/PSY.0000000000000474},
   Abstract = {OBJECTIVE: The aim of the study was to determine the
             relationship of lifestyle factors and neurocognitive
             functioning in older adults with vascular risk factors and
             cognitive impairment, no dementia (CIND). METHODS: One
             hundred sixty adults (M [SD] = 65.4 [6.8] years) with CIND
             completed neurocognitive assessments of executive function,
             processing speed, and memory. Objective measures of physical
             activity using accelerometry, aerobic capacity determined by
             exercise testing, and dietary habits quantified by the Food
             Frequency Questionnaire and 4-Day Food Diary to assess
             adherence to the Mediterranean and Dietary Approaches to
             Stop Hypertension (DASH) diets were obtained to assess
             direct effects with neurocognition. Potential indirect
             associations of high-sensitivity C-reactive protein and the
             Framingham Stroke Risk Profile also were examined. RESULTS:
             Greater aerobic capacity (β = 0.24) and daily physical
             activity (β = 0.15) were associated with better executive
             functioning/processing speed and verbal memory (βs = 0.24;
             0.16). Adherence to the DASH diet was associated with better
             verbal memory (β = 0.17). Greater high-sensitivity
             C-reactive protein (βs = -0.14; -0.21) and Framingham
             Stroke Risk Profile (β = -0.18; -0.18) were associated with
             poorer executive functioning/processing speed and verbal
             memory. Greater stroke risk partially mediated the
             association of aerobic capacity with executive
             functioning/processing speed, and verbal memory and greater
             inflammation partially mediated the association of physical
             activity and aerobic fitness, with verbal memory.
             CONCLUSIONS: Higher levels of physical activity, aerobic
             fitness, and adherence to the DASH diet are associated with
             better neurocognitive performance in adults with CIND. These
             findings suggest that the adoption of healthy lifestyle
             habits could reduce the risk of neurocognitive decline in
             vulnerable older adults. CLINICAL TRIAL REGISTRATION:
             NCT01573546.},
   Doi = {10.1097/PSY.0000000000000474},
   Key = {fds326771}
}

@article{fds322773,
   Author = {Weninger, S and Carrillo, MC and Dunn, B and Aisen, PS and Bateman, RJ and Kotz, JD and Langbaum, JB and Mills, SL and Reiman, EM and Sperling, R and Santacruz, AM and Tariot, PN and Welsh-Bohmer,
             KA},
   Title = {Collaboration for Alzheimer's Prevention: Principles to
             guide data and sample sharing in preclinical Alzheimer's
             disease trials.},
   Journal = {Alzheimers Dement},
   Volume = {12},
   Number = {5},
   Pages = {631-632},
   Year = {2016},
   Month = {May},
   url = {http://dx.doi.org/10.1016/j.jalz.2016.04.001},
   Doi = {10.1016/j.jalz.2016.04.001},
   Key = {fds322773}
}

@article{fds322221,
   Author = {Ridge, PG and Hoyt, KB and Boehme, K and Mukherjee, S and Crane, PK and Haines, JL and Mayeux, R and Farrer, LA and Pericak-Vance, MA and Schellenberg, GD and Kauwe, JSK and Alzheimer's Disease Genetics
             Consortium (ADGC)},
   Title = {Assessment of the genetic variance of late-onset Alzheimer's
             disease.},
   Journal = {Neurobiol Aging},
   Volume = {41},
   Pages = {200.e13-200.e20},
   Year = {2016},
   Month = {May},
   url = {http://dx.doi.org/10.1016/j.neurobiolaging.2016.02.024},
   Abstract = {Alzheimer's disease (AD) is a complex genetic disorder with
             no effective treatments. More than 20 common markers have
             been identified, which are associated with AD. Recently,
             several rare variants have been identified in Amyloid
             Precursor Protein (APP), Triggering Receptor Expressed On
             Myeloid Cells 2 (TREM2) and Unc-5 Netrin Receptor C (UNC5C)
             that affect risk for AD. Despite the many successes, the
             genetic architecture of AD remains unsolved. We used
             Genome-wide Complex Trait Analysis to (1) estimate
             phenotypic variance explained by genetics; (2) calculate
             genetic variance explained by known AD single nucleotide
             polymorphisms (SNPs); and (3) identify the genomic locations
             of variation that explain the remaining unexplained genetic
             variance. In total, 53.24% of phenotypic variance is
             explained by genetics, but known AD SNPs only explain 30.62%
             of the genetic variance. Of the unexplained genetic
             variance, approximately 41% is explained by unknown SNPs in
             regions adjacent to known AD SNPs, and the remaining
             unexplained genetic variance outside these
             regions.},
   Doi = {10.1016/j.neurobiolaging.2016.02.024},
   Key = {fds322221}
}

@article{fds322220,
   Author = {Lutz, MW and Crenshaw, D and Welsh-Bohmer, KA and Burns, DK and Roses,
             AD},
   Title = {New Genetic Approaches to AD: Lessons from APOE-TOMM40
             Phylogenetics.},
   Journal = {Curr Neurol Neurosci Rep},
   Volume = {16},
   Number = {5},
   Pages = {48},
   Year = {2016},
   Month = {May},
   url = {http://dx.doi.org/10.1007/s11910-016-0643-8},
   Abstract = {Clinical trials for Alzheimer's disease are now focusing on
             the earliest stages of the disease with the goal of delaying
             dementia onset. There is great utility in using genetic
             variants to identify individuals at high age-dependent risk
             when the goal is to begin treatment before the development
             of any cognitive symptoms. Genetic variants identified
             through large-scale genome-wide association studies have not
             substantially improved the accuracy provided by APOE
             genotype to identify people at high risk of late-onset
             Alzheimer's disease (LOAD). We describe novel approaches,
             focused on molecular phylogenetics, to finding genetic
             variants that predict age at LOAD onset with sufficient
             accuracy and precision to be useful. We highlight the
             discovery of a polymorphism in TOMM40 that, in addition to
             APOE, may improve risk prediction and review how TOMM40
             genetic variants may impact the develop of LOAD
             independently from APOE. The analysis methods described in
             this review may be useful for other genetically complex
             human diseases.},
   Doi = {10.1007/s11910-016-0643-8},
   Key = {fds322220}
}

@article{fds327247,
   Author = {Perry, DC and Sturm, VE and Peterson, MJ and Pieper, CF and Bullock, T and Boeve, BF and Miller, BL and Guskiewicz, KM and Berger, MS and Kramer,
             JH and Welsh-Bohmer, KA},
   Title = {Association of traumatic brain injury with subsequent
             neurological and psychiatric disease: a meta-analysis.},
   Journal = {J Neurosurg},
   Volume = {124},
   Number = {2},
   Pages = {511-526},
   Year = {2016},
   Month = {February},
   url = {http://dx.doi.org/10.3171/2015.2.JNS14503},
   Abstract = {OBJECTIVE: Mild traumatic brain injury (TBI) has been
             proposed as a risk factor for the development of Alzheimer's
             disease, Parkinson's disease, depression, and other
             illnesses. This study's objective was to determine the
             association of prior mild TBI with the subsequent diagnosis
             (that is, at least 1 year postinjury) of neurological or
             psychiatric disease. METHODS: All studies from January 1995
             to February 2012 reporting TBI as a risk factor for
             diagnoses of interest were identified by searching PubMed,
             study references, and review articles. Reviewers abstracted
             the data and assessed study designs and characteristics.
             RESULTS: Fifty-seven studies met the inclusion criteria. A
             random effects meta-analysis revealed a significant
             association of prior TBI with subsequent neurological and
             psychiatric diagnoses. The pooled odds ratio (OR) for the
             development of any illness subsequent to prior TBI was 1.67
             (95% CI 1.44-1.93, p < 0.0001). Prior TBI was independently
             associated with both neurological (OR 1.55, 95% CI
             1.31-1.83, p < 0.0001) and psychiatric (OR 2.00, 95% CI
             1.50-2.66, p < 0.0001) outcomes. Analyses of individual
             diagnoses revealed higher odds of Alzheimer's disease,
             Parkinson's disease, mild cognitive impairment, depression,
             mixed affective disorders, and bipolar disorder in
             individuals with previous TBI as compared to those without
             TBI. This association was present when examining only
             studies of mild TBI and when considering the influence of
             study design and characteristics. Analysis of a subset of
             studies demonstrated no evidence that multiple TBIs were
             associated with higher odds of disease than a single TBI.
             CONCLUSIONS: History of TBI, including mild TBI, is
             associated with the development of neurological and
             psychiatric illness. This finding indicates that either TBI
             is a risk factor for heterogeneous pathological processes or
             that TBI may contribute to a common pathological
             mechanism.},
   Doi = {10.3171/2015.2.JNS14503},
   Key = {fds327247}
}

@article{fds327851,
   Author = {Jun, G and Ibrahim-Verbaas, CA and Vronskaya, M and Lambert, J-C and Chung, J and Naj, AC and Kunkle, BW and Wang, L-S and Bis, JC and Bellenguez, C and Harold, D and Lunetta, KL and Destefano, AL and Grenier-Boley, B and Sims, R and Beecham, GW and Smith, AV and Chouraki,
             V and Hamilton-Nelson, KL and Ikram, MA and Fievet, N and Denning, N and Martin, ER and Schmidt, H and Kamatani, Y and Dunstan, ML and Valladares, O and Laza, AR and Zelenika, D and Ramirez, A and Foroud,
             TM and Choi, S-H and Boland, A and Becker, T and Kukull, WA and van der
             Lee, SJ and Pasquier, F and Cruchaga, C and Beekly, D and Fitzpatrick,
             AL and Hanon, O and Gill, M and Barber, R and Gudnason, V and Campion, D and Love, S and Bennett, DA and Amin, N and Berr, C and Tsolaki, M and Buxbaum,
             JD and Lopez, OL and Deramecourt, V and Fox, NC and Cantwell, LB and Tárraga, L and Dufouil, C and Hardy, J and Crane, PK and Eiriksdottir,
             G and Hannequin, D and Clarke, R and Evans, D and Mosley, TH and Letenneur,
             L and Brayne, C and Maier, W and De Jager and P and Emilsson, V and Dartigues,
             J-F and Hampel, H and Kamboh, MI and de Bruijn, RFAG and Tzourio, C and Pastor, P and Larson, EB and Rotter, JI and O'Donovan, MC and Montine,
             TJ and Nalls, MA and Mead, S and Reiman, EM and Jonsson, PV and Holmes, C and St George-Hyslop and PH and Boada, M and Passmore, P and Wendland, JR and Schmidt, R and Morgan, K and Winslow, AR and Powell, JF and Carasquillo,
             M and Younkin, SG and Jakobsdóttir, J and Kauwe, JSK and Wilhelmsen,
             KC and Rujescu, D and Nöthen, MM and Hofman, A and Jones, L and IGAP
             Consortium, and Haines, JL and Psaty, BM and Van Broeckhoven and C and Holmans, P and Launer, LJ and Mayeux, R and Lathrop, M and Goate, AM and Escott-Price, V and Seshadri, S and Pericak-Vance, MA and Amouyel, P and Williams, J and van Duijn, CM and Schellenberg, GD and Farrer,
             LA},
   Title = {A novel Alzheimer disease locus located near the gene
             encoding tau protein.},
   Journal = {Mol Psychiatry},
   Volume = {21},
   Number = {1},
   Pages = {108-117},
   Year = {2016},
   Month = {January},
   url = {http://dx.doi.org/10.1038/mp.2015.23},
   Abstract = {APOE ɛ4, the most significant genetic risk factor for
             Alzheimer disease (AD), may mask effects of other loci. We
             re-analyzed genome-wide association study (GWAS) data from
             the International Genomics of Alzheimer's Project (IGAP)
             Consortium in APOE ɛ4+ (10 352 cases and 9207 controls)
             and APOE ɛ4- (7184 cases and 26 968 controls) subgroups
             as well as in the total sample testing for interaction
             between a single-nucleotide polymorphism (SNP) and APOE ɛ4
             status. Suggestive associations (P<1 × 10(-4)) in stage 1
             were evaluated in an independent sample (stage 2) containing
             4203 subjects (APOE ɛ4+: 1250 cases and 536 controls; APOE
             ɛ4-: 718 cases and 1699 controls). Among APOE ɛ4-
             subjects, novel genome-wide significant (GWS) association
             was observed with 17 SNPs (all between KANSL1 and LRRC37A on
             chromosome 17 near MAPT) in a meta-analysis of the stage 1
             and stage 2 data sets (best SNP, rs2732703, P=5·8 ×
             10(-9)). Conditional analysis revealed that rs2732703
             accounted for association signals in the entire 100-kilobase
             region that includes MAPT. Except for previously identified
             AD loci showing stronger association in APOE ɛ4+ subjects
             (CR1 and CLU) or APOE ɛ4- subjects (MS4A6A/MS4A4A/MS4A6E),
             no other SNPs were significantly associated with AD in a
             specific APOE genotype subgroup. In addition, the finding in
             the stage 1 sample that AD risk is significantly influenced
             by the interaction of APOE with rs1595014 in TMEM106B
             (P=1·6 × 10(-7)) is noteworthy, because TMEM106B variants
             have previously been associated with risk of frontotemporal
             dementia. Expression quantitative trait locus analysis
             revealed that rs113986870, one of the GWS SNPs near
             rs2732703, is significantly associated with four KANSL1
             probes that target transcription of the first translated
             exon and an untranslated exon in hippocampus (P ⩽ 1.3 ×
             10(-8)), frontal cortex (P ⩽ 1.3 × 10(-9)) and temporal
             cortex (P⩽1.2 × 10(-11)). Rs113986870 is also strongly
             associated with a MAPT probe that targets transcription of
             alternatively spliced exon 3 in frontal cortex (P=9.2 ×
             10(-6)) and temporal cortex (P=2.6 × 10(-6)). Our
             APOE-stratified GWAS is the first to show GWS association
             for AD with SNPs in the chromosome 17q21.31 region.
             Replication of this finding in independent samples is needed
             to verify that SNPs in this region have significantly
             stronger effects on AD risk in persons lacking APOE ɛ4
             compared with persons carrying this allele, and if this is
             found to hold, further examination of this region and
             studies aimed at deciphering the mechanism(s) are
             warranted.},
   Doi = {10.1038/mp.2015.23},
   Key = {fds327851}
}

@article{fds322224,
   Author = {Lutz, MW and Sundseth, SS and Burns, DK and Saunders, AM and Hayden, KM and Burke, JR and Welsh-Bohmer, KA and Roses, AD},
   Title = {A Genetics-based Biomarker Risk Algorithm for Predicting
             Risk of Alzheimer's Disease.},
   Journal = {Alzheimers Dement (N Y)},
   Volume = {2},
   Number = {1},
   Pages = {30-44},
   Year = {2016},
   Month = {January},
   url = {http://dx.doi.org/10.1016/j.trci.2015.12.002},
   Abstract = {BACKGROUND: A straightforward, reproducible blood-based test
             that predicts age dependent risk of Alzheimer's disease (AD)
             could be used as an enrichment tool for clinical development
             of therapies. This study evaluated the prognostic
             performance of a genetics-based biomarker risk algorithm
             (GBRA) established on a combination of Apolipoprotein E
             (APOE)/Translocase of outer mitochondrial membrane 40
             homolog (TOMM40) genotypes and age, then compare it to
             cerebrospinal fluid (CSF) biomarkers, neuroimaging and
             neurocognitive tests using data from two independent AD
             cohorts. METHODS: The GBRA was developed using data from the
             prospective Bryan-ADRC study (n=407; 86 conversion events
             (mild cognitive impairment (MCI) or late onset Alzheimer's
             disease (LOAD)). The performance of the algorithm was tested
             using data from the ADNI study (n=660; 457 individuals
             categorized as MCI or LOAD). RESULTS: The positive
             predictive values (PPV) and negative predictive values (NPV)
             of the GBRA are in the range of 70-80%. The relatively high
             odds ratio (approximately 3-5) and significant net
             reclassification index (NRI) scores comparing the GBRA to a
             version based on APOE and age alone support the value of the
             GBRA in risk prediction for MCI due to LOAD. Performance of
             the GBRA compares favorably with CSF and imaging (fMRI)
             biomarkers. In addition, the GBRA "high" and "low" AD-risk
             categorizations correlated well with pathological CSF
             biomarker levels, PET amyloid burden and neurocognitive
             scores. CONCLUSIONS: Unlike dynamic markers (i.e., imaging,
             protein or lipid markers) that may be influenced by factors
             unrelated to disease, genomic DNA is easily collected,
             stable, and the technical methods for measurement are
             robust, inexpensive, and widely available. The performance
             characteristics of the GBRA support its use as a
             pharmacogenetic enrichment tool for LOAD delay of onset
             clinical trials, and merits further evaluation for its
             clinical utility in evaluating therapeutic
             efficacy.},
   Doi = {10.1016/j.trci.2015.12.002},
   Key = {fds322224}
}

@article{fds322774,
   Author = {Reiman, EM and Langbaum, JB and Tariot, PN and Lopera, F and Bateman,
             RJ and Morris, JC and Sperling, RA and Aisen, PS and Roses, AD and Welsh-Bohmer, KA and Carrillo, MC and Weninger,
             S},
   Title = {CAP--advancing the evaluation of preclinical Alzheimer
             disease treatments.},
   Journal = {Nat Rev Neurol},
   Volume = {12},
   Number = {1},
   Pages = {56-61},
   Year = {2016},
   Month = {January},
   url = {http://dx.doi.org/10.1038/nrneurol.2015.177},
   Abstract = {If we are to find treatments to postpone, reduce the risk
             of, or completely prevent the clinical onset of Alzheimer
             disease (AD), we need faster methods to evaluate promising
             preclinical AD treatments, new ways to work together in
             support of common goals, and a determination to expedite the
             initiation and performance of preclinical AD trials. In this
             article, we note some of the current challenges,
             opportunities and emerging strategies in preclinical AD
             treatment. We describe the Collaboration for Alzheimer's
             Prevention (CAP)-a convening, harmonizing and
             consensus-building initiative to help stakeholders advance
             AD prevention research with rigour, care and maximal
             impact-and we demonstrate the impact of CAP on the goals and
             design of new preclinical AD trials.},
   Doi = {10.1038/nrneurol.2015.177},
   Key = {fds322774}
}

@article{fds371165,
   Author = {Morrison, R and Pei, H and Novak, G and Kaufer, D and Welsh-Bohmer, K and Ruhmel, S and Narayan, VA},
   Title = {Validation of a Novel Computerized Self-Administered
             Memory-Screening Test With Automated Reporting (SAMSTAR) in
             Patients With Mild Cognitive Impairment and Normal Control
             Participants: A Randomized, Crossover, Controlled
             Study},
   Journal = {NEUROPSYCHOPHARMACOLOGY},
   Volume = {41},
   Pages = {S345-S346},
   Year = {2016},
   Key = {fds371165}
}

@article{fds322222,
   Author = {Karch, CM and Ezerskiy, LA and Bertelsen, S and Alzheimer’s Disease
             Genetics Consortium (ADGC), and Goate, AM},
   Title = {Alzheimer's Disease Risk Polymorphisms Regulate Gene
             Expression in the ZCWPW1 and the CELF1 Loci.},
   Journal = {PLoS One},
   Volume = {11},
   Number = {2},
   Pages = {e0148717},
   Editor = {Huang, Q},
   Year = {2016},
   url = {http://dx.doi.org/10.1371/journal.pone.0148717},
   Abstract = {Late onset Alzheimer's disease (LOAD) is a genetically
             complex and clinically heterogeneous disease. Recent
             large-scale genome wide association studies (GWAS) have
             identified more than twenty loci that modify risk for AD.
             Despite the identification of these loci, little progress
             has been made in identifying the functional variants that
             explain the association with AD risk. Thus, we sought to
             determine whether the novel LOAD GWAS single nucleotide
             polymorphisms (SNPs) alter expression of LOAD GWAS genes and
             whether expression of these genes is altered in AD brains.
             The majority of LOAD GWAS SNPs occur in gene dense regions
             under large linkage disequilibrium (LD) blocks, making it
             unclear which gene(s) are modified by the SNP. Thus, we
             tested for brain expression quantitative trait loci (eQTLs)
             between LOAD GWAS SNPs and SNPs in high LD with the LOAD
             GWAS SNPs in all of the genes within the GWAS loci. We found
             a significant eQTL between rs1476679 and PILRB and GATS,
             which occurs within the ZCWPW1 locus. PILRB and GATS
             expression levels, within the ZCWPW1 locus, were also
             associated with AD status. Rs7120548 was associated with
             MTCH2 expression, which occurs within the CELF1 locus.
             Additionally, expression of several genes within the CELF1
             locus, including MTCH2, were highly correlated with one
             another and were associated with AD status. We further
             demonstrate that PILRB, as well as other genes within the
             GWAS loci, are most highly expressed in microglia. These
             findings together with the function of PILRB as a DAP12
             receptor supports the critical role of microglia and
             neuroinflammation in AD risk.},
   Doi = {10.1371/journal.pone.0148717},
   Key = {fds322222}
}

@article{fds322223,
   Author = {Foster, CM and Addis, DR and Ford, JH and Kaufer, DI and Burke, JR and Browndyke, JN and Welsh-Bohmer, KA and Giovanello,
             KS},
   Title = {Prefrontal contributions to relational encoding in amnestic
             mild cognitive impairment.},
   Journal = {Neuroimage Clin},
   Volume = {11},
   Pages = {158-166},
   Year = {2016},
   url = {http://dx.doi.org/10.1016/j.nicl.2016.01.008},
   Abstract = {Relational memory declines are well documented as an early
             marker for amnestic mild cognitive impairment (aMCI).
             Episodic memory formation relies on relational processing
             supported by two mnemonic mechanisms, generation and
             binding. Neuroimaging studies using functional magnetic
             resonance imaging (fMRI) have primarily focused on binding
             deficits which are thought to be mediated by medial temporal
             lobe dysfunction. In this study, prefrontal contributions to
             relational encoding were also investigated using fMRI by
             parametrically manipulating generation demands during the
             encoding of word triads. Participants diagnosed with aMCI
             and healthy control subjects encoded word triads consisting
             of a category word with either, zero, one, or two
             semantically related exemplars. As the need to generate
             increased (i.e., two- to one- to zero-link triads), both
             groups recruited a core set of regions associated with the
             encoding of word triads including the parahippocampal gyrus,
             superior temporal gyrus, and superior parietal lobule.
             Participants diagnosed with aMCI also parametrically
             recruited several frontal regions including the inferior
             frontal gyrus and middle frontal gyrus as the need to
             generate increased, whereas the control participants did not
             show this modulation. While there is some functional overlap
             in regions recruited by generation demands between the
             groups, the recruitment of frontal regions in the aMCI
             participants coincides with worse memory performance, likely
             representing a form of neural inefficiency associated with
             Alzheimer's disease.},
   Doi = {10.1016/j.nicl.2016.01.008},
   Key = {fds322223}
}

@article{fds322775,
   Author = {Monsell, SE and Dodge, HH and Zhou, X-H and Bu, Y and Besser, LM and Mock,
             C and Hawes, SE and Kukull, WA and Weintraub, S and Neuropsychology Work
             Group Advisory to the Clinical Task Force},
   Title = {Results From the NACC Uniform Data Set Neuropsychological
             Battery Crosswalk Study.},
   Journal = {Alzheimer Dis Assoc Disord},
   Volume = {30},
   Number = {2},
   Pages = {134-139},
   Year = {2016},
   url = {http://dx.doi.org/10.1097/WAD.0000000000000111},
   Abstract = {INTRODUCTION: Four new nonproprietary tests were recommended
             for use in the National Alzheimer's Coordinating Center's
             Uniform Data Set Neuropsychological Battery. These tests are
             similar to previous tests but also allow for continuity of
             longitudinal data collection and wide dissemination among
             research collaborators. METHODS: A Crosswalk Study was
             conducted in early 2014 to assess the correlation between
             each set of new and previous tests. Tests with good
             correlation were equated using equipercentile equating. The
             resulting conversion tables allow scores on the new tests to
             be converted to equivalent scores on the previous tests.
             RESULTS: All pairs of tests had good correlation (ρ=0.68 to
             0.78). Learning effects were detected for Logical Memory
             only. Confidence intervals were narrow at each point
             estimate, and prediction accuracy was high. DISCUSSION: The
             recommended new tests are well correlated with the previous
             tests. The equipercentile equating method produced
             conversion tables that provide a useful reference for
             clinicians and researchers.},
   Doi = {10.1097/WAD.0000000000000111},
   Key = {fds322775}
}

@article{fds322225,
   Author = {Mukherjee, S and Walter, S and Kauwe, JSK and Saykin, AJ and Bennett,
             DA and Larson, EB and Crane, PK and Glymour, MM and Adult Changes in
             Thought Study Investigators, and Religious Orders Study/Memory and Aging Project Investigators, and Alzheimer's Disease
             Genetics Consortium},
   Title = {Genetically predicted body mass index and Alzheimer's
             disease-related phenotypes in three large samples: Mendelian
             randomization analyses.},
   Journal = {Alzheimers Dement},
   Volume = {11},
   Number = {12},
   Pages = {1439-1451},
   Year = {2015},
   Month = {December},
   url = {http://dx.doi.org/10.1016/j.jalz.2015.05.015},
   Abstract = {Observational research shows that higher body mass index
             (BMI) increases Alzheimer's disease (AD) risk, but it is
             unclear whether this association is causal. We applied
             genetic variants that predict BMI in Mendelian randomization
             analyses, an approach that is not biased by reverse
             causation or confounding, to evaluate whether higher BMI
             increases AD risk. We evaluated individual-level data from
             the AD Genetics Consortium (ADGC: 10,079 AD cases and 9613
             controls), the Health and Retirement Study (HRS: 8403
             participants with algorithm-predicted dementia status), and
             published associations from the Genetic and Environmental
             Risk for AD consortium (GERAD1: 3177 AD cases and 7277
             controls). No evidence from individual single-nucleotide
             polymorphisms or polygenic scores indicated BMI increased AD
             risk. Mendelian randomization effect estimates per BMI point
             (95% confidence intervals) were as follows: ADGC, odds ratio
             (OR) = 0.95 (0.90-1.01); HRS, OR = 1.00 (0.75-1.32); GERAD1,
             OR = 0.96 (0.87-1.07). One subscore (cellular processes not
             otherwise specified) unexpectedly predicted lower AD
             risk.},
   Doi = {10.1016/j.jalz.2015.05.015},
   Key = {fds322225}
}

@article{fds322226,
   Author = {Ghani, M and Reitz, C and Cheng, R and Vardarajan, BN and Jun, G and Sato,
             C and Naj, A and Rajbhandary, R and Wang, L-S and Valladares, O and Lin,
             C-F and Larson, EB and Graff-Radford, NR and Evans, D and De Jager and PL and Crane, PK and Buxbaum, JD and Murrell, JR and Raj, T and Ertekin-Taner,
             N and Logue, M and Baldwin, CT and Green, RC and Barnes, LL and Cantwell,
             LB and Fallin, MD and Go, RCP and Griffith, PA and Obisesan, TO and Manly,
             JJ and Lunetta, KL and Kamboh, MI and Lopez, OL and Bennett, DA and Hendrie, H and Hall, KS and Goate, AM and Byrd, GS and Kukull, WA and Foroud, TM and Haines, JL and Farrer, LA and Pericak-Vance, MA and Lee,
             JH and Schellenberg, GD and St George-Hyslop and P and Mayeux, R and Rogaeva, E and Alzheimer’s Disease Genetics
             Consortium},
   Title = {Association of Long Runs of Homozygosity With Alzheimer
             Disease Among African American Individuals.},
   Journal = {JAMA Neurol},
   Volume = {72},
   Number = {11},
   Pages = {1313-1323},
   Year = {2015},
   Month = {November},
   url = {http://dx.doi.org/10.1001/jamaneurol.2015.1700},
   Abstract = {IMPORTANCE: Mutations in known causal Alzheimer disease (AD)
             genes account for only 1% to 3% of patients and almost all
             are dominantly inherited. Recessive inheritance of complex
             phenotypes can be linked to long (>1-megabase [Mb]) runs of
             homozygosity (ROHs) detectable by single-nucleotide
             polymorphism (SNP) arrays. OBJECTIVE: To evaluate the
             association between ROHs and AD in an African American
             population known to have a risk for AD up to 3 times higher
             than white individuals. DESIGN, SETTING, AND PARTICIPANTS:
             Case-control study of a large African American data set
             previously genotyped on different genome-wide SNP arrays
             conducted from December 2013 to January 2015. Global and
             locus-based ROH measurements were analyzed using raw or
             imputed genotype data. We studied the raw genotypes from 2
             case-control subsets grouped based on SNP array: Alzheimer's
             Disease Genetics Consortium data set (871 cases and 1620
             control individuals) and Chicago Health and Aging
             Project-Indianapolis Ibadan Dementia Study data set (279
             cases and 1367 control individuals). We then examined the
             entire data set using imputed genotypes from 1917 cases and
             3858 control individuals. MAIN OUTCOMES AND MEASURES: The
             ROHs larger than 1 Mb, 2 Mb, or 3 Mb were investigated
             separately for global burden evaluation, consensus regions,
             and gene-based analyses. RESULTS: The African American
             cohort had a low degree of inbreeding (F ~ 0.006). In
             the Alzheimer's Disease Genetics Consortium data set, we
             detected a significantly higher proportion of cases with
             ROHs greater than 2 Mb (P = .004) or greater than 3 Mb
             (P = .02), as well as a significant 114-kilobase
             consensus region on chr4q31.3 (empirical P value
             2 = .04; ROHs >2 Mb). In the Chicago Health and Aging
             Project-Indianapolis Ibadan Dementia Study data set, we
             identified a significant 202-kilobase consensus region on
             Chr15q24.1 (empirical P value 2 = .02; ROHs >1 Mb) and a
             cluster of 13 significant genes on Chr3p21.31 (empirical P
             value 2 = .03; ROHs >3 Mb). A total of 43 of 49
             nominally significant genes common for both data sets also
             mapped to Chr3p21.31. Analyses of imputed SNP data from the
             entire data set confirmed the association of AD with global
             ROH measurements (12.38 ROHs >1 Mb in cases vs 12.11 in
             controls; 2.986 Mb average size of ROHs >2 Mb in cases vs
             2.889 Mb in controls; and 22% of cases with ROHs >3 Mb vs
             19% of controls) and a gene-cluster on Chr3p21.31 (empirical
             P value 2 = .006-.04; ROHs >3 Mb). Also, we detected a
             significant association between AD and CLDN17 (empirical P
             value 2 = .01; ROHs >1 Mb), encoding a protein from the
             Claudin family, members of which were previously suggested
             as AD biomarkers. CONCLUSIONS AND RELEVANCE: To our
             knowledge, we discovered the first evidence of increased
             burden of ROHs among patients with AD from an outbred
             African American population, which could reflect either the
             cumulative effect of multiple ROHs to AD or the contribution
             of specific loci harboring recessive mutations and risk
             haplotypes in a subset of patients. Sequencing is required
             to uncover AD variants in these individuals.},
   Doi = {10.1001/jamaneurol.2015.1700},
   Key = {fds322226}
}

@article{fds325786,
   Author = {Cao, L and Pokorney, SD and Hayden, K and Welsh-Bohmer, K and Newby,
             LK},
   Title = {Cognitive Function: Is There More to Anticoagulation in
             Atrial Fibrillation Than Stroke?},
   Journal = {J Am Heart Assoc},
   Volume = {4},
   Number = {8},
   Pages = {e001573},
   Year = {2015},
   Month = {August},
   url = {http://dx.doi.org/10.1161/JAHA.114.001573},
   Doi = {10.1161/JAHA.114.001573},
   Key = {fds325786}
}

@article{fds276902,
   Author = {Mistridis, P and Egli, SC and Iverson, GL and Berres, M and Willmes, K and Welsh-Bohmer, KA and Monsch, AU},
   Title = {Considering the base rates of low performance in cognitively
             healthy older adults improves the accuracy to identify
             neurocognitive impairment with the Consortium to Establish a
             Registry for Alzheimer's Disease-Neuropsychological
             Assessment Battery (CERAD-NAB).},
   Journal = {Eur Arch Psychiatry Clin Neurosci},
   Volume = {265},
   Number = {5},
   Pages = {407-417},
   Year = {2015},
   Month = {August},
   ISSN = {0940-1334},
   url = {http://dx.doi.org/10.1007/s00406-014-0571-z},
   Abstract = {It is common for some healthy older adults to obtain low
             test scores when a battery of neuropsychological tests is
             administered, which increases the risk of the clinician
             misdiagnosing cognitive impairment. Thus, base rates of
             healthy individuals' low scores are required to more
             accurately interpret neuropsychological results. At present,
             this information is not available for the German version of
             the Consortium to Establish a Registry for Alzheimer's
             Disease-Neuropsychological Assessment Battery (CERAD-NAB), a
             frequently used battery in the USA and in German-speaking
             Europe. This study aimed to determine the base rates of low
             scores for the CERAD-NAB and to tabulate a summary figure of
             cut-off scores and numbers of low scores to aid in clinical
             decision making. The base rates of low scores on the ten
             German CERAD-NAB subscores were calculated from the German
             CERAD-NAB normative sample (N = 1,081) using six different
             cut-off scores (i.e., 1st, 2.5th, 7th, 10th, 16th, and 25th
             percentile). Results indicate that high percentages of one
             or more "abnormal" scores were obtained, irrespective of the
             cut-off criterion. For example, 60.6% of the normative
             sample obtained one or more scores at or below the 10th
             percentile. These findings illustrate the importance of
             considering the prevalence of low scores in healthy
             individuals. The summary figure of CERAD-NAB base rates is
             an important supplement for test interpretation and can be
             used to improve the diagnostic accuracy of neurocognitive
             disorders.},
   Doi = {10.1007/s00406-014-0571-z},
   Key = {fds276902}
}

@article{fds327852,
   Author = {Østergaard, SD and Mukherjee, S and Sharp, SJ and Proitsi, P and Lotta,
             LA and Day, F and Perry, JRB and Boehme, KL and Walter, S and Kauwe, JS and Gibbons, LE and Alzheimer’s Disease Genetics Consortium, and GERAD1 Consortium, and EPIC-InterAct Consortium, and Larson, EB and Powell, JF and Langenberg, C and Crane, PK and Wareham, NJ and Scott,
             RA},
   Title = {Associations between Potentially Modifiable Risk Factors and
             Alzheimer Disease: A Mendelian Randomization
             Study.},
   Journal = {PLoS Med},
   Volume = {12},
   Number = {6},
   Pages = {e1001841},
   Year = {2015},
   Month = {June},
   url = {http://dx.doi.org/10.1371/journal.pmed.1001841},
   Abstract = {BACKGROUND: Potentially modifiable risk factors including
             obesity, diabetes, hypertension, and smoking are associated
             with Alzheimer disease (AD) and represent promising targets
             for intervention. However, the causality of these
             associations is unclear. We sought to assess the causal
             nature of these associations using Mendelian randomization
             (MR). METHODS AND FINDINGS: We used SNPs associated with
             each risk factor as instrumental variables in MR analyses.
             We considered type 2 diabetes (T2D, NSNPs = 49), fasting
             glucose (NSNPs = 36), insulin resistance (NSNPs = 10), body
             mass index (BMI, NSNPs = 32), total cholesterol (NSNPs =
             73), HDL-cholesterol (NSNPs = 71), LDL-cholesterol (NSNPs =
             57), triglycerides (NSNPs = 39), systolic blood pressure
             (SBP, NSNPs = 24), smoking initiation (NSNPs = 1), smoking
             quantity (NSNPs = 3), university completion (NSNPs = 2), and
             years of education (NSNPs = 1). We calculated MR estimates
             of associations between each exposure and AD risk using an
             inverse-variance weighted approach, with summary statistics
             of SNP-AD associations from the International Genomics of
             Alzheimer's Project, comprising a total of 17,008
             individuals with AD and 37,154 cognitively normal elderly
             controls. We found that genetically predicted higher SBP was
             associated with lower AD risk (odds ratio [OR] per standard
             deviation [15.4 mm Hg] of SBP [95% CI]: 0.75 [0.62-0.91]; p
             = 3.4 × 10(-3)). Genetically predicted higher SBP was also
             associated with a higher probability of taking
             antihypertensive medication (p = 6.7 × 10(-8)). Genetically
             predicted smoking quantity was associated with lower AD risk
             (OR per ten cigarettes per day [95% CI]: 0.67 [0.51-0.89]; p
             = 6.5 × 10(-3)), although we were unable to stratify by
             smoking history; genetically predicted smoking initiation
             was not associated with AD risk (OR = 0.70 [0.37, 1.33]; p =
             0.28). We saw no evidence of causal associations between
             glycemic traits, T2D, BMI, or educational attainment and
             risk of AD (all p > 0.1). Potential limitations of this
             study include the small proportion of intermediate trait
             variance explained by genetic variants and other implicit
             limitations of MR analyses. CONCLUSIONS: Inherited lifetime
             exposure to higher SBP is associated with lower AD risk.
             These findings suggest that higher blood pressure--or some
             environmental exposure associated with higher blood
             pressure, such as use of antihypertensive medications--may
             reduce AD risk.},
   Doi = {10.1371/journal.pmed.1001841},
   Key = {fds327852}
}

@article{fds276903,
   Author = {Wang, L-S and Naj, AC and Graham, RR and Crane, PK and Kunkle, BW and Cruchaga, C and Murcia, JDG and Cannon-Albright, L and Baldwin, CT and Zetterberg, H and Blennow, K and Kukull, WA and Faber, KM and Schupf, N and Norton, MC and Tschanz, JT and Munger, RG and Corcoran, CD and Rogaeva,
             E and Alzheimer's Disease Genetics Consortium, and Lin, C-F and Dombroski, BA and Cantwell, LB and Partch, A and Valladares, O and Hakonarson, H and St George-Hyslop and P and Green, RC and Goate, AM and Foroud, TM and Carney, RM and Larson, EB and Behrens, TW and Kauwe, JSK and Haines, JL and Farrer, LA and Pericak-Vance, MA and Mayeux, R and Schellenberg, GD and National Institute on Aging-Late-Onset
             Alzheimer’s Disease (NIA-LOAD) Family Study, and Albert, MS and Albin, RL and Apostolova, LG and Arnold, SE and Barber, R and Barmada,
             M and Barnes, LL and Beach, TG and Becker, JT and Beecham, GW and Beekly,
             D and Bennett, DA and Bigio, EH and Bird, TD and Blacker, D and Boeve, BF and Bowen, JD and Boxer, A and Burke, JR and Buxbaum, JD and Cairns, NJ and Cao, C and Carlson, CS and Carroll, SL and Chui, HC and Clark, DG and Cribbs, DH and Crocco, EA and DeCarli, C and DeKosky, ST and Demirci,
             FY and Dick, M and Dickson, DW and Duara, R and Ertekin-Taner, N and Fallon, KB and Farlow, MR and Ferris, S and Frosch, MP and Galasko, DR and Ganguli, M and Gearing, M and Geschwind, DH and Ghetti, B and Gilbert,
             JR and Glass, JD and Graff-Radford, NR and Growdon, JH and Hamilton, RL and Hamilton-Nelson, KL and Harrell, LE and Head, E and Honig, LS and Hulette, CM and Hyman, BT and Jarvik, GP and Jicha, GA and Jin, L-W and Jun, G and Kamboh, MI and Karydas, A and Kaye, JA and Kim, R and Koo, EH and Kowall, NW and Kramer, JH and LaFerla, FM and Lah, JJ and Leverenz, JB and Levey, AI and Li, G and Lieberman, AP and Lopez, OL and Lunetta, KL and Lyketsos, CG and Mack, WJ and Marson, DC and Martin, ER and Martiniuk,
             F and Mash, DC and Masliah, E and McCormick, WC and McCurry, SM and McDavid, AN and McKee, AC and Mesulam, WM and Miller, BL and Miller, CA and Miller, JW and Montine, TJ and Morris, JC and Murrell, JR and Olichney,
             JM and Parisi, JE and Perry, W and Peskind, E and Petersen, RC and Pierce,
             A and Poon, WW and Potter, H and Quinn, JF and Raj, A and Raskind, M and Reiman, EM and Reisberg, B and Reitz, C and Ringman, JM and Roberson,
             ED and Rosen, HJ and Rosenberg, RN and Sano, M and Saykin, AJ and Schneider, JA and Schneider, LS and Seeley, WW and Smith, AG and Sonnen,
             JA and Spina, S and Stern, RA and Tanzi, RE and Thornton-Wells, TA and Trojanowski, JQ and Troncoso, JC and Tsuang, DW and Van Deerlin and VM and Van Eldik and LJ and Vardarajan, BN and Vinters, HV and Vonsattel, JP and Weintraub, S and Welsh-Bohmer, KA and Williamson, J and Wishnek, S and Woltjer, RL and Wright, CB and Younkin, SG and Yu, C-E and Yu,
             L},
   Title = {Rarity of the Alzheimer disease-protective APP A673T variant
             in the United States.},
   Journal = {JAMA Neurol},
   Volume = {72},
   Number = {2},
   Pages = {209-216},
   Year = {2015},
   Month = {February},
   ISSN = {2168-6149},
   url = {http://dx.doi.org/10.1001/jamaneurol.2014.2157},
   Abstract = {IMPORTANCE: Recently, a rare variant in the amyloid
             precursor protein gene (APP) was described in a population
             from Iceland. This variant, in which alanine is replaced by
             threonine at position 673 (A673T), appears to protect
             against late-onset Alzheimer disease (AD). We evaluated the
             frequency of this variant in AD cases and cognitively normal
             controls to determine whether this variant will
             significantly contribute to risk assessment in individuals
             in the United States. OBJECTIVE: To determine the frequency
             of the APP A673T variant in a large group of elderly
             cognitively normal controls and AD cases from the United
             States and in 2 case-control cohorts from Sweden. DESIGN,
             SETTING, AND PARTICIPANTS: Case-control association analysis
             of variant APP A673T in US and Swedish white individuals
             comparing AD cases with cognitively intact elderly controls.
             Participants were ascertained at multiple
             university-associated medical centers and clinics across the
             United States and Sweden by study-specific sampling methods.
             They were from case-control studies, community-based
             prospective cohort studies, and studies that ascertained
             multiplex families from multiple sources. MAIN OUTCOMES AND
             MEASURES: Genotypes for the APP A673T variant were
             determined using the Infinium HumanExome V1 Beadchip
             (Illumina, Inc) and by TaqMan genotyping (Life
             Technologies). RESULTS: The A673T variant genotypes were
             evaluated in 8943 US AD cases, 10 480 US cognitively
             normal controls, 862 Swedish AD cases, and 707 Swedish
             cognitively normal controls. We identified 3 US individuals
             heterozygous for A673T, including 1 AD case (age at onset,
             89 years) and 2 controls (age at last examination, 82 and 77
             years). The remaining US samples were homozygous for the
             alanine (A673) allele. In the Swedish samples, 3 controls
             were heterozygous for A673T and all AD cases were homozygous
             for the A673 allele. We also genotyped a US family
             previously reported to harbor the A673T variant and found a
             mother-daughter pair, both cognitively normal at ages 72 and
             84 years, respectively, who were both heterozygous for
             A673T; however, all individuals with AD in the family were
             homozygous for A673. CONCLUSIONS AND RELEVANCE: The A673T
             variant is extremely rare in US cohorts and does not play a
             substantial role in risk for AD in this population. This
             variant may be primarily restricted to Icelandic and
             Scandinavian populations.},
   Doi = {10.1001/jamaneurol.2014.2157},
   Key = {fds276903}
}

@article{fds371166,
   Author = {Budur, K and Martenyi, F and Welsh-Bohmer, KA and Burns, DK and Chiang,
             C and O'Neil, J and Runyan, G and Schuster, J and Crenshaw, DG and Lutz,
             MW and Metz, CA and Saunders, AM and Yarbrough, D and Yarnall, D and Lai,
             E and Brannan, SK and Roses, AD},
   Title = {A Pharmacogenetics Supported Clinical Trial to Delay Onset
             of Mild Cognitive Impairment due to Alzheimer's Disease
             Using Low Dose Pioglitazone: The Tommorrow
             Study},
   Journal = {NEUROPSYCHOPHARMACOLOGY},
   Volume = {39},
   Pages = {S342-S342},
   Publisher = {NATURE PUBLISHING GROUP},
   Year = {2014},
   Month = {December},
   Key = {fds371166}
}

@article{fds276898,
   Author = {Chuang, Y-F and Breitner, JCS and Chiu, Y-L and Khachaturian, A and Hayden, K and Corcoran, C and Tschanz, J and Norton, M and Munger, R and Welsh-Bohmer, K and Zandi, PP and Cache County
             Investigators},
   Title = {Use of diuretics is associated with reduced risk of
             Alzheimer's disease: the Cache County Study.},
   Journal = {Neurobiol Aging},
   Volume = {35},
   Number = {11},
   Pages = {2429-2435},
   Year = {2014},
   Month = {November},
   ISSN = {0197-4580},
   url = {http://dx.doi.org/10.1016/j.neurobiolaging.2014.05.002},
   Abstract = {Although the use of antihypertensive medications has been
             associated with reduced risk of Alzheimer's disease (AD), it
             remains unclear which class provides the most benefit. The
             Cache County Study of Memory Health and Aging is a
             prospective longitudinal cohort study of dementing illnesses
             among the elderly population of Cache County, Utah. Using
             waves I to IV data of the Cache County Study, 3417
             participants had a mean of 7.1 years of follow-up.
             Time-varying use of antihypertensive medications including
             different class of diuretics, angiotensin converting enzyme
             inhibitors, β-blockers, and calcium channel blockers was
             used to predict the incidence of AD using Cox proportional
             hazards analyses. During follow-up, 325 AD cases were
             ascertained with a total of 23,590 person-years. Use of any
             antihypertensive medication was associated with lower
             incidence of AD (adjusted hazard ratio [aHR], 0.77; 95%
             confidence interval [CI], 0.61-0.97). Among different
             classes of antihypertensive medications, thiazide (aHR, 0.7;
             95% CI, 0.53-0.93), and potassium-sparing diuretics (aHR,
             0.69; 95% CI, 0.48-0.99) were associated with the greatest
             reduction of AD risk. Thiazide and potassium-sparing
             diuretics were associated with decreased risk of AD. The
             inverse association of potassium-sparing diuretics confirms
             an earlier finding in this cohort, now with longer
             follow-up, and merits further investigation.},
   Doi = {10.1016/j.neurobiolaging.2014.05.002},
   Key = {fds276898}
}

@article{fds276904,
   Author = {Naj, AC and Jun, G and Reitz, C and Kunkle, BW and Perry, W and Park, YS and Beecham, GW and Rajbhandary, RA and Hamilton-Nelson, KL and Wang,
             L-S and Kauwe, JSK and Huentelman, MJ and Myers, AJ and Bird, TD and Boeve,
             BF and Baldwin, CT and Jarvik, GP and Crane, PK and Rogaeva, E and Barmada,
             MM and Demirci, FY and Cruchaga, C and Kramer, PL and Ertekin-Taner, N and Hardy, J and Graff-Radford, NR and Green, RC and Larson, EB and St
             George-Hyslop, PH and Buxbaum, JD and Evans, DA and Schneider, JA and Lunetta, KL and Kamboh, MI and Saykin, AJ and Reiman, EM and De Jager,
             PL and Bennett, DA and Morris, JC and Montine, TJ and Goate, AM and Blacker, D and Tsuang, DW and Hakonarson, H and Kukull, WA and Foroud,
             TM and Martin, ER and Haines, JL and Mayeux, RP and Farrer, LA and Schellenberg, GD and Pericak-Vance, MA and Alzheimer Disease
             Genetics Consortium, and Albert, MS and Albin, RL and Apostolova,
             LG and Arnold, SE and Barber, R and Barnes, LL and Beach, TG and Becker,
             JT and Beekly, D and Bigio, EH and Bowen, JD and Boxer, A and Burke, JR and Cairns, NJ and Cantwell, LB and Cao, C and Carlson, CS and Carney, RM and Carrasquillo, MM and Carroll, SL and Chui, HC and Clark, DG and Corneveaux, J and Cribbs, DH and Crocco, EA and DeCarli, C and DeKosky,
             ST and Dick, M and Dickson, DW and Duara, R and Faber, KM and Fallon, KB and Farlow, MR and Ferris, S and Frosch, MP and Galasko, DR and Ganguli, M and Gearing, M and Geschwind, DH and Ghetti, B and Gilbert, JR and Glass,
             JD and Growdon, JH and Hamilton, RL and Harrell, LE and Head, E and Honig,
             LS and Hulette, CM and Hyman, BT and Jicha, GA and Jin, L-W and Karydas, A and Kaye, JA and Kim, R and Koo, EH and Kowall, NW and Kramer, JH and LaFerla,
             FM and Lah, JJ and Leverenz, JB and Levey, AI and Li, G and Lieberman, AP and Lin, C-F and Lopez, OL and Lyketsos, CG and Mack, WJ and Martiniuk, F and Mash, DC and Masliah, E and McCormick, WC and McCurry, SM and McDavid,
             AN and McKee, AC and Mesulam, M and Miller, BL and Miller, CA and Miller,
             JW and Murrell, JR and Olichney, JM and Pankratz, VS and Parisi, JE and Paulson, HL and Peskind, E and Petersen, RC and Pierce, A and Poon, WW and Potter, H and Quinn, JF and Raj, A and Raskind, M and Reisberg, B and Ringman, JM and Roberson, ED and Rosen, HJ and Rosenberg, RN and Sano,
             M and Schneider, LS and Seeley, WW and Smith, AG and Sonnen, JA and Spina,
             S and Stern, RA and Tanzi, RE and Thornton-Wells, TA and Trojanowski,
             JQ and Troncoso, JC and Valladares, O and Van Deerlin and VM and Van Eldik,
             LJ and Vardarajan, BN and Vinters, HV and Vonsattel, JP and Weintraub,
             S and Welsh-Bohmer, KA and Williamson, J and Wishnek, S and Woltjer, RL and Wright, CB and Younkin, SG and Yu, C-E and Yu, L},
   Title = {Effects of multiple genetic loci on age at onset in
             late-onset Alzheimer disease: a genome-wide association
             study.},
   Journal = {JAMA Neurol},
   Volume = {71},
   Number = {11},
   Pages = {1394-1404},
   Year = {2014},
   Month = {November},
   ISSN = {2168-6149},
   url = {http://dx.doi.org/10.1001/jamaneurol.2014.1491},
   Abstract = {IMPORTANCE: Because APOE locus variants contribute to risk
             of late-onset Alzheimer disease (LOAD) and to differences in
             age at onset (AAO), it is important to know whether other
             established LOAD risk loci also affect AAO in affected
             participants. OBJECTIVES: To investigate the effects of
             known Alzheimer disease risk loci in modifying AAO and to
             estimate their cumulative effect on AAO variation using data
             from genome-wide association studies in the Alzheimer
             Disease Genetics Consortium. DESIGN, SETTING, AND
             PARTICIPANTS: The Alzheimer Disease Genetics Consortium
             comprises 14 case-control, prospective, and family-based
             data sets with data on 9162 participants of white
             race/ethnicity with Alzheimer disease occurring after age 60
             years who also had complete AAO information, gathered
             between 1989 and 2011 at multiple sites by participating
             studies. Data on genotyped or imputed single-nucleotide
             polymorphisms most significantly associated with risk at 10
             confirmed LOAD loci were examined in linear modeling of AAO,
             and individual data set results were combined using a
             random-effects, inverse variance-weighted meta-analysis
             approach to determine whether they contribute to variation
             in AAO. Aggregate effects of all risk loci on AAO were
             examined in a burden analysis using genotype scores weighted
             by risk effect sizes. MAIN OUTCOMES AND MEASURES: Age at
             disease onset abstracted from medical records among
             participants with LOAD diagnosed per standard criteria.
             RESULTS: Analysis confirmed the association of APOE with
             earlier AAO (P = 3.3 × 10(-96)), with associations in CR1
             (rs6701713, P = 7.2 × 10(-4)), BIN1 (rs7561528, P = 4.8 ×
             10(-4)), and PICALM (rs561655, P = 2.2 × 10(-3)) reaching
             statistical significance (P < .005). Risk alleles
             individually reduced AAO by 3 to 6 months. Burden analyses
             demonstrated that APOE contributes to 3.7% of the variation
             in AAO (R(2) = 0.256) over baseline (R(2) = 0.221), whereas
             the other 9 loci together contribute to 2.2% of the
             variation (R(2) = 0.242). CONCLUSIONS AND RELEVANCE: We
             confirmed an association of APOE (OMIM 107741) variants with
             AAO among affected participants with LOAD and observed novel
             associations of CR1 (OMIM 120620), BIN1 (OMIM 601248), and
             PICALM (OMIM 603025) with AAO. In contrast to earlier
             hypothetical modeling, we show that the combined effects of
             Alzheimer disease risk variants on AAO are on the scale of,
             but do not exceed, the APOE effect. While the aggregate
             effects of risk loci on AAO may be significant, additional
             genetic contributions to AAO are individually likely to be
             small.},
   Doi = {10.1001/jamaneurol.2014.1491},
   Key = {fds276904}
}

@article{fds276906,
   Author = {Roses, AD and Lutz, MW and Saunders, AM and Goldgaber, D and Saul, R and Sundseth, SS and Akkari, PA and Roses, SM and Gottschalk, WK and Whitfield, KE and Vostrov, AA and Hauser, MA and Allingham, RR and Burns, DK and Chiba-Falek, O and Welsh-Bohmer,
             KA},
   Title = {African-American TOMM40'523-APOE haplotypes are admixture of
             West African and Caucasian alleles.},
   Journal = {Alzheimers Dement},
   Volume = {10},
   Number = {6},
   Pages = {592-601.e2},
   Year = {2014},
   Month = {November},
   ISSN = {1552-5260},
   url = {http://dx.doi.org/10.1016/j.jalz.2014.06.009},
   Abstract = {BACKGROUND: Several studies have demonstrated a lower
             apolipoprotein E4 (APOE ε4) allele frequency in
             African-Americans, but yet an increased age-related
             prevalence of AD. An algorithm for prevention clinical
             trials incorporating TOMM40'523 (Translocase of Outer
             Mitochondria Membrane) and APOE depends on accurate
             TOMM40'523-APOE haplotypes. METHODS: We have compared the
             APOE and TOMM40'523 phased haplotype frequencies of a 9.5 kb
             TOMM40/APOE genomic region in West African, Caucasian, and
             African-American cohorts. RESULTS: African-American
             haplotype frequency scans of poly-T lengths connected in
             phase with either APOE ε4 or APOE ε3 differ from both West
             Africans and Caucasians and represent admixture of several
             distinct West African and Caucasian haplotypes. A new West
             African TOMM40'523 haplotype, with APOE ε4 connected to a
             short TOMM40'523 allele, is observed in African-Americans
             but not Caucasians. CONCLUSION: These data have therapeutic
             implications for the age of onset risk algorithm estimates
             and the design of a prevention trial for African-Americans
             or other mixed ethnic populations.},
   Doi = {10.1016/j.jalz.2014.06.009},
   Key = {fds276906}
}

@article{fds276907,
   Author = {Hayden, KM and Makeeva, OA and Newby, LK and Plassman, BL and Markova,
             VV and Dunham, A and Romero, HR and Melikyan, ZA and Germain, CM and Welsh-Bohmer, KA and Roses, AD and TOMSK-DUKE Study Group
             Investigators},
   Title = {A comparison of neuropsychological performance between US
             and Russia: preparing for a global clinical
             trial.},
   Journal = {Alzheimers Dement},
   Volume = {10},
   Number = {6},
   Pages = {760-768.e1},
   Year = {2014},
   Month = {November},
   ISSN = {1552-5260},
   url = {http://dx.doi.org/10.1016/j.jalz.2014.02.008},
   Abstract = {BACKGROUND: Understanding regional differences in cognitive
             performance is important for interpretation of data from
             large multinational clinical trials. METHODS: Data from
             Durham and Cabarrus Counties in North Carolina, USA and
             Tomsk, Russia (n = 2972) were evaluated. The Montreal
             Cognitive Assessment (MoCA), Trail Making Test Part B
             (Trails B), Consortium to Establish a Registry for
             Alzheimer's Disease Word List Memory Test (WLM) delayed
             recall, and self-report Alzheimer's Disease Cooperative
             Studies Mail-In Cognitive Function Screening Instrument
             (MCFSI) were administered at each site. Multilevel modeling
             measured the variance explained by site and predictors of
             cognitive performance. RESULTS: Site differences accounted
             for 11% of the variation in the MoCA, 1.6% in Trails B, 1.7%
             in WLM, and 0.8% in MCFSI scores. Prior memory testing was
             significantly associated with WLM. Diabetes and stroke were
             significantly associated with Trails B and MCFSI.
             CONCLUSIONS: Sources of variation include cultural
             differences, health conditions, and exposure to test
             stimuli. Findings highlight the importance of local norms to
             interpret test performance.},
   Doi = {10.1016/j.jalz.2014.02.008},
   Key = {fds276907}
}

@article{fds276916,
   Author = {Hayden, KM and Kuchibhatla, M and Romero, HR and Plassman, BL and Burke,
             JR and Browndyke, JN and Welsh-Bohmer, KA},
   Title = {Pre-clinical cognitive phenotypes for Alzheimer disease: a
             latent profile approach.},
   Journal = {Am J Geriatr Psychiatry},
   Volume = {22},
   Number = {11},
   Pages = {1364-1374},
   Year = {2014},
   Month = {November},
   url = {http://www.ncbi.nlm.nih.gov/pubmed/24080384},
   Abstract = {BACKGROUND: Cognitive profiles for pre-clinical Alzheimer
             disease (AD) can be used to identify groups of individuals
             at risk for disease and better characterize pre-clinical
             disease. Profiles or patterns of performance as pre-clinical
             phenotypes may be more useful than individual test scores or
             measures of global decline. OBJECTIVE: To evaluate patterns
             of cognitive performance in cognitively normal individuals
             to derive latent profiles associated with later onset of
             disease using a combination of factor analysis and latent
             profile analysis. METHODS: The National Alzheimer
             Coordinating Centers collect data, including a battery of
             neuropsychological tests, from participants at 29 National
             Institute on Aging-funded Alzheimer Disease Centers across
             the United States. Prior factor analyses of this battery
             demonstrated a four-factor structure comprising memory,
             attention, language, and executive function. Factor scores
             from these analyses were used in a latent profile approach
             to characterize cognition among a group of cognitively
             normal participants (N = 3,911). Associations between
             latent profiles and disease outcomes an average of 3 years
             later were evaluated with multinomial regression models.
             Similar analyses were used to determine predictors of
             profile membership. RESULTS: Four groups were identified;
             each with distinct characteristics and significantly
             associated with later disease outcomes. Two groups were
             significantly associated with development of cognitive
             impairment. In post hoc analyses, both the Trail Making Test
             Part B, and a contrast score (Delayed Recall - Trails B),
             significantly predicted group membership and later cognitive
             impairment. CONCLUSIONS: Latent profile analysis is a useful
             method to evaluate patterns of cognition in large samples
             for the identification of preclinical AD phenotypes;
             comparable results, however, can be achieved with very
             sensitive tests and contrast scores.},
   Doi = {10.1016/j.jagp.2013.07.008},
   Key = {fds276916}
}

@article{fds276919,
   Author = {Greene, D and Tschanz, JT and Smith, KR and Ostbye, T and Corcoran, C and Welsh-Bohmer, KA and Norton, MC and Cache County
             Investigators},
   Title = {Impact of offspring death on cognitive health in late life:
             the Cache County study.},
   Journal = {Am J Geriatr Psychiatry},
   Volume = {22},
   Number = {11},
   Pages = {1307-1315},
   Year = {2014},
   Month = {November},
   url = {http://www.ncbi.nlm.nih.gov/pubmed/23954042},
   Abstract = {OBJECTIVE: Experiencing the death of a child is associated
             with negative short-term mental health consequences, but
             less is known about cognitive outcomes and whether such
             associations extend to late life. We tested the hypothesis
             that experiencing an offspring death (OD) is associated with
             an increased rate of cognitive decline in late life.
             METHODS: This population-based longitudinal study observed
             four cognitive statuses spaced 3-4 years apart, linked to an
             extensive database containing objective genealogic and vital
             statistics data. Home visits were conducted with 3,174
             residents of a rural county in northern Utah, initially
             without dementia, aged 65-105. Cognitive status was measured
             with the Modified Mini-Mental State Exam at baseline and at
             3-, 7-, and 10-year follow-ups. OD was obtained from the
             Utah Population Database, which contains statewide birth and
             death records. RESULTS: In linear mixed models, controlling
             for age, gender, education, and apolipoprotein E status,
             subjects who experienced OD while younger than age 31 years
             experienced a significantly faster rate of cognitive decline
             in late life, but only if they had an ε4 allele.
             Reclassifying all OD (regardless of age) according to
             subsequent birth of another child, OD was only related to
             faster cognitive decline when there were no subsequent
             births. CONCLUSION: Experiencing OD in early adulthood has a
             long-term association with cognitive functioning in late
             life, with a gene-environment interaction at the
             apolipoprotein E locus. Subsequent birth of another child
             attenuates this association.},
   Doi = {10.1016/j.jagp.2013.05.002},
   Key = {fds276919}
}

@article{fds276910,
   Author = {Linnertz, C and Lutz, MW and Ervin, JF and Allen, J and Miller, NR and Welsh-Bohmer, KA and Roses, AD and Chiba-Falek,
             O},
   Title = {The genetic contributions of SNCA and LRRK2 genes to Lewy
             Body pathology in Alzheimer's disease.},
   Journal = {Hum Mol Genet},
   Volume = {23},
   Number = {18},
   Pages = {4814-4821},
   Year = {2014},
   Month = {September},
   ISSN = {0964-6906},
   url = {http://dx.doi.org/10.1093/hmg/ddu196},
   Abstract = {The molecular genetic basis that leads to Lewy Body (LB)
             pathology in 15-20% of Alzheimer disease cases (LBV/AD) was
             largely unknown. Alpha-synuclein (SNCA) and Leucine-rich
             repeat kinase2 (LRRK2) have been implicated in the
             pathogenesis of Parkinson's disease (PD), the prototype of
             LB spectrum disorders. We tested the association of SNCA
             variants with LB pathology in AD. We then stratified the
             SNCA association analyses by LRRK2 genotype. We also
             investigated the expression regulation of SNCA and LRRK2 in
             relation to LB pathology. We evaluated the differences in
             SNCA-mRNA and LRRK2-mRNA levels as a function of LB
             pathology in the temporal cortex (TC) from autopsy-confirmed
             LBV/AD cases and AD controls. We further investigated the
             cis-effect of the LB pathology-associated genetic variants
             within the SNCA and LRRK2 loci on the mRNA expression of
             these genes. SNCA SNPs rs3857059 and rs2583988 showed
             significant associations with increased risk for LB
             pathology. When the analyses were stratified by
             LRRK2-rs1491923 genotype, the associations became stronger
             for both SNPs and an association was also observed with
             rs2619363. Expression analysis demonstrated that SNCA- and
             LRRK2-mRNA levels were significantly higher in TC from
             LBV/AD brains compared with AD controls. Furthermore,
             SNCA-mRNA expression level in the TC was associated with
             rs3857059; homozygotes for the minor allele showed
             significant higher expression. LRRK2-transcript levels were
             increased in carriers of rs1491923 minor allele. Our
             findings demonstrated that SNCA contributes to LB pathology
             in AD patients, possibly via interaction with LRRK2, and
             suggested that expression regulation of these genes may be
             the molecular basis underlying the observed LB
             associations.},
   Doi = {10.1093/hmg/ddu196},
   Key = {fds276910}
}

@article{fds276913,
   Author = {Linnertz, C and Anderson, L and Gottschalk, W and Crenshaw, D and Lutz,
             MW and Allen, J and Saith, S and Mihovilovic, M and Burke, JR and Welsh-Bohmer, KA and Roses, AD and Chiba-Falek,
             O},
   Title = {The cis-regulatory effect of an Alzheimer's
             disease-associated poly-T locus on expression of TOMM40 and
             apolipoprotein E genes.},
   Journal = {Alzheimers Dement},
   Volume = {10},
   Number = {5},
   Pages = {541-551},
   Year = {2014},
   Month = {September},
   ISSN = {1552-5260},
   url = {http://dx.doi.org/10.1016/j.jalz.2013.08.280},
   Abstract = {BACKGROUND: We investigated the genomic region spanning the
             Translocase of the Outer Mitochondrial Membrane 40-kD
             (TOMM40) and Apolipoprotein E (APOE) genes, that has been
             associated with the risk and age of onset of late-onset
             Alzheimer's disease (LOAD) to determine whether a highly
             polymorphic, intronic poly-T within this region (rs10524523;
             hereafter, 523) affects expression of the APOE and TOMM40
             genes. Alleles of this locus are classified as S, short; L,
             long; and VL, very long based on the number of T residues.
             METHODS: We evaluated differences in APOE messenger RNA
             (mRNA) and TOMM40 mRNA levels as a function of the 523
             genotype in two brain regions from APOE ε3/ε3 white
             autopsy-confirmed LOAD cases and normal controls. We further
             investigated the effect of the 523 locus in its native
             genomic context using a luciferase expression system.
             RESULTS: The expression of both genes was significantly
             increased with disease. Mean expression of APOE and TOMM40
             mRNA levels were higher in VL homozygotes compared with S
             homozygotes in the temporal and occipital cortexes from
             normal and LOAD cases. Results of a luciferase reporter
             system were consistent with the human brain mRNA analysis;
             the 523 VL poly-T resulted in significantly higher
             expression than the S poly-T. Although the effect of poly-T
             length on reporter expression was the same in HepG2 hepatoma
             and SH-SY5Y neuroblastoma cells, the magnitude of the effect
             was greater in the neuroblastoma than in the hepatoma cells,
             which implies tissue-specific modulation of the 523 poly-T.
             CONCLUSIONS: These results suggest that the 523 locus may
             contribute to LOAD susceptibility by modulating the
             expression of TOMM40 and/or APOE transcription.},
   Doi = {10.1016/j.jalz.2013.08.280},
   Key = {fds276913}
}

@article{fds276922,
   Author = {Peterson, D and Munger, C and Crowley, J and Corcoran, C and Cruchaga,
             C and Goate, AM and Norton, MC and Green, RC and Munger, RG and Breitner,
             JCS and Welsh-Bohmer, KA and Lyketsos, C and Tschanz, J and Kauwe, JSK and Alzheimer's Disease Neuroimaging Initiative},
   Title = {Variants in PPP3R1 and MAPT are associated with more rapid
             functional decline in Alzheimer's disease: the Cache County
             Dementia Progression Study.},
   Journal = {Alzheimers Dement},
   Volume = {10},
   Number = {3},
   Pages = {366-371},
   Year = {2014},
   Month = {May},
   url = {http://www.ncbi.nlm.nih.gov/pubmed/23727081},
   Abstract = {BACKGROUND: Single-nucleotide polymorphisms (SNPs) located
             in the gene encoding the regulatory subunit of the protein
             phosphatase 2B (PPP3R1, rs1868402) and the
             microtubule-associated protein tau (MAPT, rs3785883) gene
             were recently associated with higher cerebrospinal fluid
             (CSF) tau levels in samples from the Knight Alzheimer's
             Disease Research Center at Washington University (WU) and
             Alzheimer's Disease Neuroimaging Initiative (ADNI). In these
             same samples, these SNPs were also associated with faster
             functional decline, or progression of Alzheimer's disease
             (AD) as measured by the Clinical Dementia Rating sum of
             boxes scores (CDR-sb). We attempted to validate the latter
             association in an independent, population-based sample of
             incident AD cases from the Cache County Dementia Progression
             Study (DPS). METHODS: All 92 AD cases from the DPS with a
             global CDR-sb ≤1 (mild) at initial clinical assessment who
             were later assessed on CDR-sb data on at least two other
             time points were genotyped at the two SNPs of interest
             (rs1868402 and rs3785883). We used linear mixed models to
             estimate associations between these SNPs and CDR-sb
             trajectory. All analyses were performed using Proc Mixed in
             SAS. RESULTS: Although we observed no association between
             rs3785883 or rs1868402 alone and change in CDR-sb (P > .10),
             there was a significant association between a combined
             genotype model and change in CDR-sb: carriers of the
             high-risk genotypes at both loci progressed >2.9 times
             faster than noncarriers (P = .015). When data from DPS were
             combined with previously published data from WU and ADNI,
             change in CDR-sb was 30% faster for each copy of the
             high-risk allele at rs3785883 (P = .0082) and carriers of
             both high-risk genotypes at both loci progressed 6 times
             faster (P < .0001) than all others combined. CONCLUSIONS: We
             replicate a previous report by Cruchaga et al that specific
             variations in rs3785883 and rs1868402 are associated with
             accelerated progression of AD. Further characterization of
             this association will provide a better understanding of how
             genetic factors influence the rate of progression of AD and
             could provide novel insights into preventative and
             therapeutic strategies.},
   Doi = {10.1016/j.jalz.2013.02.010},
   Key = {fds276922}
}

@article{fds276920,
   Author = {Steinberg, M and Hess, K and Corcoran, C and Mielke, MM and Norton, M and Breitner, J and Green, R and Leoutsakos, J and Welsh-Bohmer, K and Lyketsos, C and Tschanz, J},
   Title = {Vascular risk factors and neuropsychiatric symptoms in
             Alzheimer's disease: the Cache County Study.},
   Journal = {Int J Geriatr Psychiatry},
   Volume = {29},
   Number = {2},
   Pages = {153-159},
   Year = {2014},
   Month = {February},
   ISSN = {0885-6230},
   url = {http://dx.doi.org/10.1002/gps.3980},
   Abstract = {OBJECTIVE: Knowledge of potentially modifiable risk factors
             for neuropsychiatric symptoms (NPS) in Alzheimer's disease
             (AD) is important. This study longitudinally explores
             modifiable vascular risk factors for NPS in AD. METHODS:
             Participants enrolled in the Cache County Study on Memory in
             Aging with no dementia at baseline were subsequently
             assessed over three additional waves, and those with
             incident (new onset) dementia were invited to join the
             Dementia Progression Study for longitudinal follow-up. A
             total of 327 participants with incident AD were identified
             and assessed for the following vascular factors: atrial
             fibrillation, hypertension, diabetes mellitus, angina,
             coronary artery bypass surgery, myocardial infarction,
             cerebrovascular accident, and use of antihypertensive or
             diabetes medicines. A vascular index (VI) was also
             calculated. NPS were assessed over time using the
             Neuropsychiatric Inventory (NPI). Affective and Psychotic
             symptom clusters were assessed separately. The association
             between vascular factors and change in NPI total score was
             analyzed using linear mixed model and in symptom clusters
             using a random effects model. RESULTS: No individual
             vascular risk factors or the VI significantly predicted
             change in any individual NPS. The use of antihypertensive
             medications more than four times per week was associated
             with higher total NPI and Affective cluster scores.
             CONCLUSIONS: Use of antihypertensive medication was
             associated with higher total NPI and Affective cluster
             scores. The results of this study do not otherwise support
             vascular risk factors as modifiers of longitudinal change in
             NPS in AD.},
   Doi = {10.1002/gps.3980},
   Key = {fds276920}
}

@article{fds276911,
   Author = {Romero, HR and Welsh-Bohmer, KA and Gwyther, LP and Edmonds, HL and Plassman, BL and Germain, CM and McCart, M and Hayden, KM and Pieper, C and Roses, AD},
   Title = {Community engagement in diverse populations for Alzheimer
             disease prevention trials.},
   Journal = {Alzheimer Dis Assoc Disord},
   Volume = {28},
   Number = {3},
   Pages = {269-274},
   Year = {2014},
   ISSN = {0893-0341},
   url = {http://dx.doi.org/10.1097/WAD.0000000000000029},
   Abstract = {The recruitment of asymptomatic volunteers has been
             identified as a critical factor that is delaying the
             development and validation of preventive therapies for
             Alzheimer disease (AD). Typical recruitment strategies
             involve the use of convenience samples or soliciting
             participation of older adults with a family history of AD
             from clinics and outreach efforts. However, high-risk
             groups, such as ethnic/racial minorities, are traditionally
             less likely to be recruited for AD prevention studies, thus
             limiting the ability to generalize findings for a
             significant proportion of the aging population. A
             community-engagement approach was used to create a registry
             of 2311 research-ready, healthy adult volunteers who reflect
             the ethnically diverse local community. Furthermore, the
             registry's actual commitment to research was examined,
             through demonstrated participation rates in a clinical
             study. The approach had varying levels of success in
             establishing a large, diverse pool of individuals who are
             interested in participating in pharmacological prevention
             trials and meet the criteria for primary prevention research
             trials designed to delay the onset of AD. Our efforts
             suggest that entry criteria for the clinical trials need to
             be carefully considered to be inclusive of African
             Americans, and that sustained effort is needed to engage
             African Americans in pharmacological prevention
             approaches.},
   Doi = {10.1097/WAD.0000000000000029},
   Key = {fds276911}
}

@article{fds276917,
   Author = {Wengreen, H and Munger, RG and Cutler, A and Quach, A and Bowles, A and Corcoran, C and Tschanz, JT and Norton, MC and Welsh-Bohmer,
             KA},
   Title = {Prospective study of Dietary Approaches to Stop
             Hypertension- and Mediterranean-style dietary patterns and
             age-related cognitive change: the Cache County Study on
             Memory, Health and Aging.},
   Journal = {Am J Clin Nutr},
   Volume = {98},
   Number = {5},
   Pages = {1263-1271},
   Year = {2013},
   Month = {November},
   url = {http://www.ncbi.nlm.nih.gov/pubmed/24047922},
   Abstract = {BACKGROUND: Healthy dietary patterns may protect against
             age-related cognitive decline, but results of studies have
             been inconsistent. OBJECTIVE: We examined associations
             between Dietary Approaches to Stop Hypertension (DASH)- and
             Mediterranean-style dietary patterns and age-related
             cognitive change in a prospective, population-based study.
             DESIGN: Participants included 3831 men and women ≥65 y of
             age who were residents of Cache County, UT, in 1995.
             Cognitive function was assessed by using the Modified
             Mini-Mental State Examination (3MS) ≤4 times over 11 y.
             Diet-adherence scores were computed by summing across the
             energy-adjusted rank-order of individual food and nutrient
             components and categorizing participants into quintiles of
             the distribution of the diet accordance score. Mixed-effects
             repeated-measures models were used to examine 3MS scores
             over time across increasing quintiles of dietary accordance
             scores and individual food components that comprised each
             score. RESULTS: The range of rank-order DASH and
             Mediterranean diet scores was 1661-25,596 and 2407-26,947,
             respectively. Higher DASH and Mediterranean diet scores were
             associated with higher average 3MS scores. People in
             quintile 5 of DASH averaged 0.97 points higher than those in
             quintile 1 (P = 0.001). The corresponding difference for
             Mediterranean quintiles was 0.94 (P = 0.001). These
             differences were consistent over 11 y. Higher intakes of
             whole grains and nuts and legumes were also associated with
             higher average 3MS scores [mean quintile 5 compared with 1
             differences: 1.19 (P < 0.001), 1.22 (P < 0.001),
             respectively]. CONCLUSIONS: Higher levels of accordance with
             both the DASH and Mediterranean dietary patterns were
             associated with consistently higher levels of cognitive
             function in elderly men and women over an 11-y period. Whole
             grains and nuts and legumes were positively associated with
             higher cognitive functions and may be core neuroprotective
             foods common to various healthy plant-centered diets around
             the globe.},
   Doi = {10.3945/ajcn.112.051276},
   Key = {fds276917}
}

@article{fds276924,
   Author = {Peters, ME and Rosenberg, PB and Steinberg, M and Norton, MC and Welsh-Bohmer, KA and Hayden, KM and Breitner, J and Tschanz, JT and Lyketsos, CG and Cache County Investigators},
   Title = {Neuropsychiatric symptoms as risk factors for progression
             from CIND to dementia: the Cache County Study.},
   Journal = {Am J Geriatr Psychiatry},
   Volume = {21},
   Number = {11},
   Pages = {1116-1124},
   Year = {2013},
   Month = {November},
   url = {http://www.ncbi.nlm.nih.gov/pubmed/23567370},
   Abstract = {OBJECTIVES: To examine the association of neuropsychiatric
             symptom (NPS) severity with risk of transition to all-cause
             dementia, Alzheimer disease (AD), and vascular dementia
             (VaD). DESIGN: Survival analysis of time to dementia, AD, or
             VaD onset. SETTING: Population-based study. PARTICIPANTS:
             230 participants diagnosed with cognitive impairment, no
             dementia (CIND) from the Cache County Study of Memory Health
             and Aging were followed for a mean of 3.3 years.
             MEASUREMENTS: The Neuropsychiatric Inventory (NPI) was used
             to quantify the presence, frequency, and severity of NPS.
             Chi-squared statistics, t-tests, and Cox proportional hazard
             ratios were used to assess associations. RESULTS: The
             conversion rate from CIND to all-cause dementia was 12% per
             year, with risk factors including an APOE ε4 allele, lower
             Mini-Mental State Examination, lower 3MS, and higher CDR
             sum-of-boxes. The presence of at least one NPS was a risk
             factor for all-cause dementia, as was the presence of NPS
             with mild severity. Nighttime behaviors were a risk factor
             for all-cause dementia and of AD, whereas hallucinations
             were a risk factor for VaD. CONCLUSIONS: These data confirm
             that NPS are risk factors for conversion from CIND to
             dementia. Of special interest is that even NPS of mild
             severity are a risk for all-cause dementia or
             AD.},
   Doi = {10.1016/j.jagp.2013.01.049},
   Key = {fds276924}
}

@article{fds277148,
   Author = {Tschanz, JT and Pfister, R and Wanzek, J and Corcoran, C and Smith, K and Tschanz, BT and Steffens, DC and Østbye, T and Welsh-Bohmer, KA and Norton, MC},
   Title = {Stressful life events and cognitive decline in late life:
             moderation by education and age. The Cache County
             Study.},
   Journal = {Int J Geriatr Psychiatry},
   Volume = {28},
   Number = {8},
   Pages = {821-830},
   Year = {2013},
   Month = {August},
   url = {http://www.ncbi.nlm.nih.gov/pubmed/23037866},
   Abstract = {OBJECTIVE: Stressful life events (SLE) have been associated
             with increased dementia risk, but their association with
             cognitive decline has been inconsistent. In a longitudinal
             population-based study of older individuals, we examined the
             association between SLE and cognitive decline, and the role
             of potential effect modifiers. METHODS: A total of 2665
             non-demented participants of the Cache County Memory Study
             completed an SLE questionnaire at Wave 2 and were revisited
             4 and 7 years later. The events were represented via
             several scores: total number, subjective rating (negative,
             positive, and unexpected), and a weighted summary based on
             their impact. Cognition was assessed at each visit with the
             modified Mini-Mental State Exam. General linear models were
             used to examine the association between SLE scores and
             cognition. Effect modification by age, education, and APOE
             genotype was tested. RESULTS: Years of formal education
             (p = 0.006) modified the effect of number of SLE, and
             age (p = 0.009) modified the effect of negative SLE on
             the rate of cognitive decline. Faster decline was observed
             among those with fewer years of education experiencing more
             SLE and also among younger participants experiencing more
             negative SLE. There was no association between other
             indicators of SLE and cognitive decline. APOE genotype did
             not modify any of the aforementioned associations.
             CONCLUSIONS: The effects of SLE on cognition in late life
             are complex and vary by individual factors such as age and
             education. These results may explain some of the
             contradictory findings in the literature.},
   Doi = {10.1002/gps.3888},
   Key = {fds277148}
}

@article{fds277104,
   Author = {Browndyke, JN and Giovanello, K and Petrella, J and Hayden, K and Chiba-Falek, O and Tucker, KA and Burke, JR and Welsh-Bohmer,
             KA},
   Title = {Phenotypic regional functional imaging patterns during
             memory encoding in mild cognitive impairment and Alzheimer's
             disease.},
   Journal = {Alzheimers Dement},
   Volume = {9},
   Number = {3},
   Pages = {284-294},
   Year = {2013},
   Month = {May},
   url = {http://www.ncbi.nlm.nih.gov/pubmed/22841497},
   Abstract = {BACKGROUND: Reliable blood-oxygen-level-dependent (BOLD)
             functional magnetic resonance imaging (fMRI) phenotypic
             biomarkers of Alzheimer's disease (AD) or mild cognitive
             impairment (MCI) are likely to emerge only from a
             systematic, quantitative, and aggregate examination of the
             functional neuroimaging research literature. METHODS: A
             series of random-effects activation likelihood estimation
             (ALE) meta-analyses were conducted on studies of episodic
             memory encoding operations in AD and MCI samples relative to
             normal controls. ALE analyses were based on a thorough
             literature search for all task-based functional neuroimaging
             studies in AD and MCI published up to January 2010. Analyses
             covered 16 fMRI studies, which yielded 144 distinct foci for
             ALE meta-analysis. RESULTS: ALE results indicated several
             regional task-based BOLD consistencies in MCI and AD
             patients relative to normal control subjects across the
             aggregate BOLD functional neuroimaging research literature.
             Patients with AD and those at significant risk (MCI) showed
             statistically significant consistent activation differences
             during episodic memory encoding in the medial temporal lobe,
             specifically parahippocampal gyrus, as well superior frontal
             gyrus, precuneus, and cuneus, relative to normal control
             subjects. CONCLUSIONS: ALE consistencies broadly support the
             presence of frontal compensatory activity, medial temporal
             lobe activity alteration, and posterior midline "default
             mode" hyperactivation during episodic memory encoding
             attempts in the diseased or prospective predisease
             condition. Taken together, these robust commonalities may
             form the foundation for a task-based fMRI phenotype of
             memory encoding in AD.},
   Doi = {10.1016/j.jalz.2011.12.006},
   Key = {fds277104}
}

@article{fds371167,
   Author = {Welsh-Bohmer, KA and Fillenbaum, GG and Morris,
             JC},
   Title = {Albert Heyman, MD (1916–2012)},
   Journal = {Neurology},
   Volume = {80},
   Number = {13},
   Pages = {1184-1185},
   Publisher = {Ovid Technologies (Wolters Kluwer Health)},
   Year = {2013},
   Month = {March},
   url = {http://dx.doi.org/10.1212/wnl.0b013e3182897160},
   Doi = {10.1212/wnl.0b013e3182897160},
   Key = {fds371167}
}

@article{fds276925,
   Author = {Potter, GG and Wagner, HR and Burke, JR and Plassman, BL and Welsh-Bohmer, KA and Steffens, DC},
   Title = {Neuropsychological predictors of dementia in late-life major
             depressive disorder.},
   Journal = {Am J Geriatr Psychiatry},
   Volume = {21},
   Number = {3},
   Pages = {297-306},
   Year = {2013},
   Month = {March},
   url = {http://www.ncbi.nlm.nih.gov/pubmed/23395197},
   Abstract = {OBJECTIVE: Major depressive disorder is a likely risk factor
             for dementia, but some cases of major depressive disorder in
             older adults may actually represent a prodrome of this
             condition. The purpose of this study was to use
             neuropsychological test scores to predict conversion to
             dementia in a sample of depressed older adults diagnosed as
             nondemented at the time of neuropsychological testing.
             DESIGN: Longitudinal, with mean follow-up of 5.45 years.
             SETTING: Outpatient depression treatment study at Duke
             University. PARTICIPANTS: Thirty nondemented individuals
             depressed at the time of neuropsychological testing and
             later diagnosed with incident dementia; 149 nondemented
             individuals depressed at the time of neuropsychological
             testing and a diagnosis of cognitively normal. METHODOLOGY:
             All participants received clinical assessment of depression,
             were assessed to rule out prevalent dementia at the time of
             study enrollment, completed neuropsychological testing at
             the time of study enrollment, and were diagnosed for
             cognitive disorders on an annual basis. RESULTS:
             Nondemented, acutely depressed older adults who converted to
             dementia during the study period exhibited broadly lower
             cognitive performances at baseline than acutely depressed
             individuals who remained cognitively normal. Discriminant
             function analysis indicated that 2 neuropsychological tests,
             Recognition Memory (from the Consortium to Establish a
             Registry for Alzheimer's Disease neuropsychological battery)
             and Trail Making B, best predicted dementia conversion.
             CONCLUSIONS: Depressed older adults with cognitive deficits
             in the domains of memory and executive functions during
             acute depression are at higher risk for developing dementia.
             Some cases of late-life depression may reflect a prodrome of
             dementia in which clinical manifestation of mood changes may
             co-occur with emerging cognitive deficits.},
   Doi = {10.1016/j.jagp.2012.12.009},
   Key = {fds276925}
}

@article{fds276929,
   Author = {Crenshaw, DG and Gottschalk, WK and Lutz, MW and Grossman, I and Saunders, AM and Burke, JR and Welsh-Bohmer, KA and Brannan, SK and Burns, DK and Roses, AD},
   Title = {Using genetics to enable studies on the prevention of
             Alzheimer's disease.},
   Journal = {Clin Pharmacol Ther},
   Volume = {93},
   Number = {2},
   Pages = {177-185},
   Year = {2013},
   Month = {February},
   url = {http://www.ncbi.nlm.nih.gov/pubmed/23249780},
   Abstract = {Curing Alzheimer's disease (AD) remains an elusive goal;
             indeed, it may even prove to be impossible, given the nature
             of the disease. Although modulating disease progression is
             an attractive target and will alleviate the burden of the
             most severe stages, this strategy will not reduce the
             prevalence of the disease itself. Preventing or (as
             described in this article) delaying the onset of cognitive
             impairment and AD will provide the greatest benefit to
             individuals and society by pushing the onset of disease into
             the later years of life. Because of the high variability in
             the age of onset of the disease, AD prevention studies that
             do not stratify participants by age-dependent disease risk
             will be operationally challenging, being large in size and
             of long duration. We present a composite genetic biomarker
             to stratify disease risk so as to facilitate clinical
             studies in high-risk populations. In addition, we discuss
             the rationale for the use of pioglitazone to delay the onset
             of AD in individuals at high risk.},
   Doi = {10.1038/clpt.2012.222},
   Key = {fds276929}
}

@article{fds371168,
   Author = {Foster, C and Addis, D and Ford, J and Kaufer, D and Browndyke, J and Welsh-Bohmer, K and Giovanello, K},
   Title = {PREFRONTAL CONTRIBUTIONS TO RELATIONAL ENCODING IN HEALTHY
             AGING AND MILD COGNITIVE IMPAIRMENT},
   Journal = {JOURNAL OF COGNITIVE NEUROSCIENCE},
   Pages = {197-197},
   Publisher = {MIT PRESS},
   Year = {2013},
   Month = {January},
   Key = {fds371168}
}

@article{fds277115,
   Author = {Blumenthal, JA and Smith, PJ and Welsh-Bohmer, K and Babyak, MA and Browndyke, J and Lin, P-H and Doraiswamy, PM and Burke, J and Kraus, W and Hinderliter, A and Sherwood, A},
   Title = {Can lifestyle modification improve neurocognition? Rationale
             and design of the ENLIGHTEN clinical trial.},
   Journal = {Contemp Clin Trials},
   Volume = {34},
   Number = {1},
   Pages = {60-69},
   Year = {2013},
   Month = {January},
   url = {http://www.ncbi.nlm.nih.gov/pubmed/23000080},
   Abstract = {BACKGROUND: Risk factors for cardiovascular disease (CVD)
             not only increase the risk for clinical CVD events, but also
             are associated with a cascade of neurophysiologic and
             neuroanatomic changes that increase the risk of cognitive
             impairment and dementia. Although epidemiological studies
             have shown that exercise and diet are associated with lower
             CVD risk and reduced incidence of dementia, no randomized
             controlled trial (RCT) has examined the independent effects
             of exercise and diet on neurocognitive function among
             individuals at risk for dementia. The ENLIGHTEN trial is a
             RCT of patients with CVD risk factors who also are
             characterized by subjective cognitive complaints and
             objective evidence of neurocognitive impairment without
             dementia (CIND) STUDY DESIGN: A 2 by 2 design will examine
             the independent and combined effects of diet and exercise on
             neurocognition. 160 participants diagnosed with CIND will be
             randomly assigned to 6 months of aerobic exercise, the DASH
             diet, or a combination of both exercise and diet; a
             (control) group will receive health education but otherwise
             will maintain their usual dietary and activity habits.
             Participants will complete comprehensive assessments of
             neurocognitive functioning along with biomarkers of CVD risk
             including measures of blood pressure, glucose, endothelial
             function, and arterial stiffness. CONCLUSION: The ENLIGHTEN
             trial will (a) evaluate the effectiveness of aerobic
             exercise and the DASH diet in improving neurocognitive
             functioning in CIND patients with CVD risk factors; (b)
             examine possible mechanisms by which exercise and diet
             improve neurocognition; and (c) consider potential
             moderators of treatment, including subclinical
             CVD.},
   Doi = {10.1016/j.cct.2012.09.004},
   Key = {fds277115}
}

@article{fds276923,
   Author = {Hendrix, SB and Welsh-Bohmer, KA},
   Title = {Separation of cognitive domains to improve prediction of
             progression from mild cognitive impairment to Alzheimer's
             disease.},
   Journal = {Alzheimers Res Ther},
   Volume = {5},
   Number = {3},
   Pages = {22},
   Year = {2013},
   url = {http://www.ncbi.nlm.nih.gov/pubmed/23680123},
   Abstract = {Addressing causes of heterogeneity in cognitive outcomes is
             becoming more critical as Alzheimer's disease (AD) research
             focuses on earlier disease. One of the causes of this
             heterogeneity may be that individuals with deficiencies in
             different cognitive domains may perform similarly on a
             neuropsychological (NP) test for very different reasons.
             Tatsuoka and colleagues have applied a Bayesian model in
             order to integrate knowledge about cognitive domains
             relevant to each NP test with the observed outcomes from the
             Alzheimer's Disease Neuroimaging Initiative (ADNI) mild
             cognitive impairment data. This approach resulted in better
             prediction of AD diagnosis than more traditional
             approaches.},
   Doi = {10.1186/alzrt176},
   Key = {fds276923}
}

@article{fds276927,
   Author = {Whitcomb, DC and LaRusch, J and Krasinskas, AM and Klei, L and Smith,
             JP and Brand, RE and Neoptolemos, JP and Lerch, MM and Tector, M and Sandhu, BS and Guda, NM and Orlichenko, L and Alzheimer's Disease
             Genetics Consortium, and Alkaade, S and Amann, ST and Anderson, MA and Baillie, J and Banks, PA and Conwell, D and Coté, GA and Cotton, PB and DiSario, J and Farrer, LA and Forsmark, CE and Johnstone, M and Gardner,
             TB and Gelrud, A and Greenhalf, W and Haines, JL and Hartman, DJ and Hawes,
             RA and Lawrence, C and Lewis, M and Mayerle, J and Mayeux, R and Melhem,
             NM and Money, ME and Muniraj, T and Papachristou, GI and Pericak-Vance,
             MA and Romagnuolo, J and Schellenberg, GD and Sherman, S and Simon, P and Singh, VP and Slivka, A and Stolz, D and Sutton, R and Weiss, FU and Wilcox, CM and Zarnescu, NO and Wisniewski, SR and O'Connell, MR and Kienholz, ML and Roeder, K and Barmada, MM and Yadav, D and Devlin,
             B},
   Title = {Common genetic variants in the CLDN2 and PRSS1-PRSS2 loci
             alter risk for alcohol-related and sporadic
             pancreatitis.},
   Journal = {Nat Genet},
   Volume = {44},
   Number = {12},
   Pages = {1349-1354},
   Year = {2012},
   Month = {December},
   url = {http://www.ncbi.nlm.nih.gov/pubmed/23143602},
   Abstract = {Pancreatitis is a complex, progressively destructive
             inflammatory disorder. Alcohol was long thought to be the
             primary causative agent, but genetic contributions have been
             of interest since the discovery that rare PRSS1, CFTR and
             SPINK1 variants were associated with pancreatitis risk. We
             now report two associations at genome-wide significance
             identified and replicated at PRSS1-PRSS2 (P < 1 × 10(-12))
             and X-linked CLDN2 (P < 1 × 10(-21)) through a two-stage
             genome-wide study (stage 1: 676 cases and 4,507 controls;
             stage 2: 910 cases and 4,170 controls). The PRSS1 variant
             likely affects disease susceptibility by altering expression
             of the primary trypsinogen gene. The CLDN2 risk allele is
             associated with atypical localization of claudin-2 in
             pancreatic acinar cells. The homozygous (or hemizygous in
             males) CLDN2 genotype confers the greatest risk, and its
             alleles interact with alcohol consumption to amplify risk.
             These results could partially explain the high frequency of
             alcohol-related pancreatitis in men (male hemizygote
             frequency is 0.26, whereas female homozygote frequency is
             0.07).},
   Doi = {10.1038/ng.2466},
   Key = {fds276927}
}

@article{fds276972,
   Author = {Rosenberg, PB and Mielke, MM and Han, D and Leoutsakos, JS and Lyketsos,
             CG and Rabins, PV and Zandi, PP and Breitner, JCS and Norton, MC and Welsh-Bohmer, KA and Zuckerman, IH and Rattinger, GB and Green, RC and Corcoran, C and Tschanz, JT},
   Title = {The association of psychotropic medication use with the
             cognitive, functional, and neuropsychiatric trajectory of
             Alzheimer's disease.},
   Journal = {Int J Geriatr Psychiatry},
   Volume = {27},
   Number = {12},
   Pages = {1248-1257},
   Year = {2012},
   Month = {December},
   url = {http://www.ncbi.nlm.nih.gov/pubmed/22374884},
   Abstract = {OBJECTIVE: The use of psychotropic medications in
             Alzheimer's disease (AD) has been associated with both
             deleterious and potentially beneficial outcomes. We examined
             the longitudinal association of psychotropic medication use
             with cognitive, functional, and neuropsychiatric symptom
             (NPS) trajectories among community-ascertained incident AD
             cases from the Cache County Dementia Progression Study.
             METHODS: A total of 230 participants were followed for a
             mean of 3.7 years. Persistency index (PI) was calculated for
             all antidepressants, selective serotonin reuptake inhibitors
             (SSRIs), antipsychotics (atypical and typical), and
             benzodiazepines as the proportion of observed time of
             medication exposure. Mixed-effects models were used to
             examine the association between PI for each medication class
             and Mini-Mental State Exam (MMSE), Clinical Dementia Rating
             Sum of Boxes (CDR-Sum), and Neuropsychiatric Inventory -
             Total (NPI-Total) trajectories, controlling for appropriate
             demographic and clinical covariates. RESULTS: At baseline,
             psychotropic medication use was associated with greater
             severity of dementia and poorer medical status. Higher PI
             for all medication classes was associated with a more rapid
             decline in MMSE. For antidepressant, SSRI, benzodiazepine,
             and typical antipsychotic use, a higher PI was associated
             with a more rapid increase in CDR-Sum. For SSRIs,
             antipsychotics, and typical antipsychotics, a higher PI was
             associated with more rapid increase in NPI-Total.
             CONCLUSIONS: Psychotropic medication use was associated with
             more rapid cognitive and functional decline in AD, and not
             with improved NPS. Clinicians may tend to prescribe
             psychotropic medications to AD patients at risk of poorer
             outcomes, but one cannot rule out the possibility of poorer
             outcomes being caused by psychotropic medications.},
   Doi = {10.1002/gps.3769},
   Key = {fds276972}
}

@article{fds277070,
   Author = {Sheline, YI and Disabato, BM and Hranilovich, J and Morris, C and D'Angelo, G and Pieper, C and Toffanin, T and Taylor, WD and MacFall,
             JR and Wilkins, C and Barch, DM and Welsh-Bohmer, KA and Steffens, DC and Krishnan, RR and Doraiswamy, PM},
   Title = {Treatment course with antidepressant therapy in late-life
             depression.},
   Journal = {Am J Psychiatry},
   Volume = {169},
   Number = {11},
   Pages = {1185-1193},
   Year = {2012},
   Month = {November},
   url = {http://www.ncbi.nlm.nih.gov/pubmed/23034601},
   Abstract = {OBJECTIVE: In order to assess the effect of gray matter
             volumes and cortical thickness on antidepressant treatment
             response in late-life depression, the authors examined the
             relationship between brain regions identified a priori and
             Montgomery-Åsberg Depression Rating Scale (MADRS) scores
             over the course of an antidepressant treatment trial.
             METHOD: In a nonrandomized prospective trial, 168 patients
             who were at least 60 years of age and met DSM-IV criteria
             for major depression underwent MRI and were enrolled in a
             12-week treatment study. Exclusion criteria included
             cognitive impairment or severe medical disorders. The
             volumes or cortical thicknesses of regions of interest that
             differed between the depressed group and a comparison group
             (N=50) were determined. These regions of interest were used
             in analyses of the depressed group to predict antidepressant
             treatment outcome. Mixed-model analyses adjusting for age,
             education, age at depression onset, race, baseline MADRS
             score, scanner, and interaction with time examined
             predictors of MADRS scores over time. RESULTS: Smaller
             hippocampal volumes predicted a slower response to
             treatment. With the inclusion of white matter
             hyper-intensity severity and neuropsychological factor
             scores, the best model included hippocampal volume and
             cognitive processing speed to predict rate of response over
             time. A secondary analysis showed that hippocampal volume
             and frontal pole thickness differed between patients who
             achieved remission and those who did not. CONCLUSIONS: These
             data expand our understanding of the prediction of treatment
             course in late-life depression. The authors propose that the
             primary variables of hippocampal volume and cognitive
             processing speed, subsuming other contributing variables
             (episodic memory, executive function, language processing)
             predict antidepressant response.},
   Doi = {10.1176/appi.ajp.2012.12010122},
   Key = {fds277070}
}

@article{fds277050,
   Author = {Shao, H and Breitner, JCS and Whitmer, RA and Wang, J and Hayden, K and Wengreen, H and Corcoran, C and Tschanz, J and Norton, M and Munger, R and Welsh-Bohmer, K and Zandi, PP and Cache County
             Investigators},
   Title = {Hormone therapy and Alzheimer disease dementia: new findings
             from the Cache County Study.},
   Journal = {Neurology},
   Volume = {79},
   Number = {18},
   Pages = {1846-1852},
   Year = {2012},
   Month = {October},
   ISSN = {0028-3878},
   url = {http://dx.doi.org/10.1212/WNL.0b013e318271f823},
   Abstract = {OBJECTIVES: Observational studies suggest reduced risk of
             Alzheimer disease (AD) in users of hormone therapy (HT), but
             trials show higher risk. We examined whether the association
             of HT with AD varies with timing or type of HT use. METHODS:
             Between 1995 and 2006, the population-based Cache County
             Study followed 1,768 women who had provided a detailed
             history on age at menopause and use of HT. During this
             interval, 176 women developed incident AD. Cox proportional
             hazard models evaluated the association of HT use with AD,
             overall and in relation to timing, duration of use, and type
             (opposed vs unopposed) of HT. RESULTS: Women who used any
             type of HT within 5 years of menopause had 30% less risk of
             AD (95% confidence interval 0.49-0.99), especially if use
             was for 10 or more years. By contrast, AD risk was not
             reduced among those who had initiated HT 5 or more years
             after menopause. Instead, rates were increased among those
             who began "opposed" estrogen-progestin compounds within the
             3 years preceding the Cache County Study baseline (adjusted
             hazard ratio 1.93; 95% confidence interval 0.94-3.96). This
             last hazard ratio was similar to the ratio of 2.05 reported
             in randomized trial participants assigned to opposed HT.
             CONCLUSIONS: Association of HT use and risk of AD may depend
             on timing of use. Although possibly beneficial if taken
             during a critical window near menopause, HT (especially
             opposed compounds) initiated in later life may be associated
             with increased risk. The relation of AD risk to timing and
             type of HT deserves further study.},
   Doi = {10.1212/WNL.0b013e318271f823},
   Key = {fds277050}
}

@article{fds276971,
   Author = {Leoutsakos, J-MS and Han, D and Mielke, MM and Forrester, SN and Tschanz, JT and Corcoran, CD and Green, RC and Norton, MC and Welsh-Bohmer, KA and Lyketsos, CG},
   Title = {Effects of general medical health on Alzheimer's
             progression: the Cache County Dementia Progression
             Study.},
   Journal = {Int Psychogeriatr},
   Volume = {24},
   Number = {10},
   Pages = {1561-1570},
   Year = {2012},
   Month = {October},
   url = {http://www.ncbi.nlm.nih.gov/pubmed/22687143},
   Abstract = {BACKGROUND: Several observational studies have suggested a
             link between health status and rate of decline among
             individuals with Alzheimer's disease (AD). We sought to
             quantify the relationship in a population-based study of
             incident AD, and to compare global comorbidity ratings to
             counts of comorbid conditions and medications as predictors
             of AD progression. METHODS: This was a case-only cohort
             study arising from a population-based longitudinal study of
             memory and aging, in Cache County, Utah. Participants
             comprised 335 individuals with incident AD followed for up
             to 11 years. Patient descriptors included sex, age,
             education, dementia duration at baseline, and APOE genotype.
             Measures of health status made at each visit included the
             General Medical Health Rating (GMHR), number of comorbid
             medical conditions, and number of non-psychiatric
             medications. Dementia outcomes included the Mini-Mental
             State Examination (MMSE), Clinical Dementia Rating - sum of
             boxes (CDR-sb), and the Neuropsychiatric Inventory (NPI).
             RESULTS: Health status tended to fluctuate over time within
             individuals. None of the baseline medical variables (GMHR,
             comorbidities, and non-psychiatric medications) was
             associated with differences in rates of decline in
             longitudinal linear mixed effects models. Over time, low
             GMHR ratings, but not comorbidities or medications, were
             associated with poorer outcomes (MMSE: β = -1.07 p = 0.01;
             CDR-sb: β = 1.79 p < 0.001; NPI: β = 4.57 p = 0.01).
             CONCLUSIONS: Given that time-varying GMHR, but not baseline
             GMHR, was associated with the outcomes, it seems likely that
             there is a dynamic relationship between medical and
             cognitive health. GMHR is a more sensitive measure of health
             than simple counts of comorbidities or medications. Since
             health status is a potentially modifiable risk factor,
             further study is warranted.},
   Doi = {10.1017/S104161021200049X},
   Key = {fds276971}
}

@article{fds276928,
   Author = {Peters, ME and Rosenberg, PB and Steinberg, M and Norton, MC and Welsh-Bohmer, KA and Hayden, KM and Breitner, J and Tschanz, JT and Lyketsos, CG and the Cache County Investigators},
   Title = {Neuropsychiatric Symptoms as Risk Factors for Progression
             From CIND to Dementia: The Cache County Study.},
   Journal = {Am J Geriatr Psychiatry},
   Year = {2012},
   Month = {September},
   url = {http://www.ncbi.nlm.nih.gov/pubmed/22976786},
   Abstract = {OBJECTIVES:: To examine the association of neuropsychiatric
             symptom (NPS) severity with risk of transition to all-cause
             dementia, Alzheimer disease (AD), and vascular dementia
             (VaD). DESIGN:: Survival analysis of time to dementia, AD,
             or VaD onset. SETTING:: Population-based study.
             PARTICIPANTS:: 230 participants diagnosed with cognitive
             impairment, no dementia (CIND) from the Cache County Study
             of Memory Health and Aging were followed for a mean of 3.3
             years. MEASUREMENTS:: The Neuropsychiatric Inventory (NPI)
             was used to quantify the presence, frequency, and severity
             of NPS. Chi-squared statistics, t-tests, and Cox
             proportional hazard ratios were used to assess associations.
             RESULTS:: The conversion rate from CIND to all-cause
             dementia was 12% per year, with risk factors including an
             APOE [Latin Small Letter Open E]4 allele, lower Mini-Mental
             State Examination, lower 3MS, and higher CDR sum-of-boxes.
             The presence of at least one NPS was a risk factor for
             all-cause dementia, as was the presence of NPS with mild
             severity. Nighttime behaviors were a risk factor for
             all-cause dementia and of AD, whereas hallucinations were a
             risk factor for VaD. CONCLUSIONS:: These data confirm that
             NPS are risk factors for conversion from CIND to dementia.
             Of special interest is that even NPS of mild severity are a
             risk for all-cause dementia or AD.},
   Doi = {10.1097/JGP.0b013e318267014b},
   Key = {fds276928}
}

@article{fds277108,
   Author = {Giovanello, KS and De Brigard and F and Hennessey Ford and J and Kaufer,
             DI and Burke, JR and Browndyke, JN and Welsh-Bohmer,
             KA},
   Title = {Event-related functional magnetic resonance imaging changes
             during relational retrieval in normal aging and amnestic
             mild cognitive impairment.},
   Journal = {J Int Neuropsychol Soc},
   Volume = {18},
   Number = {5},
   Pages = {886-897},
   Year = {2012},
   Month = {September},
   url = {http://www.ncbi.nlm.nih.gov/pubmed/22622022},
   Abstract = {The earliest cognitive deficits observed in amnestic mild
             cognitive impairment (aMCI) appear to center on memory tasks
             that require relational memory (RM), the ability to link or
             integrate unrelated pieces of information. RM impairments in
             aMCI likely reflect neural changes in the medial temporal
             lobe (MTL) and posterior parietal cortex (PPC). We tested
             the hypothesis that individuals with aMCI, as compared to
             cognitively normal (CN) controls, would recruit neural
             regions outside of the MTL and PPC to support relational
             memory. To this end, we directly compared the neural
             underpinnings of successful relational retrieval in aMCI and
             CN groups, using event-related functional magnetic resonance
             imaging (fMRI), holding constant the stimuli and encoding
             task. The fMRI data showed that the CN, compared to the
             aMCI, group activated left precuneus, left angular gyrus,
             right posterior cingulate, and right parahippocampal cortex
             during relational retrieval, while the aMCI group, relative
             to the CN group, activated superior temporal gyrus and
             supramarginal gyrus for this comparison. Such findings
             indicate an early shift in the functional neural
             architecture of relational retrieval in aMCI, and may prove
             useful in future studies aimed at capitalizing on
             functionally intact neural regions as targets for treatment
             and slowing of the disease course. (JINS, 2012, 18,
             1-12).},
   Doi = {10.1017/S1355617712000689},
   Key = {fds277108}
}

@article{fds277118,
   Author = {Hayden, KM and McEvoy, JM and Linnertz, C and Attix, D and Kuchibhatla,
             M and Saunders, AM and Lutz, MW and Welsh-Bohmer, KA and Roses, AD and Chiba-Falek, O},
   Title = {A homopolymer polymorphism in the TOMM40 gene contributes to
             cognitive performance in aging.},
   Journal = {Alzheimers Dement},
   Volume = {8},
   Number = {5},
   Pages = {381-388},
   Year = {2012},
   Month = {September},
   url = {http://www.ncbi.nlm.nih.gov/pubmed/22863908},
   Abstract = {INTRODUCTION: A highly polymorphic T homopolymer was
             recently found to be associated with late-onset Alzheimer's
             disease risk and age of onset. OBJECTIVE: To explore the
             effects of the polymorphic polyT tract (rs10524523, referred
             as '523') on cognitive performance in cognitively healthy
             elderly individuals. METHODS: One hundred eighty-one
             participants were recruited from local independent-living
             retirement communities. Informed consent was obtained, and
             participants completed demographic questionnaires, a
             conventional paper-and-pencil neuropsychological battery,
             and the computerized Cambridge Neuropsychological Test
             Automated Battery (CANTAB). Saliva samples were collected
             for determination of the TOMM40 '523' (S, L, VL) and the
             apolipoprotein E (APOE) (ɛ2, 3, 4) genotypes. From the
             initial sample of 181 individuals, 127 were eligible for the
             association analysis. Participants were divided into three
             groups based on '523' genotypes (S/S, S/L-S/VL, and
             L/L-L/VL-VL/VL). Generalized linear models were used to
             evaluate the association between the '523' genotypes and
             neuropsychological test performance. Analyses were adjusted
             for age, sex, education, depression, and APOE ɛ4 status. A
             planned subanalysis was undertaken to evaluate the
             association between '523' genotypes and test performance in
             a sample restricted to APOE ɛ3 homozygotes. RESULTS: The S
             homozygotes performed better, although not significantly,
             than the S/L-S/VL and the VL/L-L/VL-VL/VL genotype groups on
             measures associated with memory (CANTAB Paired Associates
             Learning, Verbal Recognition Memory free recall) and
             executive function (CANTAB measures of Intra-Extra
             Dimensional Set Shift). Follow-up analysis of APOE ɛ3
             homozygotes only showed that the S/S group performed
             significantly better than the S/VL group on measures of
             episodic memory (CANTAB Paired Associates Learning and
             Verbal Recognition Memory free recall), attention (CANTAB
             Rapid Visual Information Processing latency), and executive
             function (Digit Symbol Substitution). The S/S group
             performed marginally better than the VL/VL group on
             Intra-Extra Dimensional Set Shift. None of the associations
             remained significant after applying a Bonferroni correction
             for multiple testing. CONCLUSIONS: Results suggest important
             APOE-independent associations between the TOMM40 '523'
             polymorphism and specific cognitive domains of memory and
             executive control that are preferentially affected in
             early-stage Alzheimer's disease.},
   Doi = {10.1016/j.jalz.2011.10.005},
   Key = {fds277118}
}

@article{fds276926,
   Author = {Coppola, G and Chinnathambi, S and Lee, JJ and Dombroski, BA and Baker,
             MC and Soto-Ortolaza, AI and Lee, SE and Klein, E and Huang, AY and Sears,
             R and Lane, JR and Karydas, AM and Kenet, RO and Biernat, J and Wang, L-S and Cotman, CW and Decarli, CS and Levey, AI and Ringman, JM and Mendez, MF and Chui, HC and Le Ber and I and Brice, A and Lupton, MK and Preza, E and Lovestone, S and Powell, J and Graff-Radford, N and Petersen, RC and Boeve, BF and Lippa, CF and Bigio, EH and Mackenzie, I and Finger, E and Kertesz, A and Caselli, RJ and Gearing, M and Juncos, JL and Ghetti, B and Spina, S and Bordelon, YM and Tourtellotte, WW and Frosch, MP and Vonsattel, JPG and Zarow, C and Beach, TG and Albin, RL and Lieberman,
             AP and Lee, VM and Trojanowski, JQ and Van Deerlin and VM and Bird, TD and Galasko, DR and Masliah, E and White, CL and Troncoso, JC and Hannequin,
             D and Boxer, AL and Geschwind, MD and Kumar, S and Mandelkow, E-M and Wszolek, ZK and Uitti, RJ and Dickson, DW and Haines, JL and Mayeux, R and Pericak-Vance, MA and Farrer, LA and Alzheimer's Disease Genetics
             Consortium, and Ross, OA and Rademakers, R and Schellenberg, GD and Miller, BL and Mandelkow, E and Geschwind, DH},
   Title = {Evidence for a role of the rare p.A152T variant in MAPT in
             increasing the risk for FTD-spectrum and Alzheimer's
             diseases.},
   Journal = {Hum Mol Genet},
   Volume = {21},
   Number = {15},
   Pages = {3500-3512},
   Year = {2012},
   Month = {August},
   url = {http://www.ncbi.nlm.nih.gov/pubmed/22556362},
   Abstract = {Rare mutations in the gene encoding for tau (MAPT,
             microtubule-associated protein tau) cause frontotemporal
             dementia-spectrum (FTD-s) disorders, including FTD,
             progressive supranuclear palsy (PSP) and corticobasal
             syndrome, and a common extended haplotype spanning across
             the MAPT locus is associated with increased risk of PSP and
             Parkinson's disease. We identified a rare tau variant
             (p.A152T) in a patient with a clinical diagnosis of PSP and
             assessed its frequency in multiple independent series of
             patients with neurodegenerative conditions and controls, in
             a total of 15 369 subjects. Tau p.A152T significantly
             increases the risk for both FTD-s (n = 2139, OR = 3.0, CI:
             1.6-5.6, P = 0.0005) and Alzheimer's disease (AD) (n = 3345,
             OR = 2.3, CI: 1.3-4.2, P = 0.004) compared with 9047
             controls. Functionally, p.A152T (i) decreases the binding of
             tau to microtubules and therefore promotes microtubule
             assembly less efficiently; and (ii) reduces the tendency to
             form abnormal fibers. However, there is a pronounced
             increase in the formation of tau oligomers. Importantly,
             these findings suggest that other regions of the tau protein
             may be crucial in regulating normal function, as the p.A152
             residue is distal to the domains considered responsible for
             microtubule interactions or aggregation. These data provide
             both the first genetic evidence and functional studies
             supporting the role of MAPT p.A152T as a rare risk factor
             for both FTD-s and AD and the concept that rare variants can
             increase the risk for relatively common, complex
             neurodegenerative diseases, but since no clear significance
             threshold for rare genetic variation has been established,
             some caution is warranted until the findings are further
             replicated.},
   Doi = {10.1093/hmg/dds161},
   Key = {fds276926}
}

@article{fds277069,
   Author = {Barch, DM and DʼAngelo, G and Pieper, C and Wilkins, CH and Welsh-Bohmer, K and Taylor, W and Garcia, KS and Gersing, K and Doraiswamy, PM and Sheline, YI},
   Title = {Cognitive improvement following treatment in late-life
             depression: relationship to vascular risk and age of
             onset.},
   Journal = {Am J Geriatr Psychiatry},
   Volume = {20},
   Number = {8},
   Pages = {682-690},
   Year = {2012},
   Month = {August},
   url = {http://www.ncbi.nlm.nih.gov/pubmed/22430020},
   Abstract = {OBJECTIVES: To test the hypothesis that the degree of
             vascular burden and/or age of onset may influence the degree
             to which cognition can improve during the course of
             treatment in late-life depression. DESIGN: Measurement of
             cognition both before and following 12 weeks of treatment
             with sertraline. SETTING: University medical centers
             (Washington University and Duke University). PARTICIPANTS:
             One hundred sixty-six individuals with late-life depression.
             INTERVENTION: Sertraline treatment. MEASUREMENTS: The
             cognitive tasks were grouped into five domains (language,
             processing speed, working memory, episodic memory, and
             executive function). We measured vascular risk using the
             Framingham Stroke Risk Profile measure. We measured T2-based
             white matter hyperintensities using the Fazekas criteria.
             RESULTS: Both episodic memory and executive function
             demonstrated significant improvement among adults with
             late-life depression during treatment with sertraline.
             Importantly, older age, higher vascular risk scores, and
             lower baseline Mini-Mental State Examination scores
             predicted less change in working memory. Furthermore, older
             age, later age of onset, and higher vascular risk scores
             predicted less change in executive function. CONCLUSIONS:
             These results have important clinical implications in that
             they suggest that a regular assessment of vascular risk in
             older adults with depression is necessary as a component of
             treatment planning and in predicting prognosis, both for the
             course of the depression itself and for the cognitive
             impairments that often accompany depression in later
             life.},
   Doi = {10.1097/JGP.0b013e318246b6cb},
   Key = {fds277069}
}

@article{fds277096,
   Author = {McCarthy, JJ and Saith, S and Linnertz, C and Burke, JR and Hulette, CM and Welsh-Bohmer, KA and Chiba-Falek, O},
   Title = {The Alzheimer's associated 5' region of the SORL1 gene cis
             regulates SORL1 transcripts expression.},
   Journal = {Neurobiol Aging},
   Volume = {33},
   Number = {7},
   Pages = {1485.e1-1485.e8},
   Year = {2012},
   Month = {July},
   url = {http://www.ncbi.nlm.nih.gov/pubmed/21185108},
   Abstract = {SORL1 has been identified as a major contributor to late
             onset Alzheimer's disease (LOAD). We test whether genetic
             variability in the 5' of SORL1 gene modulates the risk to
             develop LOAD via regulation of SORL1-messenger ribonucleic
             acid (mRNA) expression and splicing. Two brain structures,
             differentially vulnerable to LOAD pathology, were examined
             in 144 brain samples from 92 neurologically normal
             individuals. The temporal cortex, which is more susceptible
             to Alzheimer's pathology, demonstrated ∼2-fold increase in
             SORL1-mRNA levels in carriers of the minor alleles at single
             nucleotide polymorphisms (SNPs), rs7945931 and rs2298525,
             compared with noncarriers. No genetic effect on
             total-SORL1-mRNA levels was detected in the frontal cortex.
             However, rs11600875 minor allele was associated with
             significantly increased levels of exon-2 skipping, but only
             in frontal cortex. No correlation of SORL1-mRNAs expression
             was found between frontal and temporal cortexes.
             Collectively, these indicate the brain region specificity of
             the genetic regulation of SORL1 expression. Our results
             suggest that genetic regulation of SORL1 expression plays a
             role in disease risk and may be responsible for the reported
             LOAD associations. Further studies to detect the actual
             pathogenic variant/s are necessary.},
   Doi = {10.1016/j.neurobiolaging.2010.10.004},
   Key = {fds277096}
}

@article{fds276970,
   Author = {Mielke, MM and Leoutsakos, J-M and Corcoran, CD and Green, RC and Norton, MC and Welsh-Bohmer, KA and Tschanz, JT and Lyketsos,
             CG},
   Title = {Effects of Food and Drug Administration-approved medications
             for Alzheimer's disease on clinical progression.},
   Journal = {Alzheimers Dement},
   Volume = {8},
   Number = {3},
   Pages = {180-187},
   Year = {2012},
   Month = {May},
   url = {http://www.ncbi.nlm.nih.gov/pubmed/22301194},
   Abstract = {BACKGROUND: Observational studies suggest that
             cholinesterase inhibitors and/or memantine may delay
             clinical progression of Alzheimer's disease (AD) in 40% of
             individuals taking the medications. Given this response and
             existence of side effects, we sought to quantify medication
             use and benefits in a population-based study of incident AD
             cases. METHODS: The Cache County Dementia Progression Study
             enrolled and followed a cohort of 327 incident AD cases for
             a maximum of 9 years. Drug exposure was expressed using a
             persistency index (PI), calculated as total years of drug
             use divided by total years of observation. Linear
             mixed-effects models examined PI, and interactions with sex
             and apolipoprotein E (APOE) as predictors of clinical
             progression on the Mini-Mental State Examination and
             Clinical Dementia Rating-Sum of Boxes. RESULTS: A total of
             69 participants (21.1%) reported having ever used
             cholinesterase inhibitors or memantine. There was a strong
             three-way interaction between PI, sex, and time. Among
             women, a higher PI (i.e., greater duration of use) of
             cholinesterase inhibitors was associated with slower
             progression on the Mini-Mental State Examination and
             Clinical Dementia Rating-Sum of Boxes, particularly among
             those with an APOE ɛ4 allele. In contrast, higher PI was
             associated with faster progression in males. CONCLUSION: A
             low percentage of individuals with AD in the community are
             taking cholinesterase inhibitors or memantine. This study
             suggests that women, particularly those with an APOE ɛ4
             allele, may benefit the most from these medications. With
             the newly approved increased dose of donepezil, it will be
             imperative to determine whether a higher dose is needed in
             men or whether other factors warrant consideration.},
   Doi = {10.1016/j.jalz.2011.02.011},
   Key = {fds276970}
}

@article{fds277049,
   Author = {Peters, ME and Rosenberg, PB and Steinberg, M and Tschanz, JT and Norton, MC and Welsh-Bohmer, KA and Hayden, KM and Breitner, JCS and Lyketsos, CG and Cache County Investigators},
   Title = {Prevalence of neuropsychiatric symptoms in CIND and its
             subtypes: the Cache County Study.},
   Journal = {Am J Geriatr Psychiatry},
   Volume = {20},
   Number = {5},
   Pages = {416-424},
   Year = {2012},
   Month = {May},
   url = {http://www.ncbi.nlm.nih.gov/pubmed/22522960},
   Abstract = {OBJECTIVES: 1) To report rates of neuropsychiatric symptoms
             (NPS) in cognitive impairment, no dementia (CIND). 2) To
             compare the 30-day prevalence of NPS in CIND with that in
             dementia and cognitively normal individuals. 3) To compare
             the prevalence of NPS in amnestic MCI (aMCI) with other
             predementia syndromes. DESIGN: Comparison of prevalence
             proportions among several defined groups. SETTING:
             Population-based study. PARTICIPANTS: A subsample of the
             permanent residents of Cache County, Utah, aged 65 years or
             older in January 1995 (N = 5092) and who had completed
             clinical assessments and had an informant-completed
             Neuropsychiatric Inventory. MEASUREMENTS: Chi-square
             statistics, tests for trend, and logistic regression models
             were used to analyze the three objectives listed earlier.
             RESULTS: The most prevalent NPS in those with CIND were
             depression (16.9%), irritability (9.8%), nighttime behaviors
             (7.6%), apathy (6.9%), and anxiety (5.4%). Trend analyses
             confirmed that the CIND group had NPS prevalence rates that
             fell between the normal and dementia groups for most NPS.
             Logistic regression models showed no significant difference
             between aMCI and other CIND participants in the prevalence
             of any NPS (lowest p: 0.316). CONCLUSIONS: These data
             confirm the relatively high prevalence of NPS in CIND
             reported by other studies, especially for affective
             symptoms. No differences in NPS prevalence were found
             between aMCI and other types of CIND.},
   Doi = {10.1097/JGP.0b013e318211057d},
   Key = {fds277049}
}

@article{fds371169,
   Author = {Babyak, MA and Schwartz, S and Jiang, R and Brummett, B and Kauwe, JS and Corcoran, C and Munger, R and Welsh-Bohmer, K and Williams, RB and Norton, M},
   Title = {PSYCHOSOCIAL STRESS IN MIDLIFE MAY MODERATE EFFECTS OF
             APOE-TOMM40 RS157580 ON METABOLIC TRAITS IN LATE
             LIFE},
   Journal = {ANNALS OF BEHAVIORAL MEDICINE},
   Volume = {43},
   Pages = {S180-S180},
   Publisher = {SPRINGER},
   Year = {2012},
   Month = {April},
   Key = {fds371169}
}

@article{fds277095,
   Author = {Potter, GG and Wagner, HR and Burke, JR and Plassman, BL and Welsh-Bohmer, KA and Steffens, DC},
   Title = {Neuropsychological Predictors of Dementia in Late-Life Major
             Depressive Disorder.},
   Journal = {Am J Geriatr Psychiatry},
   Year = {2012},
   Month = {March},
   url = {http://www.ncbi.nlm.nih.gov/pubmed/22415312},
   Abstract = {OBJECTIVE:: Major depressive disorder is a likely risk
             factor for dementia, but some cases of major depressive
             disorder in older adults may actually represent a prodrome
             of this condition. The purpose of this study was to use
             neuropsychological test scores to predict conversion to
             dementia in a sample of depressed older adults diagnosed as
             nondemented at the time of neuropsychological testing.
             DESIGN:: Longitudinal, with mean follow-up of 5.45 years.
             SETTING:: Outpatient depression treatment study at Duke
             University. PARTICIPANTS:: Thirty nondemented individuals
             depressed at the time of neuropsychological testing and
             later diagnosed with incident dementia; 149 nondemented
             individuals depressed at the time of neuropsychological
             testing and a diagnosis of cognitively normal. METHODOLOGY::
             All participants received clinical assessment of depression,
             were assessed to rule out prevalent dementia at the time of
             study enrollment, completed neuropsychological testing at
             the time of study enrollment, and were diagnosed for
             cognitive disorders on an annual basis. RESULTS::
             Nondemented, acutely depressed older adults who converted to
             dementia during the study period exhibited broadly lower
             cognitive performances at baseline than acutely depressed
             individuals who remained cognitively normal. Discriminant
             function analysis indicated that 2 neuropsychological tests,
             Recognition Memory (from the Consortium to Establish a
             Registry for Alzheimer's Disease neuropsychological battery)
             and Trail Making B, best predicted dementia conversion.
             CONCLUSIONS:: Depressed older adults with cognitive deficits
             in the domains of memory and executive functions during
             acute depression are at higher risk for developing dementia.
             Some cases of late-life depression may reflect a prodrome of
             dementia in which clinical manifestation of mood changes may
             co-occur with emerging cognitive deficits.},
   Doi = {10.1097/JGP.0b013e318248764e},
   Key = {fds277095}
}

@article{fds276969,
   Author = {Norton, MC and Dew, J and Smith, H and Fauth, E and Piercy, KW and Breitner, JCS and Tschanz, J and Wengreen, H and Welsh-Bohmer, K and Cache County Investigators},
   Title = {Lifestyle behavior pattern is associated with different
             levels of risk for incident dementia and Alzheimer's
             disease: the Cache County study.},
   Journal = {J Am Geriatr Soc},
   Volume = {60},
   Number = {3},
   Pages = {405-412},
   Year = {2012},
   Month = {March},
   ISSN = {0002-8614},
   url = {http://dx.doi.org/10.1111/j.1532-5415.2011.03860.x},
   Abstract = {OBJECTIVES: To identify distinct behavioral patterns of
             diet, exercise, social interaction, church attendance,
             alcohol consumption, and smoking and to examine their
             association with subsequent dementia risk. DESIGN:
             Longitudinal, population-based dementia study. SETTING:
             Rural county in northern Utah, at-home evaluations.
             PARTICIPANTS: Two thousand four hundred ninety-one
             participants without dementia (51% male, average age 73.0 ±
             5,7; average education 13.7 ± 4.1 years) initially reported
             no problems in activities of daily living and no stroke or
             head injury within the past 5 years. MEASUREMENTS: Six
             dichotomized lifestyle behaviors were examined (diet: high
             ≥ median on the Dietary Approaches to Stop Hypertension
             scale; exercise: ≥5 h/wk of light activity and at least
             occasional moderate to vigorous activity; church attendance:
             attending church services at least weekly; social
             Interaction: spending time with family and friends at least
             twice weekly; alcohol: currently drinking alcoholic
             beverages ≥ 2 times/wk; nonsmoker: no current use or fewer
             than 100 cigarettes ever). Latent class analysis (LCA) was
             used to identify patterns among these behaviors.
             Proportional hazards regression modeled time to dementia
             onset as a function of behavioral class, age, sex,
             education, and apolipoprotein E status. Follow-up averaged
             6.3 ± 5.3 years, during which 278 cases of incident
             dementia (200 Alzheimer's disease (AD)) were diagnosed.
             RESULTS: LCA identified four distinct lifestyle classes.
             Unhealthy-religious (UH-R; 11.5%), unhealthy-nonreligious
             (UH-NR; 10.5%), healthy-moderately religious (H-MR; 38.5%),
             and healthy-very religious (H-VR; 39.5%). UH-NR (hazard
             ratio (HR) = 0.54, P = .028), H-MR (HR = 0.56, P = .003),
             and H-VR (HR = 0.58, P = .005) had significantly lower
             dementia risk than UH-R. Results were comparable for AD,
             except that UH-NR was less definitive. CONCLUSION:
             Functionally independent older adults appear to cluster into
             subpopulations with distinct patterns of lifestyle behaviors
             with different levels of risk for subsequent dementia and
             AD.},
   Doi = {10.1111/j.1532-5415.2011.03860.x},
   Key = {fds276969}
}

@article{fds277106,
   Author = {Wee, C-Y and Yap, P-T and Zhang, D and Denny, K and Browndyke, JN and Potter, GG and Welsh-Bohmer, KA and Wang, L and Shen,
             D},
   Title = {Identification of MCI individuals using structural and
             functional connectivity networks.},
   Journal = {Neuroimage},
   Volume = {59},
   Number = {3},
   Pages = {2045-2056},
   Year = {2012},
   Month = {February},
   url = {http://www.ncbi.nlm.nih.gov/pubmed/22019883},
   Abstract = {Different imaging modalities provide essential complementary
             information that can be used to enhance our understanding of
             brain disorders. This study focuses on integrating multiple
             imaging modalities to identify individuals at risk for mild
             cognitive impairment (MCI). MCI, often an early stage of
             Alzheimer's disease (AD), is difficult to diagnose due to
             its very mild or insignificant symptoms of cognitive
             impairment. Recent emergence of brain network analysis has
             made characterization of neurological disorders at a
             whole-brain connectivity level possible, thus providing new
             avenues for brain diseases classification. Employing
             multiple-kernel Support Vector Machines (SVMs), we attempt
             to integrate information from diffusion tensor imaging (DTI)
             and resting-state functional magnetic resonance imaging
             (rs-fMRI) for improving classification performance. Our
             results indicate that the multimodality classification
             approach yields statistically significant improvement in
             accuracy over using each modality independently. The
             classification accuracy obtained by the proposed method is
             96.3%, which is an increase of at least 7.4% from the single
             modality-based methods and the direct data fusion method. A
             cross-validation estimation of the generalization
             performance gives an area of 0.953 under the receiver
             operating characteristic (ROC) curve, indicating excellent
             diagnostic power. The multimodality classification approach
             hence allows more accurate early detection of brain
             abnormalities with greater sensitivity.},
   Doi = {10.1016/j.neuroimage.2011.10.015},
   Key = {fds277106}
}

@article{fds276930,
   Author = {Hayden, KM and Welsh-Bohmer, KA},
   Title = {Epidemiology of cognitive aging and Alzheimer's disease:
             contributions of the cache county utah study of memory,
             health and aging.},
   Journal = {Curr Top Behav Neurosci},
   Volume = {10},
   Pages = {3-31},
   Year = {2012},
   ISSN = {1866-3370},
   url = {http://www.ncbi.nlm.nih.gov/pubmed/21809193},
   Abstract = {Epidemiological studies of Alzheimer's disease (AD) provide
             insights into changing public health trends and their
             contribution to disease incidence. The current chapter
             considers how the population-based approach has contributed
             to our understanding of lifetime exposures that contribute
             to later disease risk and may act to modify onset of
             symptoms. We focus on the findings from a recent survey of
             an exceptionally long-lived population, the Cache County
             Utah Study of Memory, Health, and Aging. This study is
             confined to a single geographic population has allowed
             estimation of the genetic and environmental influences on AD
             expression across the expected human lifespan of 95+ years.
             Given the emphasis of this text on the behavioral
             neurosciences of aging, we highlight within the current
             chapter the particular contributions of this
             population-based study to the neuropsychology of aging and
             AD. We also discuss hypotheses generated from this survey
             with respect to factors that may either accelerate or delay
             symptom onset in AD and the conditions that appear to be
             associated with successful cognitive aging.},
   Doi = {10.1007/7854_2011_152},
   Key = {fds276930}
}

@article{fds276968,
   Author = {Tranel, D and Welsh-Bohmer, KA},
   Title = {Pervasive olfactory impairment after bilateral limbic system
             destruction.},
   Journal = {J Clin Exp Neuropsychol},
   Volume = {34},
   Number = {2},
   Pages = {117-125},
   Year = {2012},
   url = {http://www.ncbi.nlm.nih.gov/pubmed/22220560},
   Abstract = {What pattern of brain damage could completely obliterate the
             sense of olfaction in humans? We had an opportunity to
             address this intriguing question in Patient B., who has
             extensive bilateral damage to most of the limbic system,
             including the medial and lateral temporal lobes, orbital
             frontal cortex, insular cortex, anterior cingulate cortex,
             and basal forebrain, caused by herpes simplex encephalitis.
             The patient demonstrated profound impairments in odor
             identification and recognition. Moreover, he could not
             discriminate between olfactory stimuli, and he had severe
             impairments in odor detection. Reliable stimulus detection
             was obtained only for solutions of the organic solvent
             acetone and highly concentrated solutions of ethanol. In
             contrast to the more circumscribed olfactory deficits
             demonstrated in patients with damage confined to either the
             temporal lobes or orbitofrontal cortex (which tend to
             involve odor identification but not odor detection), Patient
             B. demonstrated a strikingly severe and complete anosmia.
             This contrast in olfactory abilities and deficits as a
             result of different anatomical pathology affords new
             insights into the neural substrates of olfactory processing
             in humans.},
   Doi = {10.1080/13803395.2011.633897},
   Key = {fds276968}
}

@article{fds277107,
   Author = {Wee, C-Y and Yap, P-T and Denny, K and Browndyke, JN and Potter, GG and Welsh-Bohmer, KA and Wang, L and Shen, D},
   Title = {Resting-state multi-spectrum functional connectivity
             networks for identification of MCI patients.},
   Journal = {PLoS One},
   Volume = {7},
   Number = {5},
   Pages = {e37828},
   Year = {2012},
   url = {http://www.ncbi.nlm.nih.gov/pubmed/22666397},
   Abstract = {In this paper, a high-dimensional pattern classification
             framework, based on functional associations between brain
             regions during resting-state, is proposed to accurately
             identify MCI individuals from subjects who experience normal
             aging. The proposed technique employs multi-spectrum
             networks to characterize the complex yet subtle blood
             oxygenation level dependent (BOLD) signal changes caused by
             pathological attacks. The utilization of multi-spectrum
             networks in identifying MCI individuals is motivated by the
             inherent frequency-specific properties of BOLD spectrum. It
             is believed that frequency specific information extracted
             from different spectra may delineate the complex yet subtle
             variations of BOLD signals more effectively. In the proposed
             technique, regional mean time series of each
             region-of-interest (ROI) is band-pass filtered (0.025 ≤ ƒ
             ≤ 0.100 Hz) before it is decomposed into five frequency
             sub-bands. Five connectivity networks are constructed, one
             from each frequency sub-band. Clustering coefficient of each
             ROI in relation to the other ROIs are extracted as features
             for classification. Classification accuracy was evaluated
             via leave-one-out cross-validation to ensure generalization
             of performance. The classification accuracy obtained by this
             approach is 86.5%, which is an increase of at least 18.9%
             from the conventional full-spectrum methods. A
             cross-validation estimation of the generalization
             performance shows an area of 0.863 under the receiver
             operating characteristic (ROC) curve, indicating good
             diagnostic power. It was also found that, based on the
             selected features, portions of the prefrontal cortex,
             orbitofrontal cortex, temporal lobe, and parietal lobe
             regions provided the most discriminant information for
             classification, in line with results reported in previous
             studies. Analysis on individual frequency sub-bands
             demonstrated that different sub-bands contribute differently
             to classification, providing extra evidence regarding
             frequency-specific distribution of BOLD signals. Our MCI
             classification framework, which allows accurate early
             detection of functional brain abnormalities, makes an
             important positive contribution to the treatment management
             of potential AD patients.},
   Doi = {10.1371/journal.pone.0037828},
   Key = {fds277107}
}

@article{fds277123,
   Author = {Linnertz, C and Saunders, AM and Lutz, MW and Crenshaw, DM and Grossman,
             I and Burns, DK and Whitfield, KE and Hauser, MA and McCarthy, JJ and Ulmer, M and Allingham, R and Welsh-Bohmer, KA and Roses, AD and Chiba-Falek, O},
   Title = {Characterization of the poly-T variant in the TOMM40 gene in
             diverse populations.},
   Journal = {PLoS One},
   Volume = {7},
   Number = {2},
   Pages = {e30994},
   Year = {2012},
   url = {http://www.ncbi.nlm.nih.gov/pubmed/22359560},
   Abstract = {We previously discovered that a polymorphic,
             deoxythymidine-homopolymer (poly-T, rs10524523) in intron 6
             of the TOMM40 gene is associated with age-of-onset of
             Alzheimer's disease and with cognitive performance in
             elderly. Three allele groups were defined for rs10524523,
             hereafter '523', based on the number of 'T'-residues:
             'Short' (S, T≤19), 'Long' (L, 20≤T≤29) and 'Very Long'
             (VL, T≥30). Homopolymers, particularly long homopolymers
             like '523', are difficult to genotype because 'slippage'
             occurs during PCR-amplification. We initially genotyped this
             locus by PCR-amplification followed by Sanger-sequencing.
             However, we recognized the need to develop a
             higher-throughput genotyping method that is also accurate
             and reliable. Here we describe a new '523' genotyping assay
             that is simple and inexpensive to perform in a standard
             molecular genetics laboratory. The assay is based on the
             detection of differences in PCR-fragment length using
             capillary electrophoresis. We discuss technical problems,
             solutions, and the steps taken for validation. We employed
             the novel assay to investigate the '523' allele frequencies
             in different ethnicities. Whites and Hispanics have similar
             frequencies of S/L/VL alleles (0.45/0.11/0.44 and
             0.43/0.09/0.48, respectively). In African-Americans, the
             frequency of the L-allele (0.10) is similar to Whites and
             Hispanics; however, the S-allele is more prevalent (0.65)
             and the VL-allele is concomitantly less frequent (0.25). The
             allele frequencies determined using the new methodology are
             compared to previous reports for Ghanaian, Japanese, Korean
             and Han Chinese cohorts. Finally, we studied the linkage
             pattern between TOMM40-'523' and APOE alleles. In Whites and
             Hispanics, consistent with previous reports, the L is
             primarily linked to ε4, while the majority of the VL and S
             are linked to ε3. Interestingly, in African-Americans,
             Ghanaians and Japanese, there is an increased frequency of
             the '523'S-APOEε4 haplotype. These data may be used as
             references for '523' allele and '523'-APOE haplotype
             frequencies in diverse populations for the design of
             research studies and clinical trials.},
   Doi = {10.1371/journal.pone.0030994},
   Key = {fds277123}
}

@article{fds276967,
   Author = {Kramer, PL and Xu, H and Woltjer, RL and Westaway, SK and Clark, D and Erten-Lyons, D and Kaye, JA and Welsh-Bohmer, KA and Troncoso, JC and Markesbery, WR and Petersen, RC and Turner, RS and Kukull, WA and Bennett, DA and Galasko, D and Morris, JC and Ott,
             J},
   Title = {Alzheimer disease pathology in cognitively healthy elderly:
             a genome-wide study.},
   Journal = {Neurobiol Aging},
   Volume = {32},
   Number = {12},
   Pages = {2113-2122},
   Year = {2011},
   Month = {December},
   url = {http://www.ncbi.nlm.nih.gov/pubmed/20452100},
   Abstract = {Many elderly individuals remain dementia-free throughout
             their life. However, some of these individuals exhibit
             Alzheimer disease neuropathology on autopsy, evidenced by
             neurofibrillary tangles (NFTs) in AD-specific brain regions.
             We conducted a genome-wide association study to identify
             genetic mechanisms that distinguish non-demented elderly
             with a heavy NFT burden from those with a low NFT burden.
             The study included 299 non-demented subjects with autopsy
             (185 subjects with low and 114 with high NFT levels). Both a
             genotype test, using logistic regression, and an allele test
             provided consistent evidence that variants in the RELN gene
             are associated with neuropathology in the context of
             cognitive health. Immunohistochemical data for reelin
             expression in AD-related brain regions added support for
             these findings. Reelin signaling pathways modulate
             phosphorylation of tau, the major component of NFTs, either
             directly or through β-amyloid pathways that influence tau
             phosphorylation. Our findings suggest that up-regulation of
             reelin may be a compensatory response to tau-related or
             beta-amyloid stress associated with AD even prior to the
             onset of dementia.},
   Doi = {10.1016/j.neurobiolaging.2010.01.010},
   Key = {fds276967}
}

@article{fds277097,
   Author = {Fillenbaum, GG and Burchett, BM and Unverzagt, FW and Rexroth, DF and Welsh-Bohmer, K},
   Title = {Norms for CERAD constructional praxis recall.},
   Journal = {Clin Neuropsychol},
   Volume = {25},
   Number = {8},
   Pages = {1345-1358},
   Year = {2011},
   Month = {November},
   url = {http://www.ncbi.nlm.nih.gov/pubmed/21992077},
   Abstract = {Recall of the four-item constructional praxis measure was a
             later addition to the Consortium to Establish a Registry for
             Alzheimer's Disease (CERAD) neuropsychological battery.
             Norms for this measure, based on cognitively intact African
             Americans age ≥70 (Indianapolis-Ibadan Dementia Project,
             N=372), European American participants age ≥66 (Cache
             County Study of Memory, Health and Aging, N=507), and
             European American CERAD clinic controls age ≥50
             (N = 182), are presented here. Performance varied by
             site; by sex, education, and age (African Americans in
             Indianapolis); education and age (Cache County European
             Americans); and only age (CERAD European American controls).
             Performance declined with increased age, within age with
             less education, and was poorer for women. Means, standard
             deviations, and percentiles are presented separately for
             each sample.},
   Doi = {10.1080/13854046.2011.614962},
   Key = {fds277097}
}

@article{fds276982,
   Author = {Norton, MC and Smith, KR and Østbye, T and Tschanz, JT and Schwartz, S and Corcoran, C and Breitner, JCS and Steffens, DC and Skoog, I and Rabins,
             PV and Welsh-Bohmer, KA and Cache County Investigators},
   Title = {Early parental death and remarriage of widowed parents as
             risk factors for Alzheimer disease: the Cache County
             study.},
   Journal = {Am J Geriatr Psychiatry},
   Volume = {19},
   Number = {9},
   Pages = {814-824},
   Year = {2011},
   Month = {September},
   url = {http://www.ncbi.nlm.nih.gov/pubmed/21873837},
   Abstract = {OBJECTIVES: Early parental death is associated with lifelong
             tendencies toward depression and chronic stress. We tested
             the hypothesis that early parental death is associated with
             higher risk for Alzheimer disease (AD) in offspring. DESIGN:
             A population-based epidemiological study of dementia with
             detailed clinical evaluations, linked to one of the world's
             richest sources of objective genealogical and vital
             statistics data. SETTING: Home visits with residents of a
             rural county in northern Utah. PARTICIPANTS: 4,108 subjects,
             aged 65-105. MEASUREMENTS: Multistage dementia ascertainment
             protocol implemented in four triennial waves, yielding
             expert consensus diagnoses of 570 participants with AD and
             3,538 without dementia. Parental death dates, socioeconomic
             status, and parental remarriage after widowhood were
             obtained from the Utah Population Database, a large
             genealogical database linked to statewide birth and death
             records. RESULTS: Mother's death during subject's
             adolescence was significantly associated with higher rate of
             AD in regression models that included age, gender,
             education, APOE genotype, and socioeconomic status. Father's
             death before subject age 5 showed a weaker association. In
             stratified analyses, associations were significant only when
             the widowed parent did not remarry. Parental death
             associations were not moderated by gender or APOE genotype.
             Findings were specific to AD and not found for non-AD
             dementia. CONCLUSIONS: Parental death during childhood is
             associated with higher prevalence of AD, with different
             critical periods for father's versus mother's death, with
             strength of these associations attenuated by remarriage of
             the widowed parent.},
   Doi = {10.1097/JGP.0b013e3182011b38},
   Key = {fds276982}
}

@article{fds277094,
   Author = {Plassman, BL and Langa, KM and McCammon, RJ and Fisher, GG and Potter,
             GG and Burke, JR and Steffens, DC and Foster, NL and Giordani, B and Unverzagt, FW and Welsh-Bohmer, KA and Heeringa, SG and Weir, DR and Wallace, RB},
   Title = {Incidence of dementia and cognitive impairment, not dementia
             in the United States.},
   Journal = {Ann Neurol},
   Volume = {70},
   Number = {3},
   Pages = {418-426},
   Year = {2011},
   Month = {September},
   url = {http://www.ncbi.nlm.nih.gov/pubmed/21425187},
   Abstract = {OBJECTIVE: Estimates of incident dementia, and cognitive
             impairment, not dementia (CIND) (or the related mild
             cognitive impairment) are important for public health and
             clinical care policy. In this paper, we report US national
             incidence rates for dementia and CIND. METHODS: Participants
             in the Aging, Demographic, and Memory Study (ADAMS) were
             evaluated for cognitive impairment using a comprehensive
             in-home assessment. A total of 456 individuals aged 72 years
             and older, who were not demented at baseline, were followed
             longitudinally from August 2001 to December 2009. An expert
             consensus panel assigned a diagnosis of normal cognition,
             CIND, or dementia and its subtypes. Using a
             population-weighted sample, we estimated the incidence of
             dementia, Alzheimer disease (AD), vascular dementia (VaD),
             and CIND by age. We also estimated the incidence of
             progression from CIND to dementia. RESULTS: The incidence of
             dementia was 33.3 (standard error [SE], 4.2) per 1,000
             person-years and 22.9 (SE, 2.9) per 1,000 person-years for
             AD. The incidence of CIND was 60.4 (SE, 7.2) cases per 1,000
             person-years. An estimated 120.3 (SE, 16.9) individuals per
             1,000 person-years progressed from CIND to dementia. Over a
             5.9-year period, about 3.4 million individuals aged 72 and
             older in the United States developed incident dementia, of
             whom approximately 2.3 million developed AD, and about
             637,000 developed VaD. Over this same period, almost 4.8
             million individuals developed incident CIND. INTERPRETATION:
             The incidence of CIND is greater than the incidence of
             dementia, and those with CIND are at high risk of
             progressing to dementia, making CIND a potentially valuable
             target for treatments aimed at slowing cognitive
             decline.},
   Doi = {10.1002/ana.22362},
   Key = {fds277094}
}

@article{fds276965,
   Author = {Treiber, KA and Carlson, MC and Corcoran, C and Norton, MC and Breitner,
             JCS and Piercy, KW and Deberard, MS and Stein, D and Foley, B and Welsh-Bohmer, KA and Frye, A and Lyketsos, CG and Tschanz,
             JT},
   Title = {Cognitive stimulation and cognitive and functional decline
             in Alzheimer's disease: the cache county dementia
             progression study.},
   Journal = {J Gerontol B Psychol Sci Soc Sci},
   Volume = {66},
   Number = {4},
   Pages = {416-425},
   Year = {2011},
   Month = {July},
   url = {http://www.ncbi.nlm.nih.gov/pubmed/21441386},
   Abstract = {OBJECTIVES: To examine the association of engagement in
             cognitively stimulating activities with cognitive and
             functional decline in a population-based sample of incident
             Alzheimer's disease (AD). METHOD: After diagnosis, 187
             participants (65% females) were followed semiannually for a
             mean 2.7 (SD = 0.4) years. Mean age and education were 84.6
             (SD = 5.8) and 13.2 (SD = 2.9) years. Caregivers enumerated
             cognitively stimulating leisure activities via the Lifestyle
             Activities Questionnaire. Cognition was assessed using the
             Mini-Mental State Examination and functional ability via the
             Clinical Dementia Rating sum of boxes. Linear mixed models
             tested the association between stimulating activities and
             change over time in each outcome. Covariates were
             demographic factors, estimated premorbid IQ,
             presence/absence of the APOE ε4 allele, duration of
             dementia, level of physical activity, and general health.
             RESULTS: At initial assessment, 87% of participants were
             engaged in one or more stimulating activities, with mean
             (SD) activities = 4.0 (3.0). This number declined to 2.4
             (2.0) at the final visit. There was a statistical
             interaction between dementia duration and number of
             activities in predicting rate of cognitive decline (p = .02)
             and overall functional ability (p = .006). DISCUSSION:
             Active involvement in cognitively stimulating pursuits may
             be beneficial for persons with AD.},
   Doi = {10.1093/geronb/gbr023},
   Key = {fds276965}
}

@article{fds276966,
   Author = {Tschanz, JT and Corcoran, CD and Schwartz, S and Treiber, K and Green,
             RC and Norton, MC and Mielke, MM and Piercy, K and Steinberg, M and Rabins,
             PV and Leoutsakos, J-M and Welsh-Bohmer, KA and Breitner, JCS and Lyketsos, CG},
   Title = {Progression of cognitive, functional, and neuropsychiatric
             symptom domains in a population cohort with Alzheimer
             dementia: the Cache County Dementia Progression
             study.},
   Journal = {Am J Geriatr Psychiatry},
   Volume = {19},
   Number = {6},
   Pages = {532-542},
   Year = {2011},
   Month = {June},
   url = {http://www.ncbi.nlm.nih.gov/pubmed/21606896},
   Abstract = {OBJECTIVES: Progression of Alzheimer dementia (AD) is highly
             variable. Most estimates derive from convenience samples
             from dementia clinics or research centers where there is
             substantial potential for survival bias and other
             distortions. In a population-based sample of incident AD
             cases, we examined progression of impairment in cognition,
             function, and neuropsychiatric symptoms, and the influence
             of selected variables on these domains. DESIGN:
             Longitudinal, prospective cohort study. SETTING: Cache
             County (Utah). PARTICIPANTS: Three hundred twenty-eight
             persons with a diagnosis of possible/probable AD.
             MEASUREMENTS: Mini-Mental State Exam (MMSE), Clinical
             Dementia Rating sum-of-boxes (CDR-sb), and Neuropsychiatric
             Inventory (NPI). RESULTS: Over a mean follow-up of 3.80
             (range: 0.07-12.90) years, the mean (SD) annual rates of
             change were -1.53 (2.69) scale points on the MMSE, 1.44
             (1.82) on the CDR-sb, and 2.55 (5.37) on the NPI. Among
             surviving participants, 30% to 58% progressed less than 1
             point per year on these measures, even 5 to 7 years after
             dementia onset. Rates of change were correlated between MMSE
             and CDR-sb (r = -0.62, df = 201, p < 0.001) and between the
             CDR-sb and NPI (r = 0.20, df = 206, p < 0.004). Female
             subjects (LR χ = 8.7, df = 2, p = 0.013) and those with
             younger onset (likelihood ratio [LR] χ = 5.7, df = 2, p =
             0.058) declined faster on the MMSE. Although one or more
             apolipoprotein E ε 4 alleles and ever use of FDA-approved
             antidementia medications were associated with initial MMSE
             scores, neither was related to the rate of progression in
             any domain. CONCLUSIONS: A significant proportion of persons
             with AD progresses slowly. The results underscore
             differences between population-based versus clinic-based
             samples and suggest ongoing need to identify factors that
             may slow the progression of AD.},
   Doi = {10.1097/JGP.0b013e3181faec23},
   Key = {fds276966}
}

@article{fds277120,
   Author = {Swaminathan, M and Nicoara, A and Phillips-Bute, BG and Aeschlimann,
             N and Milano, CA and Mackensen, GB and Podgoreanu, MV and Velazquez, EJ and Stafford-Smith, M and Mathew, JP and Cardiothoracic Anesthesia
             Research Endeavors (CARE) Group},
   Title = {Utility of a simple algorithm to grade diastolic dysfunction
             and predict outcome after coronary artery bypass graft
             surgery.},
   Journal = {Ann Thorac Surg},
   Volume = {91},
   Number = {6},
   Pages = {1844-1850},
   Year = {2011},
   Month = {June},
   url = {http://www.ncbi.nlm.nih.gov/pubmed/21492828},
   Abstract = {BACKGROUND: Inclusion of a measure of left ventricular
             diastolic dysfunction (LVDD) may improve risk prediction
             after cardiac surgery. Current LVDD grading guidelines rely
             on echocardiographic variables that are not always available
             or aligned to allow grading. We hypothesized that a
             simplified algorithm involving fewer variables would enable
             more patients to be assigned a LVDD grade compared with a
             comprehensive algorithm, and also be valid in identifying
             patients at risk of long-term major adverse cardiac events
             (MACE). METHODS: Intraoperative transesophageal
             echocardiography data were gathered on 905 patients
             undergoing coronary artery bypass graft surgery, including
             flow and tissue Doppler-based measurements. Two algorithms
             were constructed to categorize LVDD: a comprehensive
             four-variable algorithm, A, was compared with a simplified
             version, B, with only two variables-transmitral early flow
             velocity and early mitral annular tissue velocity-for ease
             of grading and association with MACE. RESULTS: Using
             algorithm A, only 563 patients (62%) could be graded,
             whereas 895 patients (99%) received a grade with algorithm
             B. Over the median follow-up period of 1,468 days, Cox
             modeling showed that LVDD was significantly associated with
             MACE when graded with algorithm B (p=0.013), but not
             algorithm A (p=0.79). Patients with the highest incidence of
             MACE could not be graded with algorithm A. CONCLUSIONS: We
             found that an LVDD algorithm with fewer variables enabled
             grading of a significantly greater number of coronary artery
             bypass graft patients, and was valid, as evidenced by
             worsening grades being associated with MACE. This simplified
             algorithm could be extended to similar populations as a
             valid method of characterizing LVDD.},
   Doi = {10.1016/j.athoracsur.2011.02.008},
   Key = {fds277120}
}

@article{fds277092,
   Author = {Kaddurah-Daouk, R and Rozen, S and Matson, W and Han, X and Hulette, CM and Burke, JR and Doraiswamy, PM and Welsh-Bohmer,
             KA},
   Title = {Metabolomic changes in autopsy-confirmed Alzheimer's
             disease.},
   Journal = {Alzheimers Dement},
   Volume = {7},
   Number = {3},
   Pages = {309-317},
   Year = {2011},
   Month = {May},
   url = {http://www.ncbi.nlm.nih.gov/pubmed/21075060},
   Abstract = {BACKGROUND: Metabolomics, the global science of
             biochemistry, provides powerful tools to map perturbations
             in the metabolic network and enables simultaneous
             quantification of several metabolites to identify metabolic
             perturbances that might provide insights into disease.
             METHODS: In this pilot study, we took a targeted
             electrochemistry-based metabolomics approach where liquid
             chromatography followed by coulometric array detection
             enables quantification of over 30 metabolites within key
             neurotransmitter pathways (dopamine and serotonin) and
             pathways involved in oxidative stress. RESULTS: Using
             samples from postmortem ventricular cerebrospinal fluid (15
             Alzheimer's disease [AD] and 15 nondemented subjects with
             autopsy-confirmed diagnoses) and by using regression models,
             correlations, Wilcoxon rank-sum tests, and t-tests we
             identified alterations in tyrosine, tryptophan, purine, and
             tocopherol pathways in patients with AD. Reductions in
             norepinephrine and its related metabolites were also seen,
             consistent with previously published data. CONCLUSIONS:
             These data support further investigation of metabolomics in
             larger samples of clinical AD as well as in those with
             preclinical disease for use as biomarkers.},
   Doi = {10.1016/j.jalz.2010.06.001},
   Key = {fds277092}
}

@article{fds277093,
   Author = {Naj, AC and Jun, G and Beecham, GW and Wang, L-S and Vardarajan, BN and Buros, J and Gallins, PJ and Buxbaum, JD and Jarvik, GP and Crane, PK and Larson, EB and Bird, TD and Boeve, BF and Graff-Radford, NR and De
             Jager, PL and Evans, D and Schneider, JA and Carrasquillo, MM and Ertekin-Taner, N and Younkin, SG and Cruchaga, C and Kauwe, JSK and Nowotny, P and Kramer, P and Hardy, J and Huentelman, MJ and Myers, AJ and Barmada, MM and Demirci, FY and Baldwin, CT and Green, RC and Rogaeva,
             E and St George-Hyslop and P and Arnold, SE and Barber, R and Beach, T and Bigio, EH and Bowen, JD and Boxer, A and Burke, JR and Cairns, NJ and Carlson, CS and Carney, RM and Carroll, SL and Chui, HC and Clark, DG and Corneveaux, J and Cotman, CW and Cummings, JL and DeCarli, C and DeKosky, ST and Diaz-Arrastia, R and Dick, M and Dickson, DW and Ellis,
             WG and Faber, KM and Fallon, KB and Farlow, MR and Ferris, S and Frosch,
             MP and Galasko, DR and Ganguli, M and Gearing, M and Geschwind, DH and Ghetti, B and Gilbert, JR and Gilman, S and Giordani, B and Glass, JD and Growdon, JH and Hamilton, RL and Harrell, LE and Head, E and Honig, LS and Hulette, CM and Hyman, BT and Jicha, GA and Jin, L-W and Johnson, N and Karlawish, J and Karydas, A and Kaye, JA and Kim, R and Koo, EH and Kowall,
             NW and Lah, JJ and Levey, AI and Lieberman, AP and Lopez, OL and Mack, WJ and Marson, DC and Martiniuk, F and Mash, DC and Masliah, E and McCormick,
             WC and McCurry, SM and McDavid, AN and McKee, AC and Mesulam, M and Miller,
             BL and Miller, CA and Miller, JW and Parisi, JE and Perl, DP and Peskind,
             E and Petersen, RC and Poon, WW and Quinn, JF and Rajbhandary, RA and Raskind, M and Reisberg, B and Ringman, JM and Roberson, ED and Rosenberg, RN and Sano, M and Schneider, LS and Seeley, W and Shelanski,
             ML and Slifer, MA and Smith, CD and Sonnen, JA and Spina, S and Stern, RA and Tanzi, RE and Trojanowski, JQ and Troncoso, JC and Van Deerlin and VM and Vinters, HV and Vonsattel, JP and Weintraub, S and Welsh-Bohmer, KA and Williamson, J and Woltjer, RL and Cantwell, LB and Dombroski, BA and Beekly, D and Lunetta, KL and Martin, ER and Kamboh, MI and Saykin, AJ and Reiman, EM and Bennett, DA and Morris, JC and Montine, TJ and Goate, AM and Blacker, D and Tsuang, DW and Hakonarson, H and Kukull, WA and Foroud,
             TM and Haines, JL and Mayeux, R and Pericak-Vance, MA and Farrer, LA and Schellenberg, GD},
   Title = {Common variants at MS4A4/MS4A6E, CD2AP, CD33 and EPHA1 are
             associated with late-onset Alzheimer's disease.},
   Journal = {Nat Genet},
   Volume = {43},
   Number = {5},
   Pages = {436-441},
   Year = {2011},
   Month = {May},
   url = {http://www.ncbi.nlm.nih.gov/pubmed/21460841},
   Abstract = {The Alzheimer Disease Genetics Consortium (ADGC) performed a
             genome-wide association study of late-onset Alzheimer
             disease using a three-stage design consisting of a discovery
             stage (stage 1) and two replication stages (stages 2 and 3).
             Both joint analysis and meta-analysis approaches were used.
             We obtained genome-wide significant results at MS4A4A
             (rs4938933; stages 1 and 2, meta-analysis P (P(M)) = 1.7 ×
             10(-9), joint analysis P (P(J)) = 1.7 × 10(-9); stages 1, 2
             and 3, P(M) = 8.2 × 10(-12)), CD2AP (rs9349407; stages 1, 2
             and 3, P(M) = 8.6 × 10(-9)), EPHA1 (rs11767557; stages 1, 2
             and 3, P(M) = 6.0 × 10(-10)) and CD33 (rs3865444; stages 1,
             2 and 3, P(M) = 1.6 × 10(-9)). We also replicated previous
             associations at CR1 (rs6701713; P(M) = 4.6 × 10(-10), P(J)
             = 5.2 × 10(-11)), CLU (rs1532278; P(M) = 8.3 × 10(-8),
             P(J) = 1.9 × 10(-8)), BIN1 (rs7561528; P(M) = 4.0 ×
             10(-14), P(J) = 5.2 × 10(-14)) and PICALM (rs561655; P(M) =
             7.0 × 10(-11), P(J) = 1.0 × 10(-10)), but not at EXOC3L2,
             to late-onset Alzheimer's disease susceptibility.},
   Doi = {10.1038/ng.801},
   Key = {fds277093}
}

@article{fds277026,
   Author = {Tate, DF and Neeley, ES and Norton, MC and Tschanz, JT and Miller, MJ and Wolfson, L and Hulette, C and Leslie, C and Welsh-Bohmer, KA and Plassman, B and Bigler, ED},
   Title = {Intracranial volume and dementia: some evidence in support
             of the cerebral reserve hypothesis.},
   Journal = {Brain Res},
   Volume = {1385},
   Pages = {151-162},
   Year = {2011},
   Month = {April},
   url = {http://www.ncbi.nlm.nih.gov/pubmed/21172323},
   Abstract = {The brain reserve hypothesis has been posited as being one
             important mediating factor for developing dementia,
             especially Alzheimer's disease (AD). Evidence for this
             hypothesis is mixed though different methodologies have made
             these findings difficult to interpret. We examined imaging
             data from a large cohort (N=194) of mixed dementia patients
             and controls, 65years old and older from the Cache County,
             Utah Study of Memory and Aging for evidence of the brain
             reserve hypothesis using total intracranial volume (TICV) as
             a quantitative measure of pre-morbid brain size and a
             vicarious indicator of reserve. A broader spectrum of
             non-demented elderly control subjects from previous studies
             was also included for comparison (N=423). In addition,
             non-parametric Classification and Regression Tree (CART)
             analyses were performed to model group heterogeneity and
             identify any subgroups of patients where TICV might be an
             important predictor of dementia. Parametrically, no main
             effect was found for TICV when predicting a dementia
             diagnosis; however, the CART analysis did reveal important
             TICV subgroups, including a sex differential wherein ε4
             APOE allele presence in males and low TICV predicted AD
             classification. TICV, APOE, and other potential
             mediator/moderator variables are discussed in the context of
             the brain reserve hypothesis.},
   Doi = {10.1016/j.brainres.2010.12.038},
   Key = {fds277026}
}

@article{fds277090,
   Author = {McCarthy, JJ and Linnertz, C and Saucier, L and Burke, JR and Hulette,
             CM and Welsh-Bohmer, KA and Chiba-Falek, O},
   Title = {The effect of SNCA 3' region on the levels of SNCA-112
             splicing variant.},
   Journal = {Neurogenetics},
   Volume = {12},
   Number = {1},
   Pages = {59-64},
   Year = {2011},
   Month = {February},
   url = {http://www.ncbi.nlm.nih.gov/pubmed/21046180},
   Abstract = {Genetic variability at the 3' region of SNCA locus has been
             repeatedly associated with susceptibility to sporadic
             Parkinson's disease (PD). Accumulated evidence emphasizes
             the importance of SNCA dosage and expression levels in PD
             pathogenesis. However, the mechanism through which the 3'
             region of SNCA gene modulates the risk to develop sporadic
             PD remained elusive. We studied the effect of PD
             risk-associated variants at SNCA 3' regions on SNCA112-mRNA
             (exon 5 in-frame skipping) levels in vivo in 117
             neuropathologically normal, human brain frontal cortex
             samples. SNPs tagging the SNCA 3' showed significant effects
             on the relative levels of SNCA112-mRNA from total SNCA
             transcripts levels. The "risk" alleles were correlated with
             increased expression ratio of SNCA112-mRNA from total. We
             provide evidence for functional consequences of
             PD-associated SNCA gene variants at the 3' region,
             suggesting that genetic regulation of SNCA splicing plays an
             important role in the development of the disease. Further
             studies to determine the definite functional variant/s
             within SNCA 3'and to establish their association with PD
             pathology are necessary.},
   Doi = {10.1007/s10048-010-0263-4},
   Key = {fds277090}
}

@article{fds277103,
   Author = {Wee, C-Y and Yap, P-T and Li, W and Denny, K and Browndyke, JN and Potter,
             GG and Welsh-Bohmer, KA and Wang, L and Shen, D},
   Title = {Enriched white matter connectivity networks for accurate
             identification of MCI patients.},
   Journal = {Neuroimage},
   Volume = {54},
   Number = {3},
   Pages = {1812-1822},
   Year = {2011},
   Month = {February},
   url = {http://www.ncbi.nlm.nih.gov/pubmed/20970508},
   Abstract = {Mild cognitive impairment (MCI), often a prodromal phase of
             Alzheimer's disease (AD), is frequently considered to be a
             good target for early diagnosis and therapeutic
             interventions of AD. Recent emergence of reliable network
             characterization techniques has made it possible to
             understand neurological disorders at a whole-brain
             connectivity level. Accordingly, we propose an effective
             network-based multivariate classification algorithm, using a
             collection of measures derived from white matter (WM)
             connectivity networks, to accurately identify MCI patients
             from normal controls. An enriched description of WM
             connections, utilizing six physiological parameters, i.e.,
             fiber count, fractional anisotropy (FA), mean diffusivity
             (MD), and principal diffusivities(λ(1), λ(2), and λ(3)),
             results in six connectivity networks for each subject to
             account for the connection topology and the biophysical
             properties of the connections. Upon parcellating the brain
             into 90 regions-of-interest (ROIs), these properties can be
             quantified for each pair of regions with common traversing
             fibers. For building an MCI classifier, clustering
             coefficient of each ROI in relation to the remaining ROIs is
             extracted as feature for classification. These features are
             then ranked according to their Pearson correlation with
             respect to the clinical labels, and are further sieved to
             select the most discriminant subset of features using an
             SVM-based feature selection algorithm. Finally, support
             vector machines (SVMs) are trained using the selected subset
             of features. Classification accuracy was evaluated via
             leave-one-out cross-validation to ensure generalization of
             performance. The classification accuracy given by our
             enriched description of WM connections is 88.9%, which is an
             increase of at least 14.8% from that using simple WM
             connectivity description with any single physiological
             parameter. A cross-validation estimation of the
             generalization performance shows an area of 0.929 under the
             receiver operating characteristic (ROC) curve, indicating
             excellent diagnostic power. It was also found, based on the
             selected features, that portions of the prefrontal cortex,
             orbitofrontal cortex, parietal lobe and insula regions
             provided the most discriminant features for classification,
             in line with results reported in previous studies. Our MCI
             classification framework, especially the enriched
             description of WM connections, allows accurate early
             detection of brain abnormalities, which is of paramount
             importance for treatment management of potential AD
             patients.},
   Doi = {10.1016/j.neuroimage.2010.10.026},
   Key = {fds277103}
}

@article{fds277048,
   Author = {Mayeux, R and Reitz, C and Brickman, AM and Haan, MN and Manly, JJ and Glymour, MM and Weiss, CC and Yaffe, K and Middleton, L and Hendrie, HC and Warren, LH and Hayden, KM and Welsh-Bohmer, KA and Breitner, JCS and Morris, JC},
   Title = {Operationalizing diagnostic criteria for Alzheimer's disease
             and other age-related cognitive impairment-Part
             1.},
   Journal = {Alzheimers Dement},
   Volume = {7},
   Number = {1},
   Pages = {15-34},
   Year = {2011},
   Month = {January},
   url = {http://www.ncbi.nlm.nih.gov/pubmed/21255741},
   Abstract = {In this article, the challenges faced by several noted
             population studies for Alzheimer dementia in
             operationalizing current clinical diagnostic criteria for
             Alzheimer's disease (AD) have been reviewed. Differences in
             case ascertainment, methodological biases, cultural and
             educational influences on test performance, inclusion of
             special populations such as underrepresented minorities and
             the oldest old, and detection of the earliest symptomatic
             stages of underlying AD have been considered. Classification
             of Alzheimer dementia may be improved by the incorporation
             of biomarkers for AD if the sensitivity, specificity, and
             predictive value of the biomarkers are established and if
             they are appropriate for epidemiological studies, as may
             occur should a plasma biomarker be developed. Biomarkers for
             AD could also facilitate studies of the interactions of
             various forms of neurodegenerative disorders with
             cerebrovascular disease, resulting in "mixed
             dementia".},
   Doi = {10.1016/j.jalz.2010.11.005},
   Key = {fds277048}
}

@article{fds371170,
   Author = {Norton, MC and Franklin, LM and Bingham, D and Steffens, DC and Welsh-Bohmer, K},
   Title = {INDIVIDUAL RELIGIOUS BEHAVIORS ARE INVERSELY ASSOCIATED WITH
             SUBSEQUENT GERIATRIC DEPRESSION},
   Journal = {GERONTOLOGIST},
   Volume = {51},
   Pages = {32-32},
   Year = {2011},
   Key = {fds371170}
}

@article{fds277091,
   Author = {Han, X and Rozen, S and Boyle, SH and Hellegers, C and Cheng, H and Burke,
             JR and Welsh-Bohmer, KA and Doraiswamy, PM and Kaddurah-Daouk,
             R},
   Title = {Metabolomics in early Alzheimer's disease: identification of
             altered plasma sphingolipidome using shotgun
             lipidomics.},
   Journal = {PLoS One},
   Volume = {6},
   Number = {7},
   Pages = {e21643},
   Year = {2011},
   url = {http://www.ncbi.nlm.nih.gov/pubmed/21779331},
   Abstract = {BACKGROUND: The development of plasma biomarkers could
             facilitate early detection, risk assessment and therapeutic
             monitoring in Alzheimer's disease (AD). Alterations in
             ceramides and sphingomyelins have been postulated to play a
             role in amyloidogensis and inflammatory stress related
             neuronal apoptosis; however few studies have conducted a
             comprehensive analysis of the sphingolipidome in AD plasma
             using analytical platforms with accuracy, sensitivity and
             reproducibility. METHODS AND FINDINGS: We prospectively
             analyzed plasma from 26 AD patients (mean MMSE 21) and 26
             cognitively normal controls in a non-targeted approach using
             multi-dimensional mass spectrometry-based shotgun lipidomics
             to determine the levels of over 800 molecular species of
             lipids. These data were then correlated with diagnosis,
             apolipoprotein E4 genotype and cognitive performance. Plasma
             levels of species of sphingolipids were significantly
             altered in AD. Of the 33 sphingomyelin species tested, 8
             molecular species, particularly those containing long
             aliphatic chains such as 22 and 24 carbon atoms, were
             significantly lower (p<0.05) in AD compared to controls.
             Levels of 2 ceramide species (N16:0 and N21:0) were
             significantly higher in AD (p<0.05) with a similar, but
             weaker, trend for 5 other species. Ratios of ceramide to
             sphingomyelin species containing identical fatty acyl chains
             differed significantly between AD patients and controls.
             MMSE scores were correlated with altered mass levels of both
             N20:2 SM and OH-N25:0 ceramides (p<0.004) though lipid
             abnormalities were observed in mild and moderate AD. Within
             AD subjects, there were also genotype specific differences.
             CONCLUSIONS: In this prospective study, we used a sensitive
             multimodality platform to identify and characterize an
             essentially uniform but opposite pattern of disruption in
             sphingomyelin and ceramide mass levels in AD plasma. Given
             the role of brain sphingolipids in neuronal function, our
             findings provide new insights into the AD sphingolipidome
             and the potential use of metabolomic signatures as
             peripheral biomarkers.},
   Doi = {10.1371/journal.pone.0021643},
   Key = {fds277091}
}

@article{fds277105,
   Author = {Hayden, KM and Jones, RN and Zimmer, C and Plassman, BL and Browndyke,
             JN and Pieper, C and Warren, LH and Welsh-Bohmer,
             KA},
   Title = {Factor structure of the National Alzheimer's Coordinating
             Centers uniform dataset neuropsychological battery: an
             evaluation of invariance between and within groups over
             time.},
   Journal = {Alzheimer Dis Assoc Disord},
   Volume = {25},
   Number = {2},
   Pages = {128-137},
   Year = {2011},
   url = {http://www.ncbi.nlm.nih.gov/pubmed/21606904},
   Abstract = {The neuropsychological battery from the National Alzheimer's
             Disease Coordinating Center is designed to provide a
             sensitive assessment of mild cognitive disorders for
             multicenter investigations. Comprising 8 common
             neuropsychological tests (12 measures), the battery assesses
             cognitive domains affected early in the course of Alzheimer
             disease. We examined the factor structure of the battery
             across levels of cognition [normal, mild cognitive
             impairment, dementia] based on Clinical Dementia Rating
             scores to determine cognitive domains tapped by the battery.
             Using data pooled from 29 Alzheimer's Disease Centers funded
             by National Institute on Aging, exploratory factor analysis
             was used to derive a general model using half of the sample;
             4 factors representing memory, attention, executive
             function, and language were identified. Confirmatory factor
             analysis was used on the second half of the sample to
             evaluate invariance between groups and within groups over 1
             year. Factorial invariance testing included systematic
             addition of constraints and comparisons of nested models.
             The general confirmatory factor analysis model had a good
             fit. As constraints were added, model fit deteriorated
             slightly. Comparisons within groups showed stability over 1
             year. In a range of cognition from normal to dementia,
             factor structures and factor loadings will vary little.
             Further work is needed to determine whether domains become
             more or less distinct in severely cognitively compromised
             individuals.},
   Doi = {10.1097/WAD.0b013e3181ffa76d},
   Key = {fds277105}
}

@article{fds276915,
   Author = {Romero, HR and Hayes, SM and Welsh-bohmer, KA},
   Title = {Cognitive Domains Affected by Conditions of Ageing and the
             Role of Neuropsychological Testing},
   Pages = {389-396},
   Publisher = {JOHN WILEY & SONS LTD},
   Year = {2010},
   Month = {December},
   url = {http://dx.doi.org/10.1002/9780470669600.ch62},
   Abstract = {Determining the presence of a memory disorder in older
             patients can be challenging given the similarity between the
             complaints of benign brain aging and early brain diseases,
             such as Alzheimer's disease. In this chapter is we provide
             an overview of the cognitive domains assessed by a standard
             neuropsychological evaluation with focused consideration of
             the cognitive profiles common to normal brain aging,
             Alzheimer's disease, and geriatric depression. An
             understanding of the domains affected in these clinical
             contexts and the application of targeted neurocognitive
             measures of memory and executive processes can augment
             routine screening for neurodegenerative conditions and
             geriatric depression © 2011 John Wiley & Sons,
             Ltd.},
   Doi = {10.1002/9780470669600.ch62},
   Key = {fds276915}
}

@article{fds276991,
   Author = {Wee, CY and Yap, PT and Brownyke, JN and Potter, GG and Steffens, DC and Welsh-Bohmer, K and Wang, L and Shen, D},
   Title = {Accurate identification of MCI patients via enriched
             white-matter connectivity network},
   Journal = {Lecture Notes in Computer Science (including subseries
             Lecture Notes in Artificial Intelligence and Lecture Notes
             in Bioinformatics)},
   Volume = {6357 LNCS},
   Pages = {140-147},
   Publisher = {Springer Berlin Heidelberg},
   Year = {2010},
   Month = {October},
   ISSN = {0302-9743},
   url = {http://dx.doi.org/10.1007/978-3-642-15948-0_18},
   Abstract = {Mild cognitive impairment (MCI), often a prodromal phase of
             Alzheimer's disease (AD), is frequently considered to be a
             good target for early diagnosis and therapeutic
             interventions of AD. Recent emergence of reliable network
             characterization techniques have made understanding
             neurological disorders at a whole brain connectivity level
             possible. Accordingly, we propose a network-based
             multivariate classification algorithm, using a collection of
             measures derived from white-matter (WM) connectivity
             networks, to accurately identify MCI patients from normal
             controls. An enriched description of WM connections,
             utilizing six physiological parameters, i.e., fiber
             penetration count, fractional anisotropy (FA), mean
             diffusivity (MD), and principal diffusivities (λ 1, λ 2,
             λ 3), results in six connectivity networks for each subject
             to account for the connection topology and the biophysical
             properties of the connections. Upon parcellating the brain
             into 90 regions-of-interest (ROIs), the average statistics
             of each ROI in relation to the remaining ROIs are extracted
             as features for classification. These features are then
             sieved to select the most discriminant subset of features
             for building an MCI classifier via support vector machines
             (SVMs). Cross-validation results indicate better diagnostic
             power of the proposed enriched WM connection description
             than simple description with any single physiological
             parameter. © 2010 Springer-Verlag Berlin
             Heidelberg.},
   Doi = {10.1007/978-3-642-15948-0_18},
   Key = {fds276991}
}

@article{fds276964,
   Author = {Roses, AD and Lutz, MW and Amrine-Madsen, H and Saunders, AM and Crenshaw, DG and Sundseth, SS and Huentelman, MJ and Welsh-Bohmer,
             KA and Reiman, EM},
   Title = {A TOMM40 variable-length polymorphism predicts the age of
             late-onset Alzheimer's disease.},
   Journal = {Pharmacogenomics J},
   Volume = {10},
   Number = {5},
   Pages = {375-384},
   Year = {2010},
   Month = {October},
   url = {http://www.ncbi.nlm.nih.gov/pubmed/20029386},
   Abstract = {The ɛ4 allele of the apolipoprotein E (APOE) gene is
             currently the strongest and most highly replicated genetic
             factor for risk and age of onset of late-onset Alzheimer's
             disease (LOAD). Using phylogenetic analysis, we have
             identified a polymorphic poly-T variant, rs10524523, in the
             translocase of outer mitochondrial membrane 40 homolog
             (TOMM40) gene that provides greatly increased precision in
             the estimation of age of LOAD onset for APOE ɛ3 carriers.
             In two independent clinical cohorts, longer lengths of
             rs10524523 are associated with a higher risk for LOAD. For
             APOE ɛ3/4 patients who developed LOAD after 60 years of
             age, individuals with long poly-T repeats linked to APOE ɛ3
             develop LOAD on an average of 7 years earlier than
             individuals with shorter poly-T repeats linked to APOE ɛ3
             (70.5 ± 1.2 years versus 77.6 ± 2.1 years, P=0.02, n=34).
             Independent mutation events at rs10524523 that occurred
             during Caucasian evolution have given rise to multiple
             categories of poly-T length variants at this locus. On
             replication, these results will have clinical utility for
             predictive risk estimates for LOAD and for enabling clinical
             disease prevention studies. In addition, these results show
             the effective use of a phylogenetic approach for analysis of
             haplotypes of polymorphisms, including structural
             polymorphisms, which contribute to complex
             diseases.},
   Doi = {10.1038/tpj.2009.69},
   Key = {fds276964}
}

@article{fds371171,
   Author = {Bradford, D and Smith, K and Schwartz, S and Tschanz, J and Ostbye, T and Corcoran, C and Welsh-Bohmer, K and Norton, MC},
   Title = {TIMING OF OFFSPRING DEATH IN PARENTS' LIVES AND LATE-LIFE
             COGNITIVE DECLINE. THE CACHE COUNTY STUDY},
   Journal = {GERONTOLOGIST},
   Volume = {50},
   Pages = {462-462},
   Publisher = {OXFORD UNIV PRESS INC},
   Year = {2010},
   Month = {October},
   Key = {fds371171}
}

@article{fds371172,
   Author = {Norton, MC and Bradford, D and Smith, K and Schwartz, S and Tschanz, J and Ostbye, T and Corcoran, C and Welsh-Bohmer, K},
   Title = {FAMILY SIZE MODERATES THE ASSOCIATION BETWEEN OFFSPRING
             DEATH AND RATE OF COGNITIVE DECLINE. THE CACHE COUNTY
             STUDY},
   Journal = {GERONTOLOGIST},
   Volume = {50},
   Pages = {464-464},
   Publisher = {OXFORD UNIV PRESS INC},
   Year = {2010},
   Month = {October},
   Key = {fds371172}
}

@article{fds371173,
   Author = {Tschanz, J and Pfister, R and Steffens, DC and Corcoran, C and Smith, K and Ostbye, T and Welsh-Bohmer, K and Norton, MC},
   Title = {STRESSFUL EVENTS IN LATE-LIFE: EFFECTS ON COGNITIVE DECLINE.
             THE CACHE COUNTY STUDY},
   Journal = {GERONTOLOGIST},
   Volume = {50},
   Pages = {354-354},
   Publisher = {OXFORD UNIV PRESS INC},
   Year = {2010},
   Month = {October},
   Key = {fds371173}
}

@article{fds277047,
   Author = {Hayden, KM and Breitner, JCS and Welsh-Bohmer,
             KA},
   Title = {Apolipoprotein E ε4 status and cognitive decline with and
             without dementia - Reply},
   Journal = {Archives of Neurology},
   Volume = {67},
   Number = {8},
   Pages = {1036-1037},
   Publisher = {American Medical Association (AMA)},
   Year = {2010},
   Month = {August},
   ISSN = {0003-9942},
   url = {http://dx.doi.org/10.1001/archneurol.2010.164},
   Doi = {10.1001/archneurol.2010.164},
   Key = {fds277047}
}

@article{fds276963,
   Author = {Buckley, T and Norton, MC and Deberard, MS and Welsh-Bohmer, KA and Tschanz, JT},
   Title = {A brief metacognition questionnaire for the elderly:
             comparison with cognitive performance and informant ratings
             the Cache County Study.},
   Journal = {Int J Geriatr Psychiatry},
   Volume = {25},
   Number = {7},
   Pages = {739-747},
   Year = {2010},
   Month = {July},
   url = {http://www.ncbi.nlm.nih.gov/pubmed/19823990},
   Abstract = {OBJECTIVE: To examine the utility of a brief, metacognition
             questionnaire by examining its association with objective
             cognitive testing and informant ratings. We hypothesized
             that the association between self-ratings of change and both
             outcomes would be greater among individuals without dementia
             than among those with dementia. METHODS: Participants were
             535 persons without dementia and 152 with dementia from the
             Cache County Memory Study who had completed a metacognition
             questionnaire, two administrations of the Modified
             Mini-Mental State Exam (3 MS) and who had data on the
             Informant Questionnaire of Cognitive Decline in the Elderly
             (IQCODE). Cronbach's alpha was calculated as a measure of
             internal consistency of the metacognition questionnaire.
             Multiple regression was used to examine the relationship
             between metacognition and 3 MS change. Logistic regression
             was used to examine the relationship between metacognition
             and IQCODE ratings (no change vs. worse). RESULTS:
             Cronbach's alpha was 0.75. Among individuals without
             dementia, metacognition significantly predicted 3 MS change
             (p = .027) and IQCODE ratings (OR = 4.0, 95% CI = 1.2-13.8,
             p = .029), suggesting consistency among measures. For those
             with dementia, there was a weak, inverse relationship
             between 3 MS change and metacognition (r = -0.16, p = .056).
             IQCODE ratings were not significantly associated with
             metacognition (p = .729). Degree of dementia severity did
             not modify the relationship between metacognition and either
             outcome (p > .05). CONCLUSIONS: We demonstrated adequate
             internal consistency and evidence for validity of a brief
             metacognition questionnaire. The questionnaire may provide a
             useful adjunct to memory and functional assessments for
             assessing anosognosia in elderly populations.},
   Doi = {10.1002/gps.2416},
   Key = {fds276963}
}

@article{fds277102,
   Author = {Dennis, NA and Browndyke, JN and Stokes, J and Need, A and Burke, JR and Welsh-Bohmer, KA and Cabeza, R},
   Title = {Temporal lobe functional activity and connectivity in young
             adult APOE varepsilon4 carriers.},
   Journal = {Alzheimers Dement},
   Volume = {6},
   Number = {4},
   Pages = {303-311},
   Year = {2010},
   Month = {July},
   url = {http://www.ncbi.nlm.nih.gov/pubmed/19744893},
   Abstract = {BACKGROUND: We sought to determine if the APOE epsilon4
             allele influences both the functional activation and
             connectivity of the medial temporal lobes (MTLs) during
             successful memory encoding in young adults. METHODS:
             Twenty-four healthy young adults, i.e., 12 carriers and 12
             noncarriers of the APOE epsilon4 allele, were scanned in a
             subsequent-memory paradigm, using event-related functional
             magnetic resonance imaging. The neuroanatomic correlates of
             successful encoding were measured as greater neural activity
             for subsequently remembered versus forgotten task items, or
             in short, encoding success activity (ESA). Group differences
             in ESA within the MTLs, as well as whole-brain functional
             connectivity with the MTLs, were assessed. RESULTS: In the
             absence of demographic or performance differences, APOE
             epsilon4 allele carriers exhibited greater bilateral MTL
             activity relative to noncarriers while accomplishing the
             same encoding task. Moreover, whereas epsilon4 carriers
             demonstrated a greater functional connectivity of
             ESA-related MTL activity with the posterior cingulate and
             other peri-limbic regions, reductions in overall
             connectivity were found across the anterior and posterior
             cortices. CONCLUSIONS: These results suggest that the APOE
             varepsilon4 allele may influence not only functional
             activations within the MTL, but functional connectivity of
             the MTLs to other regions implicated in memory encoding.
             Enhanced functional connectivity of the MTLs with the
             posterior cingulate in young adult epsilon4 carriers
             suggests that APOE may be expressed early in brain regions
             known to be involved in Alzheimer's disease, long before
             late-onset dementia is a practical risk or consideration.
             These functional connectivity differences may also reflect
             pleiotropic effects of APOE during early
             development.},
   Doi = {10.1016/j.jalz.2009.07.003},
   Key = {fds277102}
}

@article{fds277114,
   Author = {Smith, PJ and Blumenthal, JA and Babyak, MA and Craighead, L and Welsh-Bohmer, KA and Browndyke, JN and Strauman, TA and Sherwood,
             A},
   Title = {Effects of the dietary approaches to stop hypertension diet,
             exercise, and caloric restriction on neurocognition in
             overweight adults with high blood pressure.},
   Journal = {Hypertension},
   Volume = {55},
   Number = {6},
   Pages = {1331-1338},
   Year = {2010},
   Month = {June},
   url = {http://www.ncbi.nlm.nih.gov/pubmed/20305128},
   Abstract = {High blood pressure increases the risks of stroke, dementia,
             and neurocognitive dysfunction. Although aerobic exercise
             and dietary modifications have been shown to reduce blood
             pressure, no randomized trials have examined the effects of
             aerobic exercise combined with dietary modification on
             neurocognitive functioning in individuals with high blood
             pressure (ie, prehypertension and stage 1 hypertension). As
             part of a larger investigation, 124 participants with
             elevated blood pressure (systolic blood pressure 130 to 159
             mm Hg or diastolic blood pressure 85 to 99 mm Hg) who were
             sedentary and overweight or obese (body mass index: 25 to 40
             kg/m(2)) were randomized to the Dietary Approaches to Stop
             Hypertension (DASH) diet alone, DASH combined with a
             behavioral weight management program including exercise and
             caloric restriction, or a usual diet control group.
             Participants completed a battery of neurocognitive tests of
             executive function-memory-learning and psychomotor speed at
             baseline and again after the 4-month intervention.
             Participants on the DASH diet combined with a behavioral
             weight management program exhibited greater improvements in
             executive function-memory-learning (Cohen's D=0.562;
             P=0.008) and psychomotor speed (Cohen's D=0.480; P=0.023),
             and DASH diet alone participants exhibited better
             psychomotor speed (Cohen's D=0.440; P=0.036) compared with
             the usual diet control. Neurocognitive improvements appeared
             to be mediated by increased aerobic fitness and weight loss.
             Also, participants with greater intima-medial thickness and
             higher systolic blood pressure showed greater improvements
             in executive function-memory-learning in the group on the
             DASH diet combined with a behavioral weight management
             program. In conclusion, combining aerobic exercise with the
             DASH diet and caloric restriction improves neurocognitive
             function among sedentary and overweight/obese individuals
             with prehypertension and hypertension.},
   Doi = {10.1161/HYPERTENSIONAHA.109.146795},
   Key = {fds277114}
}

@article{fds277073,
   Author = {Hayden, KM and Norton, MC and Darcey, D and Ostbye, T and Zandi, PP and Breitner, JCS and Welsh-Bohmer, KA and Cache County Study
             Investigators},
   Title = {Occupational exposure to pesticides increases the risk of
             incident AD: the Cache County study.},
   Journal = {Neurology},
   Volume = {74},
   Number = {19},
   Pages = {1524-1530},
   Year = {2010},
   Month = {May},
   url = {http://www.ncbi.nlm.nih.gov/pubmed/20458069},
   Abstract = {BACKGROUND: Commonly used organophosphate and organochlorine
             pesticides inhibit acetylcholinesterase at synapses in the
             somatic, autonomic, and central nervous systems and may
             therefore have lasting effects on the nervous system. Few
             studies have examined the relationship of pesticide exposure
             and risk of dementia or Alzheimer disease (AD). We sought to
             examine the association of occupational pesticide exposure
             and the risk of incident dementia and AD in later life.
             METHODS: Residents of the agricultural community of Cache
             County, UT, who were aged 65 years and older as of January
             1995, were invited to participate in the study. At baseline,
             participants completed detailed occupational history
             questionnaires that included information about exposures to
             various types of pesticides. Cognitive status was assessed
             at baseline and after 3, 7, and 10 years. Standardized
             methods were used for detection and diagnosis of dementia
             and AD. Cox proportional hazards survival analyses were used
             to evaluate the risk of incident dementia and AD associated
             with pesticide exposure. RESULTS: Among 3,084 enrollees
             without dementia, more men than women reported pesticide
             exposure (p < 0.0001). Exposed individuals (n = 572) had
             more years of education (p < 0.01) but did not differ from
             others in age. Some 500 individuals developed incident
             dementia, 344 with AD. After adjustment for baseline age,
             sex, education, APOE epsilon4 status, and baseline Modified
             Mini-Mental State Examination scores, Cox proportional
             hazards models showed increased risks among
             pesticide-exposed individuals for all-cause dementia, with
             hazard ratio (HR) 1.38 and 95% confidence interval (CI)
             1.09-1.76, and for AD (HR 1.42, 95% CI 1.06-1.91). The risk
             of AD associated with organophosphate exposure (HR 1.53, 95%
             CI 1.05-2.23) was slightly higher than the risk associated
             with organochlorines (HR 1.49, 95% CI 0.99-2.24), which was
             nearly significant. CONCLUSIONS: Pesticide exposure may
             increase the risk of dementia and Alzheimer disease in late
             life.},
   Doi = {10.1212/WNL.0b013e3181dd4423},
   Key = {fds277073}
}

@article{fds276981,
   Author = {Norton, MC and Smith, KR and Østbye, T and Tschanz, JT and Corcoran, C and Schwartz, S and Piercy, KW and Rabins, PV and Steffens, DC and Skoog, I and Breitner, JCS and Welsh-Bohmer, KA and Cache County
             Investigators},
   Title = {Greater risk of dementia when spouse has dementia? The Cache
             County study.},
   Journal = {J Am Geriatr Soc},
   Volume = {58},
   Number = {5},
   Pages = {895-900},
   Year = {2010},
   Month = {May},
   url = {http://www.ncbi.nlm.nih.gov/pubmed/20722820},
   Abstract = {OBJECTIVES: To examine the effects of caring for a spouse
             with dementia on the caregiver's risk for incident dementia.
             DESIGN: Population-based study of incident dementia in
             spouses of persons with dementia. SETTING: Rural county in
             northern Utah. PARTICIPANTS: Two thousand four hundred
             forty-two subjects (1,221 married couples) aged 65 and
             older. MEASUREMENTS: Incident dementia was diagnosed in 255
             subjects, with onset defined as age when subject met
             Diagnostic and Statistical Manual of Mental Disorders, Third
             Edition, Revised, criteria for dementia. Cox proportional
             hazards regression tested the effect of time-dependent
             exposure to dementia in one's spouse, adjusted for potential
             confounders. RESULTS: A subject whose spouse experienced
             incident dementia onset had a six times greater risk for
             incident dementia as subjects whose spouses were dementia
             free (hazard rate ratio (HRR)=6.0, 95% confidence interval
             (CI)=2.2-16.2, P<.001). In sex-specific analyses, husbands
             had higher risks (HRR=11.9, 95% CI=1.7-85.5, P=.01) than
             wives (HRR=3.7, 95% CI=1.2-11.6, P=.03). CONCLUSION: The
             chronic and often severe stress associated with dementia
             caregiving may exert substantial risk for the development of
             dementia in spouse caregivers. Additional (not mutually
             exclusive) explanations for findings are
             discussed.},
   Doi = {10.1111/j.1532-5415.2010.02806.x},
   Key = {fds276981}
}

@article{fds277112,
   Author = {Smith, PJ and Blumenthal, JA and Hoffman, BM and Cooper, H and Strauman,
             TA and Welsh-Bohmer, K and Browndyke, JN and Sherwood,
             A},
   Title = {Aerobic exercise and neurocognitive performance: a
             meta-analytic review of randomized controlled
             trials.},
   Journal = {Psychosom Med},
   Volume = {72},
   Number = {3},
   Pages = {239-252},
   Year = {2010},
   Month = {April},
   url = {http://www.ncbi.nlm.nih.gov/pubmed/20223924},
   Abstract = {OBJECTIVES: To assess the effects of aerobic exercise
             training on neurocognitive performance. Although the effects
             of exercise on neurocognition have been the subject of
             several previous reviews and meta-analyses, they have been
             hampered by methodological shortcomings and are now outdated
             as a result of the recent publication of several
             large-scale, randomized, controlled trials (RCTs). METHODS:
             We conducted a systematic literature review of RCTs
             examining the association between aerobic exercise training
             on neurocognitive performance between January 1966 and July
             2009. Suitable studies were selected for inclusion according
             to the following criteria: randomized treatment allocation;
             mean age > or =18 years of age; duration of treatment >1
             month; incorporated aerobic exercise components; supervised
             exercise training; the presence of a nonaerobic-exercise
             control group; and sufficient information to derive effect
             size data. RESULTS: Twenty-nine studies met inclusion
             criteria and were included in our analyses, representing
             data from 2049 participants and 234 effect sizes.
             Individuals randomly assigned to receive aerobic exercise
             training demonstrated modest improvements in attention and
             processing speed (g = 0.158; 95% confidence interval [CI];
             0.055-0.260; p = .003), executive function (g = 0.123; 95%
             CI, 0.021-0.225; p = .018), and memory (g = 0.128; 95% CI,
             0.015-0.241; p = .026). CONCLUSIONS: Aerobic exercise
             training is associated with modest improvements in attention
             and processing speed, executive function, and memory,
             although the effects of exercise on working memory are less
             consistent. Rigorous RCTs are needed with larger samples,
             appropriate controls, and longer follow-up
             periods.},
   Doi = {10.1097/PSY.0b013e3181d14633},
   Key = {fds277112}
}

@article{fds277025,
   Author = {Van Deerlin and VM and Sleiman, PMA and Martinez-Lage, M and Chen-Plotkin, A and Wang, L-S and Graff-Radford, NR and Dickson, DW and Rademakers, R and Boeve, BF and Grossman, M and Arnold, SE and Mann,
             DMA and Pickering-Brown, SM and Seelaar, H and Heutink, P and van
             Swieten, JC and Murrell, JR and Ghetti, B and Spina, S and Grafman, J and Hodges, J and Spillantini, MG and Gilman, S and Lieberman, AP and Kaye,
             JA and Woltjer, RL and Bigio, EH and Mesulam, M and Al-Sarraj, S and Troakes, C and Rosenberg, RN and White, CL and Ferrer, I and Lladó, A and Neumann, M and Kretzschmar, HA and Hulette, CM and Welsh-Bohmer, KA and Miller, BL and Alzualde, A and Lopez de Munain and A and McKee, AC and Gearing, M and Levey, AI and Lah, JJ and Hardy, J and Rohrer, JD and Lashley, T and Mackenzie, IRA and Feldman, HH and Hamilton, RL and Dekosky, ST and van der Zee, J and Kumar-Singh, S and Van
             Broeckhoven, C and Mayeux, R and Vonsattel, JPG and Troncoso, JC and Kril, JJ and Kwok, JBJ and Halliday, GM and Bird, TD and Ince, PG and Shaw,
             PJ and Cairns, NJ and Morris, JC and McLean, CA and DeCarli, C and Ellis,
             WG and Freeman, SH and Frosch, MP and Growdon, JH and Perl, DP and Sano, M and Bennett, DA and Schneider, JA and Beach, TG and Reiman, EM and Woodruff,
             BK and Cummings, J and Vinters, HV and Miller, CA and Chui, HC and Alafuzoff, I and Hartikainen, P and Seilhean, D and Galasko, D and Masliah, E and Cotman, CW and Tuñón, MT and Martínez, MCC and Munoz,
             DG and Carroll, SL and Marson, D and Riederer, PF and Bogdanovic, N and Schellenberg, GD and Hakonarson, H and Trojanowski, JQ and Lee,
             VM-Y},
   Title = {Common variants at 7p21 are associated with frontotemporal
             lobar degeneration with TDP-43 inclusions.},
   Journal = {Nat Genet},
   Volume = {42},
   Number = {3},
   Pages = {234-239},
   Year = {2010},
   Month = {March},
   url = {http://www.ncbi.nlm.nih.gov/pubmed/20154673},
   Abstract = {Frontotemporal lobar degeneration (FTLD) is the second most
             common cause of presenile dementia. The predominant
             neuropathology is FTLD with TAR DNA-binding protein (TDP-43)
             inclusions (FTLD-TDP). FTLD-TDP is frequently familial,
             resulting from mutations in GRN (which encodes progranulin).
             We assembled an international collaboration to identify
             susceptibility loci for FTLD-TDP through a genome-wide
             association study of 515 individuals with FTLD-TDP. We found
             that FTLD-TDP associates with multiple SNPs mapping to a
             single linkage disequilibrium block on 7p21 that contains
             TMEM106B. Three SNPs retained genome-wide significance
             following Bonferroni correction (top SNP rs1990622, P = 1.08
             x 10(-11); odds ratio, minor allele (C) 0.61, 95% CI
             0.53-0.71). The association replicated in 89 FTLD-TDP cases
             (rs1990622; P = 2 x 10(-4)). TMEM106B variants may confer
             risk of FTLD-TDP by increasing TMEM106B expression. TMEM106B
             variants also contribute to genetic risk for FTLD-TDP in
             individuals with mutations in GRN. Our data implicate
             variants in TMEM106B as a strong risk factor for FTLD-TDP,
             suggesting an underlying pathogenic mechanism.},
   Doi = {10.1038/ng.536},
   Key = {fds277025}
}

@article{fds277068,
   Author = {Sheline, YI and Pieper, CF and Barch, DM and Welsh-Bohmer, K and McKinstry, RC and MacFall, JR and D'Angelo, G and Garcia, KS and Gersing, K and Wilkins, C and Taylor, W and Steffens, DC and Krishnan,
             RR and Doraiswamy, PM},
   Title = {Support for the vascular depression hypothesis in late-life
             depression: results of a 2-site, prospective, antidepressant
             treatment trial.},
   Journal = {Arch Gen Psychiatry},
   Volume = {67},
   Number = {3},
   Pages = {277-285},
   Year = {2010},
   Month = {March},
   url = {http://www.ncbi.nlm.nih.gov/pubmed/20194828},
   Abstract = {CONTEXT: Research on vascular depression has used 2
             approaches to subtype late-life depression, based on
             executive dysfunction or white matter hyperintensity
             severity. OBJECTIVE: To evaluate the relationship of
             neuropsychological performance and white matter
             hyperintensity with clinical response in late-life
             depression. DESIGN: Two-site, prospective, nonrandomized
             controlled trial. SETTING: Outpatient clinics at Washington
             University and Duke University. PARTICIPANTS: A total of 217
             subjects aged 60 years or older met DSM-IV criteria for
             major depression, scored 20 or more on the Montgomery-Asberg
             Depression Rating Scale (MADRS), and received vascular risk
             factor scores, neuropsychological testing, and magnetic
             resonance imaging; they were excluded for cognitive
             impairment or severe medical disorders. Fazekas rating was
             conducted to grade white matter hyperintensity lesions.
             Intervention Twelve weeks of sertraline treatment, titrated
             by clinical response. Main Outcome Measure Participants'
             MADRS scores over time. RESULTS: Baseline neuropsychological
             factor scores correlated negatively with baseline Fazekas
             scores. A mixed model examined effects of predictor
             variables on MADRS scores over time. Baseline episodic
             memory (P = .002), language (P = .007), working memory (P =
             .01), processing speed (P < .001), executive function factor
             scores (P = .002), and categorical Fazekas ratings (P = .05)
             predicted MADRS scores, controlling for age, education, age
             of onset, and race. Controlling for baseline MADRS scores,
             these factors remained significant predictors of decrease in
             MADRS scores, except for working memory and Fazekas ratings.
             Thirty-three percent of subjects achieved remission (MADRS <
             or =7). Remitters differed from nonremitters in baseline
             cognitive processing speed, executive function, language,
             episodic memory, and vascular risk factor scores.
             CONCLUSIONS: Comprehensive neuropsychological function and
             white matter hyperintensity severity predicted MADRS scores
             prospectively over a 12-week treatment course with selective
             serotonin reuptake inhibitors in late-life depression.
             Baseline neuropsychological function differentiated
             remitters from nonremitters and predicted time to remission
             in a proportional hazards model. Predictor variables
             correlated highly with vascular risk factor severity. These
             data support the vascular depression hypothesis and
             highlight the importance of linking subtypes based on
             neuropsychological function and white matter integrity.
             TRIAL REGISTRATION: clinicaltrials.gov Identifier:
             NCT00045773.},
   Doi = {10.1001/archgenpsychiatry.2009.204},
   Key = {fds277068}
}

@article{fds277045,
   Author = {Welsh-Bohmer, KA and Plassman, BL and Hayden, KM},
   Title = {Genetic and environmental contributions to cognitive decline
             in aging and Alzheimer's disease},
   Journal = {Annual Review of Gerontology and Geriatrics},
   Volume = {30},
   Number = {1},
   Pages = {81-114},
   Publisher = {Springer Publishing Company},
   Year = {2010},
   Month = {January},
   ISSN = {0198-8794},
   url = {http://dx.doi.org/10.1891/0198-8794.30.81},
   Abstract = {Inheritance appears to play a strong role in terms of human
             longevity and also in risk for chronic neurodegenerative
             diseases of late life such as Alzheimer's disease (AD).
             Understanding the role of genes in normal biological aging
             of the nervous system and in the expression of AD
             pathological changes is important for developing useful
             disease targets for clinical intervention and treatment. The
             extent to which gene expression can be altered through
             environmental exposures may also provide important clues for
             disease prevention. In this chapter, we consider the role of
             genetic and environmental factors in AD risk and in
             age-related cognitive decline. We focus, in particular, on
             targets that are potentially modifiable with respect to
             exerting positive effects on cognition and AD risk. In this
             context, we conclude the chapter by considering lifestyle
             approaches to disease prevention and enhancement of
             successful cognitive aging that extend from the current
             state of evidence and discuss areas for future research
             development. © 2010 Springer Publishing
             Company.},
   Doi = {10.1891/0198-8794.30.81},
   Key = {fds277045}
}

@article{fds371174,
   Author = {Romero, HR and Browndyke, JN and Burke, JR and Hayden, KM and Welsh-Bohmer, KA},
   Title = {Executive Measures are Related to Functional Ability in an
             Ethnically Diverse Cohort},
   Journal = {CLINICAL NEUROPSYCHOLOGIST},
   Volume = {24},
   Number = {4},
   Pages = {599-600},
   Publisher = {TAYLOR & FRANCIS INC},
   Year = {2010},
   Month = {January},
   Key = {fds371174}
}

@article{fds277046,
   Author = {Chuang, Y-F and Hayden, KM and Norton, MC and Tschanz, J and Breitner,
             JCS and Welsh-Bohmer, KA and Zandi, PP},
   Title = {Association between APOE epsilon4 allele and vascular
             dementia: The Cache County study.},
   Journal = {Dement Geriatr Cogn Disord},
   Volume = {29},
   Number = {3},
   Pages = {248-253},
   Year = {2010},
   url = {http://www.ncbi.nlm.nih.gov/pubmed/20375505},
   Abstract = {BACKGROUND: The APOE epsilon4 allele is an established risk
             factor for Alzheimer's disease, but reports of its
             association with vascular dementia (VaD) have been
             inconsistent. We examined the relationship between APOE
             epsilon4 allele and the risk of incident VaD in a large,
             population-based cohort of elderly adults with up to 10
             years of follow-up between 1995 and 2005. METHODS: A total
             of 3,424 elderly men and women free of dementia were
             genotyped at the baseline assessment. Incident VaD was
             identified through standardized procedures administered at 3
             follow-up assessments. Cox proportional hazards models were
             used to evaluate the risk of VaD associated with APOE
             epsilon4. RESULTS: The adjusted hazard ratio was 1.6 for the
             participants with 1 APOE epsilon4 allele (95% CI: 0.9-2.7; p
             = 0.083) and 4.4 for those with 2 APOE epsilon4 alleles (95%
             CI: 1.6-12.5; p = 0.005). The increased risk did not appear
             to be mediated by vascular risk factors. CONCLUSIONS: The
             APOE epsilon4 allele is associated with an increased risk of
             VaD in a dose-dependent fashion and accounts for almost 20%
             of VaD in the population.},
   Doi = {10.1159/000285166},
   Key = {fds277046}
}

@article{fds277089,
   Author = {Heinzen, EL and Need, AC and Hayden, KM and Chiba-Falek, O and Roses,
             AD and Strittmatter, WJ and Burke, JR and Hulette, CM and Welsh-Bohmer,
             KA and Goldstein, DB},
   Title = {Genome-wide scan of copy number variation in late-onset
             Alzheimer's disease.},
   Journal = {J Alzheimers Dis},
   Volume = {19},
   Number = {1},
   Pages = {69-77},
   Year = {2010},
   url = {http://www.ncbi.nlm.nih.gov/pubmed/20061627},
   Abstract = {Alzheimer's disease is a complex and progressive
             neurodegenerative disease leading to loss of memory,
             cognitive impairment, and ultimately death. To date, six
             large-scale genome-wide association studies have been
             conducted to identify SNPs that influence disease
             predisposition. These studies have confirmed the well-known
             APOE epsilon4 risk allele, identified a novel variant that
             influences disease risk within the APOE epsilon4 population,
             found a SNP that modifies the age of disease onset, as well
             as reported the first sex-linked susceptibility variant.
             Here we report a genome-wide scan of Alzheimer's disease in
             a set of 331 cases and 368 controls, extending analyses for
             the first time to include assessments of copy number
             variation. In this analysis, no new SNPs show genome-wide
             significance. We also screened for effects of copy number
             variation, and while nothing was significant, a duplication
             in CHRNA7 appears interesting enough to warrant further
             investigation.},
   Doi = {10.3233/JAD-2010-1212},
   Key = {fds277089}
}

@article{fds376500,
   Author = {Sheline, and Welsh-Bohmer, K},
   Title = {Support for the Vascular Depression Hypothesis in Late-Life
             Depression: Results of a 2-Site, Prospective, Antidepressant
             Treatment Trial (vol 67, pg 277, 2010)},
   Journal = {ARCHIVES OF GENERAL PSYCHIATRY},
   Volume = {67},
   Number = {10},
   Pages = {1043-1043},
   Year = {2010},
   Key = {fds376500}
}

@article{fds277044,
   Author = {Hayden, KM and Zandi, PP and West, NA and Tschanz, JT and Norton, MC and Corcoran, C and Breitner, JCS and Welsh-Bohmer, KA and Cache County
             Study Group},
   Title = {Effects of family history and apolipoprotein E epsilon4
             status on cognitive decline in the absence of Alzheimer
             dementia: the Cache County Study.},
   Journal = {Arch Neurol},
   Volume = {66},
   Number = {11},
   Pages = {1378-1383},
   Year = {2009},
   Month = {November},
   url = {http://www.ncbi.nlm.nih.gov/pubmed/19901170},
   Abstract = {OBJECTIVE: To evaluate the influences of a family history of
             Alzheimer dementia (FHxAD) and the apolipoprotein E epsilon4
             genotype (APOE epsilon4) on cognitive decline. DESIGN,
             SETTING, AND PARTICIPANTS: Residents of Cache County, Utah,
             aged 65 years or older, were invited to participate. At
             baseline, 2957 participants provided DNA for genotyping of
             APOE and a detailed FHxAD. They also completed the Modified
             Mini-Mental State Examination. Cognitive status was
             reexamined after 3 and 7 years. We used mixed-effects models
             to examine the association among FHxAD, APOE epsilon4, and
             cognitive trajectories. MAIN OUTCOME MEASURE: Modified
             Mini-Mental State Examination score trajectories over time.
             RESULTS: Compared with participants who did not have APOE
             epsilon4 or an FHxAD, those with APOE epsilon4 scored lower
             on the Modified Mini-Mental State Examination at baseline
             (-0.70 points; 95% confidence interval [CI], -1.15 to
             -0.24). Participants with an FHxAD and APOE epsilon4
             differed less, if at all, in baseline score (-0.46 points;
             95% CI, -1.09 to 0.16) but declined faster during the 7-year
             study (-9.75 points [95% CI, -10.82 to -8.67] vs -2.91
             points [95% CI, -3.37 to -2.44]). After exclusion of
             participants who developed prodromal AD or incident
             dementia, the group with an FHxAD and APOE epsilon4 declined
             much less during the 7-year study (-1.54; 95% CI, -2.59 to
             -0.50). CONCLUSIONS: Much of the association among FHxAD,
             APOE epsilon4, and cognitive decline may be attributed to
             undetected incipient (latent) disease. In the absence of
             latent disease, the 2 factors do not appear individually to
             be associated with cognitive decline, although they may be
             additive.},
   Doi = {10.1001/archneurol.2009.237},
   Key = {fds277044}
}

@article{fds277101,
   Author = {Linnertz, C and Saucier, L and Ge, D and Cronin, KD and Burke, JR and Browndyke, JN and Hulette, CM and Welsh-Bohmer, KA and Chiba-Falek,
             O},
   Title = {Genetic regulation of alpha-synuclein mRNA expression in
             various human brain tissues.},
   Journal = {PLoS One},
   Volume = {4},
   Number = {10},
   Pages = {e7480},
   Year = {2009},
   Month = {October},
   url = {http://www.ncbi.nlm.nih.gov/pubmed/19834617},
   Abstract = {Genetic variability across the SNCA locus has been
             repeatedly associated with susceptibility to sporadic
             Parkinson's disease (PD). Accumulated evidence emphasizes
             the importance of SNCA dosage and expression levels in PD
             pathogenesis. However whether genetic variability in the
             SNCA gene modulates the risk to develop sporadic PD via
             regulation of SNCA expression remained elusive. We studied
             the effect of PD risk-associated variants at SNCA 5' and
             3'regions on SNCA-mRNA levels in vivo in 228 human brain
             samples from three structures differentially vulnerable to
             PD pathology (substantia-nigra, temporal- and
             frontal-cortex) obtained from 144 neurologically normal
             cadavers. The extensively characterized PD-associated
             promoter polymorphism, Rep1, had an effect on SNCA-mRNA
             levels. Homozygous genotype of the 'protective', Rep1-259 bp
             allele, was associated with lower levels of SNCA-mRNA
             relative to individuals that carried at least one copy of
             the PD-risk associated alleles, amounting to an average
             decrease of approximately 40% and >50% in temporal-cortex
             and substantia-nigra, respectively. Furthermore, SNPs
             tagging the SNCA 3'-untranslated-region also showed effects
             on SNCA-mRNA levels in both the temporal-cortex and the
             substantia-nigra, although, in contrast to Rep1, the
             'decreased-risk' alleles were correlated with increased
             SNCA-mRNA levels. Similar to Rep1 findings, no difference in
             SNCA-mRNA level was seen with different SNCA 3'SNP alleles
             in the frontal-cortex, indicating there is brain-region
             specificity of the genetic regulation of SNCA expression. We
             provide evidence for functional consequences of
             PD-associated SNCA gene variants in disease relevant brain
             tissues, suggesting that genetic regulation of SNCA
             expression plays an important role in the development of the
             disease.},
   Doi = {10.1371/journal.pone.0007480},
   Key = {fds277101}
}

@article{fds276990,
   Author = {Potter, GG and Taylor, WD and McQuoid, DR and Steffens, DC and Welsh-Bohmer, KA and Krishnan, KRR},
   Title = {The COMT Val158Met polymorphism and cognition in depressed
             and nondepressed older adults.},
   Journal = {Int J Geriatr Psychiatry},
   Volume = {24},
   Number = {10},
   Pages = {1127-1133},
   Year = {2009},
   Month = {October},
   url = {http://www.ncbi.nlm.nih.gov/pubmed/19296553},
   Abstract = {OBJECTIVE: The objective of the current study was to examine
             the relationship between the COMT Val(158)Met polymorphism
             and neuropsychological performance in depressed and
             nondepressed older adults. METHODS: One hundred and
             twenty-six clinically depressed older adults and 105
             nondepressed comparison participants were compared on
             neuropsychological performance and COMT Val(158)Met
             (Val/Val, Val/Met, Met/Met). RESULTS: Based on multivariate
             regression models, the COMT Val(158)Met polymorphism was not
             associated with cognitive performance among depressed or
             nondepressed individuals, nor did this polymorphism account
             for the fact that depressed individuals performed worse than
             nondepressed individuals on several neuropsychological tests
             that are typically affected by depression. There was also no
             difference in frequency of the COMT Val(158)Met alleles
             between depressed and nondepressed individuals. CONCLUSIONS:
             Although the current study found no association between COMT
             Val(158)Met polymorphism on a number of clinical
             neuropsychological tests that are typically found to be
             sensitive to depression, differential effects of the COMT
             Val(158)Met polymorphism on dopamine transmission in
             psychiatric and non-psychiatric populations may be further
             clarified by clinical research with neuroscience-based
             paradigms that segregate cognitive tasks into component
             processes with precise neural substrates, particularly with
             respect to the complex functions of the prefrontal cortex.
             Negative results can be important to narrowing down target
             processes and understanding the influence of clinical and
             demographic characteristics in studies of psychiatric
             genetics.},
   Doi = {10.1002/gps.2235},
   Key = {fds276990}
}

@article{fds276961,
   Author = {Norton, MC and Piercy, KW and Rabins, PV and Green, RC and Breitner,
             JCS and Ostbye, T and Corcoran, C and Welsh-Bohmer, KA and Lyketsos, CG and Tschanz, JT},
   Title = {Caregiver-recipient closeness and symptom progression in
             Alzheimer disease. The Cache County Dementia Progression
             Study.},
   Journal = {J Gerontol B Psychol Sci Soc Sci},
   Volume = {64},
   Number = {5},
   Pages = {560-568},
   Year = {2009},
   Month = {September},
   url = {http://www.ncbi.nlm.nih.gov/pubmed/19564210},
   Abstract = {Applying Rusbult's investment model of dyadic relationships,
             we examined the effect of caregiver-care recipient
             relationship closeness (RC) on cognitive and functional
             decline in Alzheimer's disease. After diagnosis, 167
             participants completed up to six visits, observed over an
             average of 20 months. Participants were 64% women, had a
             mean age of 86 years, and mean dementia duration of 4 years.
             Caregiver-rated closeness was measured using a six-item
             scale. In mixed models adjusted for dementia severity, dyads
             with higher levels of closeness (p < .05) and with spouse
             caregivers (p = .01) had slower cognitive decline. Effect of
             higher RC on functional decline was greater with spouse
             caregivers (p = .007). These findings of attenuated
             Alzheimer's dementia (AD) decline with closer relationships,
             particularly with spouse caregivers, are consistent with
             investment theory. Future interventions designed to enhance
             the caregiving dyadic relationship may help slow decline in
             AD.},
   Doi = {10.1093/geronb/gbp052},
   Key = {fds276961}
}

@article{fds277043,
   Author = {Levin, ED and Aschner, M and Heberlein, U and Ruden, D and Welsh-Bohmer,
             KA and Bartlett, S and Berger, K and Chen, L and Corl, AB and Eddins, D and French, R and Hayden, KM and Helmcke, K and Hirsch, HVB and Linney, E and Lnenicka, G and Page, GP and Possidente, D and Possidente, B and Kirshner, A},
   Title = {Genetic aspects of behavioral neurotoxicology.},
   Journal = {Neurotoxicology},
   Volume = {30},
   Number = {5},
   Pages = {741-753},
   Year = {2009},
   Month = {September},
   url = {http://www.ncbi.nlm.nih.gov/pubmed/19647018},
   Abstract = {Considerable progress has been made over the past couple of
             decades concerning the molecular bases of neurobehavioral
             function and dysfunction. The field of neurobehavioral
             genetics is becoming mature. Genetic factors contributing to
             neurologic diseases such as Alzheimer's disease have been
             found and evidence for genetic factors contributing to other
             diseases such as schizophrenia and autism are likely. This
             genetic approach can also benefit the field of behavioral
             neurotoxicology. It is clear that there is substantial
             heterogeneity of response with behavioral impairments
             resulting from neurotoxicants. Many factors contribute to
             differential sensitivity, but it is likely that genetic
             variability plays a prominent role. Important discoveries
             concerning genetics and behavioral neurotoxicity are being
             made on a broad front from work with invertebrate and
             piscine mutant models to classic mouse knockout models and
             human epidemiologic studies of polymorphisms. Discovering
             genetic factors of susceptibility to neurobehavioral
             toxicity not only helps identify those at special risk, it
             also advances our understanding of the mechanisms by which
             toxicants impair neurobehavioral function in the larger
             population. This symposium organized by Edward Levin and
             Annette Kirshner, brought together researchers from the
             laboratories of Michael Aschner, Douglas Ruden, Ulrike
             Heberlein, Edward Levin and Kathleen Welsh-Bohmer conducting
             studies with Caenorhabditis elegans, Drosophila, fish,
             rodents and humans studies to determine the role of genetic
             factors in susceptibility to behavioral impairment from
             neurotoxic exposure.},
   Doi = {10.1016/j.neuro.2009.07.014},
   Key = {fds277043}
}

@article{fds277088,
   Author = {Hulette, CM and Ervin, JF and Edmonds, Y and Antoine, S and Stewart, N and Szymanski, MH and Hayden, KM and Pieper, CF and Burke, JR and Welsh-Bohmer, KA},
   Title = {Cerebrovascular smooth muscle actin is increased in
             nondemented subjects with frequent senile plaques at
             autopsy: implications for the pathogenesis of Alzheimer
             disease.},
   Journal = {J Neuropathol Exp Neurol},
   Volume = {68},
   Number = {4},
   Pages = {417-424},
   Year = {2009},
   Month = {April},
   ISSN = {0022-3069},
   url = {http://www.ncbi.nlm.nih.gov/pubmed/19287310},
   Abstract = {We previously found that vascular smooth muscle actin (SMA)
             is reduced in the brains of patients with late stage
             Alzheimer disease (AD) compared with brains of nondemented,
             neuropathologically normal subjects. To assess the
             pathogenetic significance and disease specificity of this
             finding, we studied 3 additional patient groups: nondemented
             subjects without significant AD type pathology ("Normal"; n
             = 20), nondemented subjects with frequent senile plaques at
             autopsy ("Preclinical AD"; n = 20), and subjects with
             frontotemporal dementia ("FTD"; n = 10). The groups were
             matched for sex and age with those previously reported; SMA
             immunohistochemistry and image analysis were performed as
             previously described. Surprisingly, SMA expression in
             arachnoid, cerebral cortex, and white matter arterioles was
             greater in the Preclinical AD group than in the Normal and
             FTD groups. The plaques were not associated with amyloid
             angiopathy or other vascular disease in this group. Smooth
             muscle actin expression in the brains of the Normal group
             was intermediate between the Preclinical AD and FTD groups.
             All 3 groups exhibited much greater SMA expression than in
             our previous report. The presence of frequent plaques and
             increased arteriolar SMA expression in the brains of
             nondemented subjects suggest that increased SMA expression
             might represent a physiological response to
             neurodegeneration that could prevent or delay overt
             expression dementia in AD.},
   Doi = {10.1097/NEN.0b013e31819e6334},
   Key = {fds277088}
}

@article{fds371175,
   Author = {Norton, MC and Tschanz, JT and Steffens, DC and Skoog, I and Welsh-Bohmer, KA and Munger, R and Breitner, JCS},
   Title = {Age Moderates the Effect of Late-life Depression on Incident
             Alzheimer's Disease Risk},
   Journal = {AMERICAN JOURNAL OF GERIATRIC PSYCHIATRY},
   Volume = {17},
   Number = {3},
   Pages = {A72-A72},
   Publisher = {LIPPINCOTT WILLIAMS & WILKINS},
   Year = {2009},
   Month = {March},
   Key = {fds371175}
}

@article{fds276989,
   Author = {Potter, GG and McQuoid, DR and Steffens, DC and Welsh-Bohmer, KA and Krishnan, KRR},
   Title = {Neuropsychological correlates of magnetic resonance
             imaging-defined subcortical ischemic depression.},
   Journal = {Int J Geriatr Psychiatry},
   Volume = {24},
   Number = {3},
   Pages = {219-225},
   Year = {2009},
   Month = {March},
   url = {http://www.ncbi.nlm.nih.gov/pubmed/18655212},
   Abstract = {OBJECTIVE: The goal of the current study was to examine the
             neuropsychological profile of magnetic resonance imaging
             (MRI)-defined subcortical ischemic depression (SID).
             METHODS: Clinically depressed older adults with MRI-defined
             SID (n = 70) and depressed elders without SID (n = 75) were
             compared on neuropsychological performance, depression
             symptoms, and medical burden. RESULTS: Group comparisons
             revealed that the SID was associated with worse performance
             on all neuropsychological measures, but also with greater
             age, higher cardiac illness burden, and greater deficits in
             the depression symptoms of self-initiation and
             concentration. In multivariate regression models, auditory
             working memory and nonverbal memory remained worse among the
             SID group after controlling for contributions of age,
             cardiovascular risk, and depression symptoms. CONCLUSIONS:
             Although auditory working memory span and nonverbal memory
             appear to be specifically associated with the ischemic
             pathology that defines SID, the typical individual with SID
             is also likely to have a broader profile of
             neuropsychological deficits than those without SID because
             they are typically older and have specific depression
             symptoms that predispose them to compromised neurocognitive
             performance.},
   Doi = {10.1002/gps.2093},
   Key = {fds276989}
}

@article{fds276962,
   Author = {Welsh-Bohmer, KA and White, CL},
   Title = {Alzheimer disease: what changes in the brain cause
             dementia?},
   Journal = {Neurology},
   Volume = {72},
   Number = {4},
   Pages = {e21},
   Year = {2009},
   Month = {January},
   url = {http://www.ncbi.nlm.nih.gov/pubmed/19171824},
   Doi = {10.1212/01.wnl.0000343818.11392.d9},
   Key = {fds276962}
}

@article{fds277042,
   Author = {Welsh-Bohmer, KA and Ostbye, T and Sanders, L and Pieper, CF and Hayden,
             KM and Tschanz, JT and Norton, MC and Cache Country Study
             Group},
   Title = {Neuropsychological performance in advanced age: influences
             of demographic factors and Apolipoprotein E: findings from
             the Cache County Memory Study.},
   Journal = {Clin Neuropsychol},
   Volume = {23},
   Number = {1},
   Pages = {77-99},
   Year = {2009},
   Month = {January},
   url = {http://www.ncbi.nlm.nih.gov/pubmed/18609337},
   Abstract = {The Cache County Study of Memory in Aging (CCMS) is an
             epidemiological study of Alzheimer's disease (AD), mild
             cognitive disorders, and aging in a population of
             exceptionally long-lived individuals (7th to 11th decade).
             Observation of population members without dementia provides
             an opportunity for establishing the range of normal
             neurocognitive performance in a representative sample of the
             very old. We examined neurocognitive performance of the
             normal participants undergoing full clinical evaluations (n
             = 507) and we tested the potential modifying effects of
             apolipoprotein E (APOE) genotype, a known genetic risk
             factor for the later development of AD. The results indicate
             that advanced age and low education are related to lower
             test scores across nearly all of the neurocognitive
             measures. Gender and APOE epsilon4 both had negligible and
             inconsistent influences, affecting only isolated measures of
             memory and expressive speech (in case of gender). The gender
             and APOE effects disappeared once age and education were
             controlled. The study of this exceptionally long-lived
             population provides useful normative information regarding
             the broad range of "normal" cognition seen in advanced age.
             Among elderly without dementia or other cognitive
             impairment, APOE does not appear to exert any major effects
             on cognition once other demographic influences are
             controlled.},
   Doi = {10.1080/13854040801894730},
   Key = {fds277042}
}

@article{fds371176,
   Author = {Peters, ME and Rosenberg, P and Steinberg, M and Tschanz, J and Welsh-Bohmer, K and Hayden, K and Breitner, J and Lyketsos,
             CG},
   Title = {Prevalence of Neuropsychiatric Symptoms in CIND and its
             Subtypes: Cache County Study},
   Journal = {AMERICAN JOURNAL OF GERIATRIC PSYCHIATRY},
   Volume = {17},
   Number = {3},
   Pages = {A117-A118},
   Year = {2009},
   Key = {fds371176}
}

@article{fds277072,
   Author = {Heinzen, EL and Ge, D and Cronin, KD and Maia, JM and Shianna, KV and Gabriel, WN and Welsh-Bohmer, KA and Hulette, CM and Denny, TN and Goldstein, DB},
   Title = {Tissue-specific genetic control of splicing: implications
             for the study of complex traits.},
   Journal = {PLoS Biol},
   Volume = {6},
   Number = {12},
   Pages = {e1},
   Year = {2008},
   Month = {December},
   url = {http://www.ncbi.nlm.nih.gov/pubmed/19222302},
   Abstract = {Numerous genome-wide screens for polymorphisms that
             influence gene expression have provided key insights into
             the genetic control of transcription. Despite this work, the
             relevance of specific polymorphisms to in vivo expression
             and splicing remains unclear. We carried out the first
             genome-wide screen, to our knowledge, for SNPs that
             associate with alternative splicing and gene expression in
             human primary cells, evaluating 93 autopsy-collected
             cortical brain tissue samples with no defined
             neuropsychiatric condition and 80 peripheral blood
             mononucleated cell samples collected from living healthy
             donors. We identified 23 high confidence associations with
             total expression and 80 with alternative splicing as
             reflected by expression levels of specific exons. Fewer than
             50% of the implicated SNPs however show effects in both
             tissue types, reflecting strong evidence for distinct
             genetic control of splicing and expression in the two tissue
             types. The data generated here also suggest the possibility
             that splicing effects may be responsible for up to 13 out of
             84 reported genome-wide significant associations with human
             traits. These results emphasize the importance of
             establishing a database of polymorphisms affecting splicing
             and expression in primary tissue types and suggest that
             splicing effects may be of more phenotypic significance than
             overall gene expression changes.},
   Doi = {10.1371/journal.pbio.1000001},
   Key = {fds277072}
}

@article{fds277041,
   Author = {Rosenberg, PB and Mielke, MM and Tschanz, J and Cook, L and Corcoran, C and Hayden, KM and Norton, M and Rabins, PV and Green, RC and Welsh-Bohmer,
             KA and Breitner, JCS and Munger, R and Lyketsos, CG},
   Title = {Effects of cardiovascular medications on rate of functional
             decline in Alzheimer disease.},
   Journal = {Am J Geriatr Psychiatry},
   Volume = {16},
   Number = {11},
   Pages = {883-892},
   Year = {2008},
   Month = {November},
   url = {http://www.ncbi.nlm.nih.gov/pubmed/18978249},
   Abstract = {BACKGROUND: Evidence suggests that cardiovascular
             medications, including statins and antihypertensive
             medications, may delay cognitive decline in patients with
             Alzheimer dementia (AD). We examined the association of
             cardiovascular medication use and rate of functional decline
             in a population-based cohort of individuals with incident
             AD. METHODS: In the Dementia Progression Study of the Cache
             County Study on Memory, Health, and Aging, 216 individuals
             with incident AD were identified and followed longitudinally
             with in-home visits for a mean of 3.0 years and 2.1
             follow-up visits. The Clinical Dementia Rating (CDR) was
             completed at each follow-up. Medication use was inventoried
             during in-home visits. Generalized least-squares
             random-effects regression was performed with CDR Sum of
             Boxes (CDR-Sum) as the outcome and cardiovascular medication
             use as the major predictors. RESULTS: CDR-Sum increased an
             average of 1.69 points annually, indicating a steady decline
             in functioning. After adjustment for demographic variables
             and the baseline presence of cardiovascular conditions, use
             of statins (p = 0.03) and beta-blockers (p = 0.04) was
             associated with a slower annual rate of increase in CDR-Sum
             (slower rate of functional decline) of 0.75 and 0.68 points
             respectively, while diuretic use was associated with a
             faster rate of increase in CDR-Sum (p = 0.01; 0.96 points
             annually). Use of calcium-channel blockers,
             angiotensin-converting enzyme inhibitors, digoxin, or
             nitrates did not affect the rate of functional decline.
             CONCLUSIONS: In this population-based study of individuals
             with incident AD, use of statins and beta-blockers was
             associated with delay of functional decline. Further studies
             are needed to confirm these results and to determine whether
             treatment with these medications may help delay AD
             progression.},
   Doi = {10.1097/JGP.0b013e318181276a},
   Key = {fds277041}
}

@article{fds276960,
   Author = {Welsh-Bohmer, K},
   Title = {Neuropsychological characterization of dementia
             patients},
   Journal = {Primary Psychiatry},
   Volume = {15},
   Number = {10 SUPPL. 6},
   Pages = {10-13},
   Year = {2008},
   Month = {October},
   ISSN = {1082-6319},
   Key = {fds276960}
}

@article{fds277117,
   Author = {Story, TJ and Potter, GG and Attix, DK and Welsh-Bohmer, KA and Steffens, DC},
   Title = {Neurocognitive correlates of response to treatment in
             late-life depression.},
   Journal = {Am J Geriatr Psychiatry},
   Volume = {16},
   Number = {9},
   Pages = {752-759},
   Year = {2008},
   Month = {September},
   url = {http://www.ncbi.nlm.nih.gov/pubmed/18697883},
   Abstract = {UNLABELLED: Depression is often associated with
             neurocognitive deficits in older adults, particularly in the
             domains of information processing speed, episodic memory,
             and executive functions. Greater neurocognitive dysfunction
             while depressed is associated with a less effective
             treatment response; however, questions remain about the
             specific variables that characterize patients showing low
             treatment response and persistent cognitive deficiencies.
             OBJECTIVES: The authors examined neurocognitive variables
             that differentiated patients who showed robust versus weak
             responses to antidepressant therapy. PARTICIPANTS: The
             baseline sample included 110 women and 67 men, with a mean
             age of 69.1 years (SD = 6.9) and mean education of 14 years
             (SD = 3.3). DESIGN: Patients enrolled in a treatment study
             completed both a structured diagnostic assessment for
             depression and neuropsychological testing at study entry and
             1-year follow-up. MEASUREMENTS: Clinicians rated patient
             depression using the Montgomery-Asberg Depression Rating
             Scale. Neuropsychological assessments consisted of prose
             recall and percent retention (Wechsler Memory Scale -III
             Logical Memory), word-list recall, attention and visuomotor
             processing speed (Trail Making A, Symbol Digit Modalities
             Test), and mental flexibility (Trail Making B).
             INTERVENTIONS: Patients underwent treatment for depression
             following the guidelines of the Duke Somatic Treatment
             Algorithm for Geriatric Depression approach. RESULTS:
             Individuals who demonstrated the greatest improvement in
             mood symptoms at follow-up exhibited better prose recall and
             faster processing speed at baseline than individuals who
             demonstrated weaker treatment responses. These differences
             remained after controlling for depression severity at both
             time-points. CONCLUSION: The current results suggest that
             better pretreatment cognitive function, particularly in
             verbal memory, is associated with a greater treatment
             response in late-life depression.},
   Doi = {10.1097/JGP.0b013e31817e739a},
   Key = {fds277117}
}

@article{fds277024,
   Author = {Carney, RM and Slifer, MA and Lin, PI and Gaskell, PC and Scott, WK and Potocky, CF and Hulette, CM and Welsh-Bohmer, KA and Schmechel, DE and Vance, JM and Pericak-Vance, MA},
   Title = {Longitudinal follow-up of late-onset Alzheimer disease
             families.},
   Journal = {Am J Med Genet B Neuropsychiatr Genet},
   Volume = {147B},
   Number = {5},
   Pages = {571-578},
   Year = {2008},
   Month = {July},
   url = {http://www.ncbi.nlm.nih.gov/pubmed/18361431},
   Abstract = {Historically, data for genetic studies are collected at one
             time point. However, for diseases with late onset or with
             complex phenotypes, such as Alzheimer disease (AD),
             restricting diagnosis to a single ascertainment contact may
             not be sufficient. Affection status may change over time and
             some initial diagnoses may be inconclusive. Follow-up
             provides the opportunity to resolve these complications.
             However, to date, previous studies have not formally
             demonstrated that longitudinally re-contacting families is
             practical or productive. To update data initially collected
             for linkage analysis of late-onset Alzheimer disease (LOAD),
             we successfully re-contacted 63 of 81 (78%) multiplex
             families (two to 17 years after ascertainment). Clinical
             status changed for 73 of the 230 (32%) non-affected
             participants. Additionally, expanded family history
             identified 20 additional affected individuals to supplement
             the data set. Furthermore, fostering ongoing relationships
             with participating families helped recruit 101 affected
             participants into an autopsy and tissue donation program.
             Despite similar presentations, discordance between clinical
             diagnosis and neuropathologic diagnosis was observed in 28%
             of those with tissue diagnoses. Most of the families were
             successfully re-contacted, and significant refinement and
             supplementation of the data was achieved. We concluded that
             serial contact with longitudinal evaluation of families has
             significant implications for genetic analyses.},
   Doi = {10.1002/ajmg.b.30590},
   Key = {fds277024}
}

@article{fds276957,
   Author = {Szekely, CA and Green, RC and Breitner, JCS and Østbye, T and Beiser,
             AS and Corrada, MM and Dodge, HH and Ganguli, M and Kawas, CH and Kuller,
             LH and Psaty, BM and Resnick, SM and Wolf, PA and Zonderman, AB and Welsh-Bohmer, KA and Zandi, PP},
   Title = {No advantage of A beta 42-lowering NSAIDs for prevention of
             Alzheimer dementia in six pooled cohort studies.},
   Journal = {Neurology},
   Volume = {70},
   Number = {24},
   Pages = {2291-2298},
   Year = {2008},
   Month = {June},
   url = {http://www.ncbi.nlm.nih.gov/pubmed/18509093},
   Abstract = {INTRODUCTION: Observational studies show reduced incidence
             of Alzheimer dementia (AD) in users of nonsteroidal
             anti-inflammatory drugs (NSAIDs). One hypothesis holds that
             the subset of NSAIDs known as selective A beta(42)-lowering
             agents (SALAs) is responsible for this apparent reduction in
             AD risk. METHODS: We pooled individual-level data from six
             prospective studies to obtain a sufficient sample to examine
             AD risk in users of SALA vs non-SALA NSAIDs. RESULTS: Of
             13,499 initially dementia-free participants (70,863
             person-years), 820 developed incident AD. Users of NSAIDs
             (29.6%) showed reduced risk of AD (adjusted hazard ratio
             [aHR] 0.77, 95% CI 0.65-0.91). The point estimates were
             similar for SALAs (aHR 0.87, CI 0.72-1.04) and non-SALAs
             (aHR 0.75, CI 0.56-1.01). Because 573 NSAID users (14.5%)
             reported taking both a SALA and non-SALA, we examined their
             use alone and in combination. Resulting aHRs were 0.82 (CI
             0.67-0.99) for SALA only, 0.60 (CI 0.40-0.90) for non-SALA
             only, and 0.87 (CI 0.57-1.33) for both NSAIDs (Wald test for
             differences, p = 0.32). The 40.7% of participants who used
             aspirin also showed reduced risk of AD, even when they used
             no other NSAIDs (aHR 0.78, CI 0.66-0.92). By contrast, there
             was no association with use of acetaminophen (aHR 0.93, CI
             0.76-1.13). CONCLUSIONS: In this pooled dataset,
             nonsteroidal anti-inflammatory drug (NSAID) use reduced the
             risk of Alzheimer dementia (AD). However, there was no
             apparent advantage in AD risk reduction for the subset of
             NSAIDs shown to selectively lower A beta(42), suggesting
             that all conventional NSAIDs including aspirin have a
             similar protective effect in humans.},
   Doi = {10.1212/01.wnl.0000313933.17796.f6},
   Key = {fds276957}
}

@article{fds276958,
   Author = {Treiber, KA and Lyketsos, CG and Corcoran, C and Steinberg, M and Norton, M and Green, RC and Rabins, P and Stein, DM and Welsh-Bohmer,
             KA and Breitner, JCS and Tschanz, JT},
   Title = {Vascular factors and risk for neuropsychiatric symptoms in
             Alzheimer's disease: the Cache County Study.},
   Journal = {Int Psychogeriatr},
   Volume = {20},
   Number = {3},
   Pages = {538-553},
   Year = {2008},
   Month = {June},
   ISSN = {1041-6102},
   url = {http://www.ncbi.nlm.nih.gov/pubmed/18289451},
   Abstract = {OBJECTIVE: To examine, in an exploratory analysis, the
             association between vascular conditions and the occurrence
             of neuropsychiatric symptoms (NPS) in a population-based
             sample of incident Alzheimer's disease (AD). METHODS: The
             sample consisted of 254 participants, identified through two
             waves of assessment. NPS were assessed using the
             Neuropsychiatric Inventory. Prior to the onset of AD, data
             regarding a history of stroke, hypertension, hyperlipidemia,
             heart attack or coronary artery bypass graft (CABG), and
             diabetes were recorded. Logistic regression procedures were
             used to examine the relationship of each vascular condition
             to individual neuropsychiatric symptoms. Covariates
             considered were age, gender, education, APOE genotype,
             dementia severity, and overall health status. RESULTS: One
             or more NPS were observed in 51% of participants. Depression
             was most common (25.8%), followed by apathy (18.6%), and
             irritability (17.7%). Least common were elation (0.8%),
             hallucinations (5.6%), and disinhibition (6.0%). Stroke
             prior to the onset of AD was associated with increased risk
             of delusions (OR = 4.76, p = 0.02), depression (OR = 3.87, p
             = 0.03), and apathy (OR = 4.48, p = 0.02). Hypertension was
             associated with increased risk of delusions (OR = 2.34, p =
             0.02), anxiety (OR = 4.10, p = 0.002), and
             agitation/aggression (OR = 2.82, p = 0.01). No associations
             were observed between NPS and diabetes, hyperlipidemia,
             heart attack or CABG, or overall health. CONCLUSIONS:
             Results suggest that a history of stroke and hypertension
             increase the risk of specific NPS in patients with AD. These
             conditions may disrupt neural circuitry in brain areas
             involved in NPS. Findings may provide an avenue for
             reduction in occurrence of NPS through the treatment or
             prevention of vascular risk conditions.},
   Doi = {10.1017/S1041610208006704},
   Key = {fds276958}
}

@article{fds277100,
   Author = {Browndyke, JN and Paskavitz, J and Sweet, LH and Cohen, RA and Tucker,
             KA and Welsh-Bohmer, KA and Burke, JR and Schmechel,
             DE},
   Title = {Neuroanatomical correlates of malingered memory impairment:
             event-related fMRI of deception on a recognition memory
             task.},
   Journal = {Brain Inj},
   Volume = {22},
   Number = {6},
   Pages = {481-489},
   Year = {2008},
   Month = {June},
   url = {http://www.ncbi.nlm.nih.gov/pubmed/18465389},
   Abstract = {PRIMARY OBJECTIVE: Event-related, functional magnetic
             resonance imaging (fMRI) data were acquired in healthy
             participants during purposefully malingered and normal
             recognition memory performances to evaluate the neural
             substrates of feigned memory impairment. METHODS AND
             PROCEDURES: Pairwise, between-condition contrasts of neural
             activity associated with discrete recognition memory
             responses were conducted to isolate dissociable neural
             activity between normal and malingered responding while
             simultaneously controlling for shared stimulus familiarity
             and novelty effects. Response timing characteristics were
             also examined for any association with observed
             between-condition activity differences. OUTCOMES AND
             RESULTS: Malingered recognition memory errors, regardless of
             type, were associated with inferior parietal and superior
             temporal activity relative to normal performance, while
             feigned recognition target misses produced additional
             dorsomedial frontal activation and feigned foil false alarms
             activated bilateral ventrolateral frontal regions.
             Malingered response times were associated with activity in
             the dorsomedial frontal, temporal and inferior parietal
             regions. Normal memory responses were associated with
             greater inferior occipitotemporal and dorsomedial parietal
             activity, suggesting greater reliance upon
             visual/attentional networks for proper task performance.
             CONCLUSIONS: The neural substrates subserving feigned
             recognition memory deficits are influenced by response
             demand and error type, producing differential activation of
             cortical regions important to complex visual processing,
             executive control, response planning and working memory
             processes.},
   Doi = {10.1080/02699050802084894},
   Key = {fds277100}
}

@article{fds276956,
   Author = {Hallam, BJ and Brown, WS and Ross, C and Buckwalter, JG and Bigler, ED and Tschanz, JT and Norton, MC and Welsh-Bohmer, KA and Breitner,
             JCS},
   Title = {Regional atrophy of the corpus callosum in
             dementia.},
   Journal = {J Int Neuropsychol Soc},
   Volume = {14},
   Number = {3},
   Pages = {414-423},
   Year = {2008},
   Month = {May},
   url = {http://www.ncbi.nlm.nih.gov/pubmed/18419840},
   Abstract = {The regional distribution of degeneration of the corpus
             callosum (CC) in dementia is not yet clear. This study
             compared regional CC size in participants (n = 179) from the
             Cache County Memory and Aging Study. Participants
             represented a range of cognitive function: Alzheimer's
             disease (AD), vascular dementia (VaD), mild ambiguous
             (MA-cognitive problems, but not severe enough for diagnosis
             of dementia), and healthy older adults. CC outlines obtained
             from midsagittal magnetic resonance images were divided into
             99 equally spaced widths. Factor analysis of these callosal
             widths identified 10 callosal regions. Multivariate analysis
             of variance revealed significant group differences for
             anterior and posterior callosal regions. Post-hoc pairwise
             comparisons of CC regions in patient groups as compared to
             the control group (controlling for age) revealed trends
             toward smaller anterior and posterior regions, but not all
             were statistically significant. As compared to controls,
             significantly smaller anterior and posterior CC regions were
             found in the AD group; significantly smaller anterior CC
             regions in the VaD group; but no significant CC regional
             differences in the MA group. Findings suggest that
             dementia-related CC atrophy occurs primarily in the anterior
             and posterior portions.},
   Doi = {10.1017/S1355617708080533},
   Key = {fds276956}
}

@article{fds276980,
   Author = {Norton, MC and Singh, A and Skoog, I and Corcoran, C and Tschanz, JT and Zandi, PP and Breitner, JCS and Welsh-Bohmer, KA and Steffens, DC and Cache County Investigators},
   Title = {Church attendance and new episodes of major depression in a
             community study of older adults: the Cache County
             Study.},
   Journal = {J Gerontol B Psychol Sci Soc Sci},
   Volume = {63},
   Number = {3},
   Pages = {P129-P137},
   Year = {2008},
   Month = {May},
   ISSN = {1079-5014},
   url = {http://www.ncbi.nlm.nih.gov/pubmed/18559677},
   Abstract = {We examined the relation between church attendance,
             membership in the Church of Jesus Christ of Latter-Day
             Saints (LDS), and major depressive episode, in a
             population-based study of aging and dementia in Cache
             County, Utah. Participants included 2,989 nondemented
             individuals aged between 65 and 100 years who were
             interviewed initially in 1995 to 1996 and again in 1998 to
             1999. LDS church members reported twice the rate of major
             depression that non-LDS members did (odds ratio = 2.56, 95%
             confidence interval = 1.07-6.08). Individuals attending
             church weekly or more often had a significantly lower risk
             for major depression. After controlling for demographic and
             health variables and the strongest predictor of future
             episodes of depression, a prior depression history, we found
             that church attendance more often than weekly remained a
             significant protectant (odds ratio = 0.51, 95% confidence
             interval = 0.28-0.92). Results suggest that there may be a
             threshold of church attendance that is necessary for a
             person to garner long-term protection from depression. We
             discuss sociological factors relevant to LDS
             culture.},
   Doi = {10.1093/geronb/63.3.p129},
   Key = {fds276980}
}

@article{fds277040,
   Author = {Fotuhi, M and Zandi, PP and Hayden, KM and Khachaturian, AS and Szekely,
             CA and Wengreen, H and Munger, RG and Norton, MC and Tschanz, JT and Lyketsos, CG and Breitner, JCS and Welsh-Bohmer,
             K},
   Title = {Better cognitive performance in elderly taking antioxidant
             vitamins E and C supplements in combination with
             nonsteroidal anti-inflammatory drugs: the Cache County
             Study.},
   Journal = {Alzheimers Dement},
   Volume = {4},
   Number = {3},
   Pages = {223-227},
   Year = {2008},
   Month = {May},
   url = {http://www.ncbi.nlm.nih.gov/pubmed/18631971},
   Abstract = {Studies have shown less cognitive decline and lower risk of
             Alzheimer's disease in elderly individuals consuming either
             antioxidant vitamins or nonsteroidal anti-inflammatory drugs
             (NSAIDs). The potential of added benefit from their combined
             use has not been studied. We therefore analyzed data from
             3,376 elderly participants of the Cache County Study who
             were given the Modified Mini-Mental State examination up to
             three times during a period of 8 years. Those who used a
             combination of vitamins E and C supplements and NSAIDs at
             baseline declined by an average 0.96 fewer points every 3
             years than nonusers (P < .05). This apparent effect was
             attributable entirely to participants with the APOE epsilon4
             allele, whose users declined by 2.25 fewer points than
             nonusers every 3 years (P < .05). These results suggest that
             among elderly individuals with an APOE epsilon4 allele,
             there is an association between using antioxidant
             supplements in combination with NSAIDs and less cognitive
             decline over time.},
   Doi = {10.1016/j.jalz.2008.01.004},
   Key = {fds277040}
}

@article{fds276954,
   Author = {Martini, B and Buffington, ALH and Welsh-Bohmer, KA and Brandt, J and ADAPT Research Group},
   Title = {Time of day affects episodic memory in older
             adults.},
   Journal = {Neuropsychol Dev Cogn B Aging Neuropsychol
             Cogn},
   Volume = {15},
   Number = {2},
   Pages = {146-164},
   Year = {2008},
   Month = {March},
   url = {http://www.ncbi.nlm.nih.gov/pubmed/17851985},
   Abstract = {The neuropsychological test scores of 2030 cognitively
             normal older adults were examined to evaluate performance
             patterns as they related to time of day (TOD) at which
             testing was initiated. Multiple regression analyses were
             used to examine the association of TOD with scores on seven
             neuropsychological tests used in the clinical evaluation of
             dementia. Episodic memory performance was significantly
             related to TOD, while memory span and verbal fluency were
             not. Best performance occurred during early morning hours
             and late afternoon; worst performance occurred mid-day
             (i.e., noon). These findings may have implications for
             clinical assessment, the design of research on dementia, and
             the daily functioning of older adults.},
   Doi = {10.1080/13825580601186643},
   Key = {fds276954}
}

@article{fds276955,
   Author = {Welsh-Bohmer, KA},
   Title = {Defining "prodromal" Alzheimer's disease, frontotemporal
             dementia, and Lewy body dementia: are we there
             yet?},
   Journal = {Neuropsychol Rev},
   Volume = {18},
   Number = {1},
   Pages = {70-72},
   Year = {2008},
   Month = {March},
   ISSN = {1040-7308},
   url = {http://www.ncbi.nlm.nih.gov/pubmed/18327642},
   Doi = {10.1007/s11065-008-9057-y},
   Key = {fds276955}
}

@article{fds277062,
   Author = {Fillenbaum, GG and van Belle, G and Morris, JC and Mohs, RC and Mirra,
             SS and Davis, PC and Tariot, PN and Silverman, JM and Clark, CM and Welsh-Bohmer, KA and Heyman, A},
   Title = {Consortium to Establish a Registry for Alzheimer's Disease
             (CERAD): the first twenty years.},
   Journal = {Alzheimers Dement},
   Volume = {4},
   Number = {2},
   Pages = {96-109},
   Year = {2008},
   Month = {March},
   url = {http://www.ncbi.nlm.nih.gov/pubmed/18631955},
   Abstract = {The Consortium to Establish a Registry for Alzheimer's
             Disease (CERAD) was funded by the National Institute on
             Aging in 1986 to develop standardized, validated measures
             for the assessment of Alzheimer's disease (AD). The present
             report describes the measures that CERAD developed during
             its first decade and their continued use in their original
             and translated forms. These measures include clinical,
             neuropsychological, neuropathologic, and behavioral
             assessments of AD and also assessment of family history and
             parkinsonism in AD. An approach to evaluating neuroimages
             did not meet the standards desired. Further evaluations that
             could not be completed because of lack of funding (but where
             some materials are available) include evaluation of very
             severe AD and of service use and need by patient and
             caregiver. The information that was developed in the U.S.
             and abroad permits standardized assessment of AD in clinical
             practice, facilitates epidemiologic studies, and provides
             information valuable for individual and public health
             planning. CERAD materials and data remain available for
             those wishing to use them.},
   Doi = {10.1016/j.jalz.2007.08.005},
   Key = {fds277062}
}

@article{fds277087,
   Author = {Steinberg, M and Shao, H and Zandi, P and Lyketsos, CG and Welsh-Bohmer,
             KA and Norton, MC and Breitner, JCS and Steffens, DC and Tschanz, JT and Cache County Investigators},
   Title = {Point and 5-year period prevalence of neuropsychiatric
             symptoms in dementia: the Cache County Study.},
   Journal = {Int J Geriatr Psychiatry},
   Volume = {23},
   Number = {2},
   Pages = {170-177},
   Year = {2008},
   Month = {February},
   ISSN = {0885-6230},
   url = {http://www.ncbi.nlm.nih.gov/pubmed/17607801},
   Abstract = {BACKGROUND: Neuropsychiatric symptoms are nearly universal
             in dementia, yet little is known about their longitudinal
             course in the community. OBJECTIVE: To estimate point and
             5-year period prevalence of neuropsychiatric symptoms in an
             incident sample of 408 dementia participants from the Cache
             County Study. METHODS: The Neuropsychiatric Inventory
             assessed symptoms at baseline and at 1.5 years, 3.0 years,
             4.1 years, and 5.3 years. Point prevalence, period
             prevalence and mean symptom severity at each time point were
             estimated. RESULTS: Point prevalence for delusions was 18%
             at baseline and 34-38% during the last three visits;
             hallucinations, 10% at baseline and 19-24% subsequently;
             agitation/aggression fluctuated between 13% and 24%;
             depression 29% at baseline and 41-47% subsequently; apathy
             increased from 20% at baseline to 51% at 5.3 years; elation
             never rose above 1%; anxiety 14% at baseline and 24-32%
             subsequently; disinhibition fluctuated between 2% and 15%;
             irritability between 17% and 27%; aberrant motor behavior
             gradually increased from 7% at baseline to 29% at 5.3 years.
             Point prevalence for any symptom was 56% at baseline and
             76-87% subsequently. Five-year period prevalence was
             greatest for depression (77%), apathy (71%), and anxiety
             (62%); lowest for elation (6%), and disinhibition (31%).
             Ninety-seven percent experienced at least one symptom.
             Symptom severity was consistently highest for apathy.
             CONCLUSIONS: Participants were most likely to develop
             depression, apathy, or anxiety, and least likely to develop
             elation or disinhibition. Give converging evidence that
             syndromal definitions may more accurately capture
             neuropsychiatric co-morbidity in dementia, future efforts to
             validate such syndromes are warranted.},
   Doi = {10.1002/gps.1858},
   Key = {fds277087}
}

@article{fds276959,
   Author = {Welsh-Bohmer, K},
   Title = {Neuropsychological characterization of dementia
             patients},
   Journal = {CNS Spectrums},
   Volume = {13},
   Number = {10 SUPPL. 16},
   Pages = {10-13},
   Year = {2008},
   ISSN = {1092-8529},
   url = {http://dx.doi.org/10.1017/s1092852900026973},
   Doi = {10.1017/s1092852900026973},
   Key = {fds276959}
}

@article{fds277071,
   Author = {Heinzen, EL and Ge, D and Cronin, KD and Maia, JM and Shianna, KV and Gabriel, WN and Welsh-Bohmer, KA and Hulette, CM and Denny, TN and Goldstein, DB},
   Title = {Tissue-specific genetic control of splicing: Implications
             for the study of complex traits},
   Journal = {PLoS Biology},
   Volume = {6},
   Number = {12},
   Pages = {2869-2879},
   Year = {2008},
   ISSN = {1544-9173},
   url = {http://dx.doi.org/10.1371/journal.pbio.1000001},
   Abstract = {Numerous genome-wide screens for polymorphisms that
             influence gene expression have provided key insights into
             the genetic control of transcription. Despite this work, the
             relevance of specific polymorphisms to in vivo expression
             and splicing remains unclear. We carried out the first
             genome-wide screen, to our knowledge, for SNPs that
             associate with alternative splicing and gene expression in
             human primary cells, evaluating 93 autopsy-collected
             cortical brain tissue samples with no defined
             neuropsychiatric condition and 80 peripheral blood
             mononucleated cell samples collected from living healthy
             donors. We identified 23 high confidence associations with
             total expression and 80 with alternative splicing as
             reflected by expression levels of specific exons. Fewer than
             50% of the implicated SNPs however show effects in both
             tissue types, reflecting strong evidence for distinct
             genetic control of splicing and expression in the two tissue
             types. The data generated here also suggest the possibility
             that splicing effects may be responsible for up to 13 out of
             84 reported genome-wide significant associations with human
             traits. These results emphasize the importance of
             establishing a database of polymorphisms affecting splicing
             and expression in primary tissue types and suggest that
             splicing effects may be of more phenotypic significance than
             overall gene expression changes. © 2008 Heinzen et
             al.},
   Doi = {10.1371/journal.pbio.1000001},
   Key = {fds277071}
}

@article{fds277086,
   Author = {Ervin, JF and Heinzen, EL and Cronin, KD and Goldstein, D and Szymanski,
             MH and Burke, JR and Welsh-Bohmer, KA and Hulette,
             CM},
   Title = {Postmortem delay has minimal effect on brain RNA
             integrity.},
   Journal = {J Neuropathol Exp Neurol},
   Volume = {66},
   Number = {12},
   Pages = {1093-1099},
   Year = {2007},
   Month = {December},
   ISSN = {0022-3069},
   url = {http://www.ncbi.nlm.nih.gov/pubmed/18090918},
   Abstract = {The Bryan Alzheimer Disease Research Center obtains
             postmortem human brain tissue from patients with Alzheimer
             disease (AD) and cognitively normal control subjects for
             molecular and genetic research programs. A growing body of
             research suggests that variations in gene transcript levels
             may contribute to the onset and progression of disease.
             Identifying how the regulation of gene expression may affect
             AD requires the use of high-quality mRNA from banked human
             brains. The present study was conducted to establish the
             quality and suitability of available banked brain tissue for
             future gene expression studies. We chose 32 AD cases with
             Braak stage IV, V, or VI. These AD cases were matched to 36
             normal control cases by age and sex when possible. Multiple
             regions from each brain were sampled, including frontal
             cortex, temporal cortex, occipital cortex, and cerebellum.
             Hippocampus was also available for study from 14 control
             cases. A comparison of several antemortem and postmortem
             variables, such as postmortem interval, agonal state,
             ventricular cerebrospinal fluid pH, and cause of death were
             analyzed. RNA was isolated from at least 1 area from every
             brain and most brains yielded intact RNA from all regions
             tested. Analysis of the clinical variables did not reveal
             any features that correlated with the ability to recover
             intact mRNA. We conclude that undegraded mRNA may be
             isolated from most brain regions many hours postmortem and
             that neither the pH of ventricular fluid nor postmortem
             interval is predictive of mRNA integrity.},
   Doi = {10.1097/nen.0b013e31815c196a},
   Key = {fds277086}
}

@article{fds277039,
   Author = {Mielke, MM and Rosenberg, PB and Tschanz, J and Cook, L and Corcoran, C and Hayden, KM and Norton, M and Rabins, PV and Green, RC and Welsh-Bohmer,
             KA and Breitner, JCS and Munger, R and Lyketsos, CG},
   Title = {Vascular factors predict rate of progression in Alzheimer
             disease.},
   Journal = {Neurology},
   Volume = {69},
   Number = {19},
   Pages = {1850-1858},
   Year = {2007},
   Month = {November},
   url = {http://www.ncbi.nlm.nih.gov/pubmed/17984453},
   Abstract = {BACKGROUND: While there is considerable epidemiologic
             evidence that cardiovascular risk factors increase risk of
             incident Alzheimer disease (AD), few studies have examined
             their effect on progression after an established AD
             diagnosis. OBJECTIVE: To examine the effect of vascular
             factors, and potential age modification, on rate of
             progression in a longitudinal study of incident dementia.
             METHODS: A total of 135 individuals with incident AD,
             identified in a population-based sample of elderly persons
             in Cache County, UT, were followed with in-home visits for a
             mean of 3.0 years (range: 0.8 to 9.5) and 2.1 follow-up
             visits (range: 1 to 5). The Clinical Dementia Rating (CDR)
             Scale and Mini-Mental State Examination (MMSE) were
             administered at each visit. Baseline vascular factors were
             determined by interview and physical examination.
             Generalized least-squares random-effects regression was
             performed with CDR Sum of Boxes (CDR-Sum) or MMSE as the
             outcome, and vascular index or individual vascular factors
             as independent variables. RESULTS: Atrial fibrillation,
             systolic hypertension, and angina were associated with more
             rapid decline on both the CDR-Sum and MMSE, while history of
             coronary artery bypass graft surgery, diabetes, and
             antihypertensive medications were associated with a slower
             rate of decline. There was an age interaction such that
             systolic hypertension, angina, and myocardial infarction
             were associated with greater decline with increasing
             baseline age. CONCLUSION: Atrial fibrillation, hypertension,
             and angina were associated with a greater rate of decline
             and may represent modifiable risk factors for secondary
             prevention in Alzheimer disease. The attenuated decline for
             diabetes and coronary artery bypass graft surgery may be due
             to selective survival. Some of these effects appear to vary
             with age.},
   Doi = {10.1212/01.wnl.0000279520.59792.fe},
   Key = {fds277039}
}

@article{fds277012,
   Author = {Potter, GG and Blackwell, AD and McQuoid, DR and Payne, ME and Steffens,
             DC and Sahakian, BJ and Welsh-Bohmer, KA and Krishnan,
             KRR},
   Title = {Prefrontal white matter lesions and prefrontal task
             impersistence in depressed and nondepressed
             elders.},
   Journal = {Neuropsychopharmacology},
   Volume = {32},
   Number = {10},
   Pages = {2135-2142},
   Year = {2007},
   Month = {October},
   ISSN = {0893-133X},
   url = {http://www.ncbi.nlm.nih.gov/pubmed/17299509},
   Abstract = {Poor task persistence is often observed among depressed
             individuals, and may be associated with some of the same
             frontal regions that are involved in depression. The current
             study explored the association between white-matter lesion
             volume in prefrontal cortex and noncompletion rates on a
             complex neurocognitive task among older adults in a
             treatment study for depression. Older adults in treatment
             for depression (n=83) and nondepressed (n=47) elders were
             administered the Stockings of Cambridge subtest (SoC) of the
             Cambridge Automated Neuropsychological Testing Battery
             (CANTAB) and completed a brain magnetic resonance imaging
             scan as part of an ongoing research study. Noncompletion of
             the SoC occurred in approximately 19% of depressed
             participants (16/83) and only 2% of nondepressed
             participants (1/47), which was statistically significant. In
             multivariate models, failure to complete the SoC was
             consistently and significantly associated with greater
             volume of white matter lesions in the anterior-most region
             of prefrontal cortex, particularly in the left hemisphere,
             and with greater age. Although SoC completion was not
             significantly associated with depression severity,
             noncompletion rates were significantly higher among
             unremitted individuals and those with comorbid anxiety at
             study entry. The inability to initiate behavior sufficient
             to sustain a complex neurocognitive task is a characteristic
             of geriatric depression which may be associated with
             integrity of left-prefrontal regions. Future research should
             investigate whether task impersistence is a construct that
             generalizes to other neurocognitive tasks, and if it is
             associated with other adverse outcomes in geriatric
             depression related to cerebrovascular pathology, such as
             poor treatment response.},
   Doi = {10.1038/sj.npp.1301339},
   Key = {fds277012}
}

@article{fds277085,
   Author = {Steffens, DC and Potter, GG and McQuoid, DR and MacFall, JR and Payne,
             ME and Burke, JR and Plassman, BL and Welsh-Bohmer,
             KA},
   Title = {Longitudinal magnetic resonance imaging vascular changes,
             apolipoprotein E genotype, and development of dementia in
             the neurocognitive outcomes of depression in the elderly
             study.},
   Journal = {Am J Geriatr Psychiatry},
   Volume = {15},
   Number = {10},
   Pages = {839-849},
   Year = {2007},
   Month = {October},
   ISSN = {1064-7481},
   url = {http://www.ncbi.nlm.nih.gov/pubmed/17623814},
   Abstract = {OBJECTIVE: Several studies suggest that depression is a risk
             factor for development of dementia in the elderly. In a
             study of older depressed individuals, the authors examined
             both neuroimaging and genetic factors in development of
             dementia. The authors hypothesized that change in
             subcortical gray matter and white matter hyperintensity
             volumes would be associated with development of dementia, as
             would presence of an apolipoprotein E (APOE) epsilon 4
             allele. METHODS: The sample consisted of 161 older depressed
             subjects without dementia who had magnetic resonance imaging
             scans at baseline and at two years. Blood samples were also
             taken to determine APOE genotype. All participants were
             treated with antidepressants using a guideline-based
             treatment algorithm. Their cognitive status was evaluated
             annually. A consensus panel of experts evaluated each case
             to determine cognitive status and assign a diagnosis.
             RESULTS: Twenty subjects became demented over the follow-up
             period (5.4 years on average). Change in white matter
             hyperintensity volume was significantly associated with
             development of dementia, especially among non-Alzheimer
             dementias. There was a trend for change in subcortical gray
             matter hyperintensity volume to be associated with incident
             dementia. APOE genotype was not associated with onset of
             dementia. CONCLUSION: Worsening cerebrovascular disease in
             older depressed adults is associated with cognitive decline
             and dementia, particularly of the non-Alzheimer disease
             type. The association of change in white matter lesion
             volume and incident dementia among depressed elders extends
             the vascular depression hypothesis of geriatric depression
             to include cognitive outcomes of depression in the
             elderly.},
   Doi = {10.1097/JGP.0b013e318048a1a0},
   Key = {fds277085}
}

@article{fds277037,
   Author = {Hayden, KM and Zandi, PP and Khachaturian, AS and Szekely, CA and Fotuhi, M and Norton, MC and Tschanz, JT and Pieper, CF and Corcoran, C and Lyketsos, CG and Breitner, JCS and Welsh-Bohmer, KA and Cache County
             Investigators},
   Title = {Does NSAID use modify cognitive trajectories in the elderly?
             The Cache County study.},
   Journal = {Neurology},
   Volume = {69},
   Number = {3},
   Pages = {275-282},
   Year = {2007},
   Month = {July},
   url = {http://www.ncbi.nlm.nih.gov/pubmed/17636065},
   Abstract = {BACKGROUND: Epidemiologic studies have suggested that
             nonsteroidal anti-inflammatory drugs (NSAIDs) may be useful
             for the prevention of Alzheimer disease (AD). By contrast,
             clinical trials have not supported NSAID use to delay or
             treat AD. Few studies have evaluated cognitive trajectories
             of NSAID users over time. METHODS: Residents of Cache
             County, UT, aged 65 or older on January 1, 1995, were
             invited to participate in the study. At baseline,
             participants provided a detailed inventory of their
             medications and completed a revised Modified Mini-Mental
             State Examination (3MS). Participants (n = 3,383) who were
             cognitively normal at baseline were re-examined after 3 and
             8 years. The association between NSAID use and 3MS scores
             over time was estimated using random effects modeling.
             RESULTS: Associations depended upon when NSAIDs were started
             and APOE genotype. In participants who started NSAID use
             prior to age 65, those with no APOE epsilon4 alleles
             performed similarly to nonusers (a difference of 0.10 points
             per year; p = 0.19), while those with one or more epsilon4
             allele(s) showed more protection (0.40 points per year; p =
             0.0005). Among participants who first used NSAIDs at or
             after age 65, those with one or more epsilon4 alleles had
             higher baseline scores (0.95 points; p = 0.03) but did not
             show subsequent difference in change in score over time
             (0.06 points per year; p = 0.56). Those without an epsilon4
             allele who started NSAID use after age 65 showed greater
             decline than nonusers (-0.16 points per year; p = 0.02).
             CONCLUSIONS: Nonsteroidal anti-inflammatory drug use may
             help to prevent cognitive decline in older adults if started
             in midlife rather than late life. This effect may be more
             notable in those who have one or more APOE epsilon4
             alleles.},
   Doi = {10.1212/01.wnl.0000265223.25679.2a},
   Key = {fds277037}
}

@article{fds277084,
   Author = {Fearing, MA and Bigler, ED and Norton, M and Tschanz, JA and Hulette, C and Leslie, C and Welsh-Bohmer, K and Cache County
             Investigators},
   Title = {Autopsy-confirmed Alzheimer's disease versus clinically
             diagnosed Alzheimer's disease in the Cache County Study on
             Memory and Aging: a comparison of quantitative MRI and
             neuropsychological findings.},
   Journal = {J Clin Exp Neuropsychol},
   Volume = {29},
   Number = {5},
   Pages = {553-560},
   Year = {2007},
   Month = {July},
   ISSN = {1380-3395},
   url = {http://www.ncbi.nlm.nih.gov/pubmed/17564920},
   Abstract = {Atrophy of specific, regional, and generalized brain
             structures occurs as a result of the Alzheimer's disease
             (AD) process. Comparing AD patients with histopathological
             confirmation of the disease at autopsy to those without
             autopsy but who were clinically diagnosed using the same
             antemortem criteria will provide further evidence of the
             utility and accuracy of neuropsychological assessments at
             the time of diagnosis, as well as the efficacy of
             quantitative magnetic resonance imaging (qMRI) in
             demonstrating gross neuropathological changes associated
             with the disease. The Cache County Study of Aging provides a
             unique opportunity to determine how closely AD subjects with
             only the clinical diagnosis match similarly diagnosed AD
             subjects but with postmortem confirmation of the disease.
             qMRI volumes of various brain structures, as well as
             neuropsychological outcome measures from an expanded
             battery, were obtained in 31 autopsy-confirmed AD subjects
             and 45 clinically diagnosed AD subjects. Of the various qMRI
             variables examined, only total temporal lobe volume was
             different, where those with postmortem confirmation had
             reduced volume. No significant differences between the two
             groups were found with any of the neuropsychological outcome
             measures. These findings confirm the similarity in
             neuroimaging and neuropsychological assessment findings
             between those with just the clinical diagnosis of AD and
             those with an autopsy-confirmed diagnosis in the
             moderate-to-severe stage of the disease at the time of
             diagnosis.},
   Doi = {10.1080/13803390600826579},
   Key = {fds277084}
}

@article{fds371177,
   Author = {Hulette, CM and Ervin, JF and Steed, E and Szymanski, M and Burke, J and Welsh-Bohmer, K},
   Title = {Arteriolar ApoE expression is increased Alzheimer disease
             cortex},
   Journal = {Journal of Neuropathology and Experimental
             Neurology},
   Volume = {66},
   Number = {5},
   Pages = {433-433},
   Publisher = {Oxford University Press (OUP)},
   Year = {2007},
   Month = {May},
   url = {http://dx.doi.org/10.1097/01.jnen.0000268866.50903.81},
   Doi = {10.1097/01.jnen.0000268866.50903.81},
   Key = {fds371177}
}

@article{fds276952,
   Author = {Onyike, CU and Sheppard, J-ME and Tschanz, JT and Norton, MC and Green,
             RC and Steinberg, M and Welsh-Bohmer, KA and Breitner, JC and Lyketsos,
             CG},
   Title = {Epidemiology of apathy in older adults: the Cache County
             Study.},
   Journal = {Am J Geriatr Psychiatry},
   Volume = {15},
   Number = {5},
   Pages = {365-375},
   Year = {2007},
   Month = {May},
   ISSN = {1064-7481},
   url = {http://www.ncbi.nlm.nih.gov/pubmed/17463187},
   Abstract = {OBJECTIVES: The objectives of this study are to describe the
             distribution of apathy in community-based older adults and
             to investigate its relationships with cognition and
             day-to-day functioning. METHODS: Data from the Cache County
             Study on Memory, Health and Aging were used to estimate the
             frequency of apathy in groups of elders defined by
             demographic, cognitive, and functional status and to examine
             the associations of apathy with impairments of cognition and
             day-to-day functioning. RESULTS: Apathy was measured with
             the Neuropsychiatric Inventory. Clinical apathy
             (Neuropsychiatric Inventory score > or = 4) was found in
             1.4% of individuals classified as cognitively normal, 3.1%
             of those with a mild cognitive syndrome, and 17.3% of those
             with dementia. Apathy status was associated with cognitive
             and functional impairments and higher levels of stress
             experienced by caregivers. Among participants with normal
             cognition, apathy was associated with worse performance on
             the Mini-Mental State Examination, the Boston Naming and
             Animal Fluency tests, and the Trail Making Test-Part B. The
             association of apathy with cognitive impairment was
             independent of its association with Neuropsychiatric
             Inventory depression. CONCLUSIONS: In a cohort of
             community-based older adults, the frequency and severity of
             apathy is positively correlated with the severity of
             cognitive impairment. In addition, apathy is associated with
             cognitive and functional impairments in elders adjudged to
             have normal cognition. The results suggest that apathy is an
             early sign of cognitive decline and that delineating
             phenotypes in which apathy and a mild cognitive syndrome
             co-occur may facilitate earlier identification of
             individuals at risk for dementia.},
   Doi = {10.1097/01.JGP.0000235689.42910.0d},
   Key = {fds276952}
}

@article{fds371178,
   Author = {Hulette, CM and Ervin, JF and Steed, E and Szymanski, M and Burke, J and Welsh-Bohmer, K},
   Title = {Arteriolar ApoE expression is increased Alzheimer disease
             cortex.},
   Journal = {FASEB JOURNAL},
   Volume = {21},
   Number = {5},
   Pages = {A72-A72},
   Publisher = {FEDERATION AMER SOC EXP BIOL},
   Year = {2007},
   Month = {April},
   Key = {fds371178}
}

@article{fds276951,
   Author = {Sugarman, J and Roter, D and Cain, C and Wallace, R and Schmechel, D and Welsh-Bohmer, KA},
   Title = {Proxies and consent discussions for dementia
             research.},
   Journal = {J Am Geriatr Soc},
   Volume = {55},
   Number = {4},
   Pages = {556-561},
   Year = {2007},
   Month = {April},
   ISSN = {0002-8614},
   url = {http://www.ncbi.nlm.nih.gov/pubmed/17397434},
   Abstract = {OBJECTIVES: To better understand the nature of informed
             consent encounters for research involving patients with
             dementia that requires proxy consent. DESIGN: Audiotaping of
             informed-consent encounters for a study of genetic markers
             for sporadic Alzheimer's disease. SETTING: Outpatients at an
             Alzheimer's disease research center. PARTICIPANTS: Patients
             with dementia and their companions. MEASUREMENTS: Audiotapes
             were analyzed to characterize communication style and
             coverage of the standard elements of informed consent and,
             using the Roter Interaction Analysis System, to capture the
             dynamics of three-way interaction between the patient, their
             companion, and the physician investigator. RESULTS: Of 26
             informed consent encounters, all involved a patient, a
             companion, and a physician. Patients had a mean Mini-Mental
             State Examination (MMSE) score of 21.8. For patients, 49% of
             their interactions involved agreement and approval (positive
             statements), 16% psychosocial information, 7% biomedical
             information, 7% asking questions, and 7% expressing emotion.
             Companion interactions involved 37% positive statements and
             19% biomedical information. Physician interactions involved
             emotional expressiveness (30%) and positive statements
             (19%). Discussion length was positively related to MMSE
             score (Spearman rho=0.45; P<.02). Coverage of required
             elements of informed consent was fairly comprehensive and
             had no relationship to patients' MMSE scores. CONCLUSION:
             These data should inform policies regarding the ethically
             appropriate ways of conducting research with cognitively
             impaired adults. For example, patients in this study were
             more silent than their companions and the physician, but
             when patients spoke, they primarily agreed with what was
             said. Although this might first seem to signal assent, such
             an interpretation should be made with caution for persons
             with dementia. In addition, previous work on informed
             consent has focused on its cognitive aspects, but these data
             reveal that the emotional and social dimensions warrant
             attention.},
   Doi = {10.1111/j.1532-5415.2007.01101.x},
   Key = {fds276951}
}

@article{fds277023,
   Author = {Hulette, CM and Welsh-Bohmer, K},
   Title = {Coronary artery disease is associated with Alzheimer disease
             neuropathology in APOE4 carriers.},
   Journal = {Neurology},
   Volume = {68},
   Number = {6},
   Pages = {471},
   Year = {2007},
   Month = {February},
   url = {http://www.ncbi.nlm.nih.gov/pubmed/17283328},
   Doi = {10.1212/01.wnl.0000256286.78188.dd},
   Key = {fds277023}
}

@article{fds276988,
   Author = {Lee, JS and Potter, GG and Wagner, HR and Welsh-Bohmer, KA and Steffens,
             DC},
   Title = {Persistent mild cognitive impairment in geriatric
             depression.},
   Journal = {Int Psychogeriatr},
   Volume = {19},
   Number = {1},
   Pages = {125-135},
   Year = {2007},
   Month = {February},
   ISSN = {1041-6102},
   url = {http://www.ncbi.nlm.nih.gov/pubmed/16834811},
   Abstract = {BACKGROUND: Cognitive impairment often occurs with geriatric
             depression and impairments may persist despite remission of
             depression. Although clinical definitions of mild cognitive
             impairment (MCI) have typically excluded depression, a
             neuropsychological model of MCI in depression has utility
             for identifying individuals whose cognitive impairments may
             persist or progress to dementia. METHODS: At baseline and
             1-year follow-up, 67 geriatric patients with depression had
             a comprehensive clinical examination that included
             depression assessment and neuropsychological testing. We
             defined MCI by a neuropsychological algorithm and examined
             the odds of MCI classification at Year 1 for remitted
             depressed individuals with baseline MCI, and examined
             clinical, functional and genetic factors associated with
             MCI. RESULTS: Fifty-four percent of the sample had MCI at
             baseline. Odds of MCI classification at Year 1 were four
             times greater among patients with baseline MCI than those
             without. Instrumental activities of daily living were
             associated with MCI at Year 1, while age and APOE genotype
             was not. CONCLUSIONS: These results confirm previous
             observations that MCI is highly prevalent among older
             depressed adults and that cognitive impairment occurring
             during acute depression may persist after depression remits.
             Self-reported decline in functional activities may be a
             marker for persistent cognitive impairment, which suggests
             that assessments of both neuropsychological and functional
             status are important prognostic factors in the evaluation of
             geriatric depression.},
   Doi = {10.1017/S1041610206003607},
   Key = {fds276988}
}

@article{fds277036,
   Author = {Hayden, KM and Welsh-Bohmer, KA and Wengreen, HJ and Zandi, PP and Lyketsos, CG and Breitner, JCS and Cache County
             Investigators},
   Title = {Risk of mortality with vitamin E supplements: the Cache
             County study.},
   Journal = {Am J Med},
   Volume = {120},
   Number = {2},
   Pages = {180-184},
   Year = {2007},
   Month = {February},
   url = {http://www.ncbi.nlm.nih.gov/pubmed/17275460},
   Abstract = {PURPOSE: A recent meta-analysis reported increased mortality
             in clinical trial participants randomized to high-dose
             vitamin E. We sought to determine whether these mortality
             risks with vitamin E reflect adverse consequences of its use
             in the presence of cardiovascular disease. METHODS: In a
             defined population aged 65 years or older, baseline
             interviews captured self- or proxy-reported history of
             cardiovascular illness. A medicine cabinet inventory
             verified nutritional supplement and medication use. Three
             sources identified subsequent deaths. Cox proportional
             hazards methods examined the association between vitamin E
             use and mortality. RESULTS: After adjustment for age and
             sex, there was no association in this population between
             vitamin E use and mortality (adjusted hazard ratio [aHR]
             0.93; 95% confidence interval [CI], 0.74-1.15). Predictably,
             deaths were more frequent with a history of diabetes,
             stroke, coronary artery bypass graft surgery, or myocardial
             infarction, and with the use of warfarin, nitrates, or
             diuretics. None of these conditions or treatments altered
             the null main effect with vitamin E, but mortality was
             increased in vitamin E users who had a history of stroke
             (aHR 3.64; CI, 1.73-7.68), coronary bypass graft surgery
             (aHR 4.40; CI, 2.83-6.83), or myocardial infarction (aHR
             1.95; CI, 1.29-2.95) and, independently, in those taking
             nitrates (aHR 3.95; CI, 2.04-7.65), warfarin (aHR 3.71; CI,
             2.22-6.21), or diuretics (aHR 1.83; CI, 1.35-2.49). Although
             not definitive, a consistent trend toward reduced mortality
             was seen in vitamin E users without these conditions or
             treatments. CONCLUSIONS: In this population-based study,
             vitamin E use was unrelated to mortality, but this
             apparently null finding seems to represent a combination of
             increased mortality in those with severe cardiovascular
             disease and a possible protective effect in those
             without.},
   Doi = {10.1016/j.amjmed.2006.03.039},
   Key = {fds277036}
}

@article{fds277022,
   Author = {Trembath, D and Ervin, JF and Broom, L and Szymanski, M and Welsh-Bohmer, K and Pieper, C and Hulette, CM},
   Title = {The distribution of cerebrovascular amyloid in Alzheimer's
             disease varies with ApoE genotype.},
   Journal = {Acta Neuropathol},
   Volume = {113},
   Number = {1},
   Pages = {23-31},
   Year = {2007},
   Month = {January},
   ISSN = {0001-6322},
   url = {http://www.ncbi.nlm.nih.gov/pubmed/17089130},
   Abstract = {We performed a comparative study to assess cerebral amyloid
             angiopathy and ApoE genotype in cases of Alzheimer's disease
             (AD). Ten ApoE 3,3 and ten ApoE 4,4 AD brains, as well as
             ten normal control brains, were selected after matching for
             age, sex, and duration of disease. Sections of middle
             frontal and inferior parietal cortex including white matter
             sections were stained with an antibody against amyloid beta
             (Abeta), and extensive analysis of arteriolar Abeta
             deposition was performed using digital image analysis.
             Quantification of the staining revealed a larger
             cross-section of arteriolar walls occupied by Abeta in ApoE
             4,4 and ApoE 3,3 AD subjects compared to controls. Our
             results show Abeta deposition in gray matter and white
             matter arterioles was predominantly found in ApoE 4,4 brains
             and, overall, Abeta deposition was greatest in these cases.
             This observation implies that there is greater vascular
             amyloid deposition (particularly in the white matter
             arterioles) in ApoE 4,4 AD individuals compared to ApoE 3,3
             AD. These observations may give insight into the etiology
             behind the increased risk for AD associated with the
             ApoE-epsilon4 allele and the pathogenesis of vascular Abeta
             deposition.},
   Doi = {10.1007/s00401-006-0162-9},
   Key = {fds277022}
}

@article{fds276953,
   Author = {Beekly, DL and Ramos, EM and Lee, WW and Deitrich, WD and Jacka, ME and Wu,
             J and Hubbard, JL and Koepsell, TD and Morris, JC and Kukull, WA and NIA
             Alzheimer's Disease Centers},
   Title = {The National Alzheimer's Coordinating Center (NACC)
             database: the Uniform Data Set.},
   Journal = {Alzheimer Dis Assoc Disord},
   Volume = {21},
   Number = {3},
   Pages = {249-258},
   Year = {2007},
   ISSN = {0893-0341},
   url = {http://dx.doi.org/10.1097/WAD.0b013e318142774e},
   Abstract = {The National Alzheimer's Coordinating Center (NACC) is
             responsible for developing and maintaining a database of
             participant information collected from the 29 Alzheimer's
             Disease Centers (ADCs) funded by the National Institute on
             Aging (NIA). The NIA appointed the ADC Clinical Task Force
             to determine and define an expanded, standardized clinical
             data set, called the Uniform Data Set (UDS). The goal of the
             UDS is to provide ADC researchers a standard set of
             assessment procedures, collected longitudinally, to better
             characterize ADC participants with mild Alzheimer disease
             and mild cognitive impairment in comparison with nondemented
             controls. NACC implemented the UDS (September 2005) by
             developing data collection forms for initial and follow-up
             visits based on Clinical Task Force definitions, a
             relational database, and a data submission system accessible
             by all ADCs. The NIA requires ADCs to submit UDS data to
             NACC for all their Clinical Core participants. Thus, the
             NACC web site (https://www.alz.washington.edu) was enhanced
             to provide efficient and secure access data submission and
             retrieval systems.},
   Doi = {10.1097/WAD.0b013e318142774e},
   Key = {fds276953}
}

@article{fds277038,
   Author = {Wengreen, HJ and Munger, RG and Corcoran, CD and Zandi, P and Hayden,
             KM and Fotuhi, M and Skoog, I and Norton, MC and Tschanz, J and Breitner,
             JCS and Welsh-Bohmer, KA},
   Title = {Antioxidant intake and cognitive function of elderly men and
             women: the Cache County Study.},
   Journal = {J Nutr Health Aging},
   Volume = {11},
   Number = {3},
   Pages = {230-237},
   Year = {2007},
   ISSN = {1279-7707},
   url = {http://www.ncbi.nlm.nih.gov/pubmed/17508099},
   Abstract = {OBJECTIVE: We prospectively examined associations between
             intakes of antioxidants (vitamins C, vitamin E, and
             carotene) and cognitive function and decline among elderly
             men and women of the Cache County Study on Memory and Aging
             in Utah. PARTICIPANTS AND DESIGN: In 1995, 3831 residents 65
             years of age or older completed a baseline survey that
             included a food frequency questionnaire and cognitive
             assessment. Cognitive function was assessed using an adapted
             version of the Modified Mini-Mental State examination (3MS)
             at baseline and at three subsequent follow-up interviews
             spanning approximately 7 years. Multivariable-mixed models
             were used to estimate antioxidant nutrient effects on
             average 3MS score over time. RESULTS: Increasing quartiles
             of vitamin C intake alone and combined with vitamin E were
             associated with higher baseline average 3MS scores (p-trend
             = 0.013 and 0.02 respectively); this association appeared
             stronger for food sources compared to supplement or food and
             supplement sources combined. Study participants with lower
             levels of intake of vitamin C, vitamin E and carotene had a
             greater acceleration of the rate of 3MS decline over time
             compared to those with higher levels of intake. CONCLUSION:
             High antioxidant intake from food and supplement sources of
             vitamin C, vitamin E, and carotene may delay cognitive
             decline in the elderly.},
   Key = {fds277038}
}

@article{fds277083,
   Author = {Heinzen, EL and Yoon, W and Weale, ME and Sen, A and Wood, NW and Burke,
             JR and Welsh-Bohmer, KA and Hulette, CM and Sisodiya, SM and Goldstein,
             DB},
   Title = {Alternative ion channel splicing in mesial temporal lobe
             epilepsy and Alzheimer's disease.},
   Journal = {Genome Biol},
   Volume = {8},
   Number = {3},
   Pages = {R32},
   Year = {2007},
   url = {http://www.ncbi.nlm.nih.gov/pubmed/17343748},
   Abstract = {BACKGROUND: Alternative gene transcript splicing permits a
             single gene to produce multiple proteins with varied
             functions. Bioinformatic investigations have identified
             numerous splice variants, but whether these transcripts are
             translated to functional proteins and the physiological
             significance of these alternative proteins are largely
             unknown. Through direct identification of splice variants
             associated with disease states, we can begin to address
             these questions and to elucidate their roles in disease
             predisposition and pathophysiology. This work specifically
             sought to identify disease-associated alternative splicing
             patterns in ion channel genes by comprehensively screening
             affected brain tissue collected from patients with mesial
             temporal lobe epilepsy and Alzheimer's disease. New
             technology permitting the screening of alternative splice
             variants in microarray format was employed. Real time
             quantitative PCR was used to verify observed splice variant
             patterns. RESULTS: This work shows for the first time that
             two common neurological conditions are associated with
             extensive changes in gene splicing, with 25% and 12% of the
             genes considered having significant changes in splicing
             patterns associated with mesial temporal lobe epilepsy and
             Alzheimer's disease, respectively. Furthermore, these
             changes were found to exhibit unique and consistent patterns
             within the disease groups. CONCLUSION: This work has
             identified a set of disease-associated, alternatively
             spliced gene products that represent high priorities for
             detailed functional investigations into how these changes
             impact the pathophysiology of mesial temporal lobe epilepsy
             and Alzheimer's disease.},
   Doi = {10.1186/gb-2007-8-3-r32},
   Key = {fds277083}
}

@article{fds371179,
   Author = {Khachaturian, AS and Zandi, PP and Lyketsos, CG and Hayden, KM and Skoog, I and Norton, MC and Tschanz, JT and Mayer, L and Welsh-Bohmer,
             KA and Breitner, JC},
   Title = {Antihypertensive medication use and incident Alzheimer
             disease: The cache county study},
   Journal = {NEUROPSYCHOPHARMACOLOGY},
   Volume = {31},
   Pages = {S39-S39},
   Publisher = {NATURE PUBLISHING GROUP},
   Year = {2006},
   Month = {December},
   Key = {fds371179}
}

@article{fds277082,
   Author = {Plassman, BL and Steffens, DC and Burke, JR and Welsh-Bohmer, KA and Newman, TN and Drosdick, D and Helms, MJ and Potter, GG and Breitner,
             JCS},
   Title = {Duke Twins Study of Memory in Aging in the NAS-NRC Twin
             Registry.},
   Journal = {Twin Res Hum Genet},
   Volume = {9},
   Number = {6},
   Pages = {950-957},
   Year = {2006},
   Month = {December},
   ISSN = {1832-4274},
   url = {http://www.ncbi.nlm.nih.gov/pubmed/17254435},
   Abstract = {The Duke Twins Study of Memory in Aging is an ongoing,
             longitudinal study of cognitive change and dementia in the
             population-based National Academy of Sciences-National
             Research Council (NAS-NRC) Twin Registry of World War II
             Male Veterans. The primary goal of this study has been to
             estimate the overall genetic and environmental contributions
             to dementia with a specific focus on Alzheimer's disease. An
             additional goal has been to examine specific genetic and
             environmental antecedents of cognitive decline and dementia.
             Since 1989, we have completed 4 waves of data collection.
             Each wave included a 2-phase telephone cognitive screening
             protocol, followed by an in-home standardized clinical
             assessment for those with suspected dementia. For many
             participants, we have obtained postmortem neuropathological
             confirmation of the diagnosis of dementia. In addition to
             data on cognition, we have also collected information on
             occupational history, medical history, medications and other
             lifetime experiences that may influence cognitive function
             in late life. We provide an overview of the study's
             methodology and describe the focus of recent
             research.},
   Doi = {10.1375/183242706779462381},
   Key = {fds277082}
}

@article{fds371180,
   Author = {Ervin, JF and Heinzen, EL and Goldstein, D and Szymanski, MH and Burke,
             JR and Welsh-Bohmer, KA and Hulette, CM},
   Title = {Human brain tissue may be retrieved up to 30 hours
             post-mortem with little effect on RNA or protein
             integrity},
   Journal = {BRAIN PATHOLOGY},
   Volume = {16},
   Pages = {S125-S125},
   Publisher = {BLACKWELL PUBLISHING},
   Year = {2006},
   Month = {September},
   Key = {fds371180}
}

@article{fds276978,
   Author = {Steinberg, M and Corcoran, C and Tschanz, JT and Huber, C and Welsh-Bohmer, K and Norton, MC and Zandi, P and Breitner, JCS and Steffens, DC and Lyketsos, CG},
   Title = {Risk factors for neuropsychiatric symptoms in dementia: the
             Cache County Study.},
   Journal = {Int J Geriatr Psychiatry},
   Volume = {21},
   Number = {9},
   Pages = {824-830},
   Year = {2006},
   Month = {September},
   ISSN = {0885-6230},
   url = {http://dx.doi.org/10.1002/gps.1567},
   Abstract = {OBJECTIVE: To investigate the probability of individual
             neuropsychiatric symptoms in dementia patients as a function
             of eight risk factors. METHODS: In the Cache County Study,
             we administered the Neuropsychiatric Inventory (NPI) to 328
             dementia patients at baseline. Approximately 18 months
             later, we re-administered the NPI to 184 participants
             available for follow-up. Generalized estimating equation
             methods were used to model the probability of individual
             neuropsychiatric symptoms as a function of: gender, age,
             education, dementia type and severity, APOE status, time of
             observation, and general medical health. RESULTS: Women
             showed increased tendency toward anxiety, [odds ratio (OR)
             2.22, 95% confidence interval (CI) 1.31-3.76] and delusions
             (OR 2.15, CI 1.22-3.78), but older persons of both sexes
             showed less tendency toward anxiety. Dementia severity
             increased the tendency toward hallucinations and agitation
             (OR 2.42, CI 1.81-3.23) and decreased risk of depression.
             Positive APOE epsilon4 status increased the tendency toward
             aberrant motor behavior (OR 1.84, CI 1.05-3.22). Among
             dementia diagnoses, those with Alzheimer's disease showed
             decreased tendency toward agitation (OR 0.58, CI 0.35-0.95),
             depression (OR 0.56, CI 0.33-0.96) and disinhibition (OR
             0.46, CI 0.24-0.88). Later time of observation increased
             risk of aberrant motor behavior and delusions, and more
             serious medical comorbidity increased risk of, agitation,
             irritability, disinhibition, and aberrant motor behavior.
             CONCLUSIONS: Gender, age, dementia severity, APOE epsilon4,
             dementia diagnosis, time of observation, and general medical
             health appear to influence the occurrence of individual
             neuropsychiatric symptoms.},
   Doi = {10.1002/gps.1567},
   Key = {fds276978}
}

@article{fds277111,
   Author = {Newman, MF and Mathew, JP and Grocott, HP and Mackensen, GB and Monk, T and Welsh-Bohmer, KA and Blumenthal, JA and Laskowitz, DT and Mark,
             DB},
   Title = {Central nervous system injury associated with cardiac
             surgery.},
   Journal = {Lancet},
   Volume = {368},
   Number = {9536},
   Pages = {694-703},
   Year = {2006},
   Month = {August},
   url = {http://www.ncbi.nlm.nih.gov/pubmed/16920475},
   Abstract = {Millions of individuals with coronary artery or valvular
             heart disease have been given a new chance at life by heart
             surgery, but the potential for neurological injury is an
             Achilles heel. Technological advancements and innovations in
             surgical and anaesthetic technique have allowed us to offer
             surgical treatment to patients at the extremes of age and
             infirmity-the group at greatest risk for neurological
             injury. Neurocognitive dysfunction is a complication of
             cardiac surgery that can restrict the improved quality of
             life that patients usually experience after heart surgery.
             With a broader understanding of the frequency and effects of
             neurological injury from cardiac surgery and its
             implications for patients in both the short term and the
             long term, we should be able to give personalised treatments
             and thus preserve both their quantity and quality of life.
             We describe these issues and the controversies that merit
             continued investigation.},
   Doi = {10.1016/S0140-6736(06)69254-4},
   Key = {fds277111}
}

@article{fds277067,
   Author = {Li, Y-J and Scott, WK and Zhang, L and Lin, P-I and Oliveira, SA and Skelly, T and Doraiswamy, MP and Welsh-Bohmer, KA and Martin, ER and Haines, JL and Pericak-Vance, MA and Vance, JM},
   Title = {Revealing the role of glutathione S-transferase omega in
             age-at-onset of Alzheimer and Parkinson diseases.},
   Journal = {Neurobiol Aging},
   Volume = {27},
   Number = {8},
   Pages = {1087-1093},
   Year = {2006},
   Month = {August},
   url = {http://www.ncbi.nlm.nih.gov/pubmed/15985314},
   Abstract = {We previously reported a linkage region on chromosome 10q
             for age-at-onset (AAO) of Alzheimer (AD) and Parkinson (PD)
             diseases. Glutathione S-transferase, omega-1 (GSTO1) and the
             adjacent gene GSTO2, located in this linkage region, were
             then reported to associate with AAO of AD and PD. To examine
             whether GSTO1 and GSTO2 (hereafter referred to as GSTO1h)
             are responsible for the linkage evidence, we identified 39
             families in AD that lead to our previous linkage and
             association findings. The evidence of linkage and
             association was markedly diminished after removing these 39
             families from the analyses, thus providing support that
             GSTO1h drives the original linkage results. The maximum
             average AAO delayed by GSTO1h SNP 7-1 (rs4825, A nucleotide)
             was 6.8 (+/-4.41) years for AD and 8.6(+/-5.71) for PD,
             respectively. This is comparable to the magnitude of AAO
             difference by APOE-4 in these same AD and PD families. These
             findings suggest the presence of genetic heterogeneity for
             GSTO1h's effect on AAO, and support GSTO1h's role in
             modifying AAO in these two disorders.},
   Doi = {10.1016/j.neurobiolaging.2005.05.013},
   Key = {fds277067}
}

@article{fds277035,
   Author = {Tschanz, JT and Welsh-Bohmer, KA and Lyketsos, CG and Corcoran, C and Green, RC and Hayden, K and Norton, MC and Zandi, PP and Toone, L and West,
             NA and Breitner, JCS and Cache County Investigators},
   Title = {Conversion to dementia from mild cognitive disorder: the
             Cache County Study.},
   Journal = {Neurology},
   Volume = {67},
   Number = {2},
   Pages = {229-234},
   Year = {2006},
   Month = {July},
   url = {http://www.ncbi.nlm.nih.gov/pubmed/16864813},
   Abstract = {OBJECTIVE: To examine 3-year rates of conversion to
             dementia, and risk factors for such conversion, in a
             population-based sample with diverse types of cognitive
             impairment. METHODS: All elderly (aged 65 or older)
             residents of Cache County, UT, were invited to undergo two
             waves of dementia screening and assessment. Three-year
             follow-up data were available for 120 participants who had
             some form of mild cognitive impairment at baseline. Of
             these, 51 had been classified at baseline with prodromal
             Alzheimer disease (proAD), and 69 with other cognitive
             syndromes (CS). RESULTS: Three-year rates of conversion to
             dementia were 46% among those with cognitive impairment at
             baseline. By comparison, 3.3% without impairment converted
             to dementia in the interval. Among converters, AD was the
             most common type of dementia. In individuals with at least
             one APOE epsilon4 allele, those with proAD or CS exhibited a
             22- to 25-fold higher risk of dementia than cognitively
             unimpaired individuals (vs 5- to 10-fold higher risk in
             those without epsilon4). CONCLUSIONS: Individuals with all
             types of mild cognitive impairment have an elevated risk of
             dementia over 3 years, more so in those with an APOE
             epsilon4 allele. These results suggest value in dementia
             surveillance for broad groups of cognitively impaired
             individuals beyond any specific category, and utility of
             APOE genotyping as a prognostic method.},
   Doi = {10.1212/01.wnl.0000224748.48011.84},
   Key = {fds277035}
}

@article{fds276950,
   Author = {Lin, PI and Martin, ER and Bronson, PG and Browning-Large, C and Small,
             GW and Schmechel, DE and Welsh-Bohmer, KA and Haines, JL and Gilbert,
             JR and Pericak-Vance, MA},
   Title = {Exploring the association of glyceraldehyde-3-phosphate
             dehydrogenase gene and Alzheimer disease.},
   Journal = {Neurology},
   Volume = {67},
   Number = {1},
   Pages = {64-68},
   Year = {2006},
   Month = {July},
   url = {http://www.ncbi.nlm.nih.gov/pubmed/16832079},
   Abstract = {BACKGROUND: Previous linkage studies have shown that
             chromosome 12 harbors susceptibility genes for late-onset
             Alzheimer disease (LOAD). However, association studies of
             several candidate genes on this chromosome region have
             produced ambiguous results. A recent study reported the
             association between the glyceraldehyde-3-phosphate
             dehydrogenase (GAPD) gene on chromosome 12p and the risk of
             LOAD. METHODS: The authors conducted family-based and
             case-control association studies in two independent LOAD
             data sets on 12 single-nucleotide polymorphisms (SNPs) in
             the GAPD gene and its paralogs. RESULTS: No association was
             found of the GAPD gene with LOAD in the family-based data
             set, but marginal evidence of association was seen in the
             later-onset subgroup when age at onset was stratified. The
             SNP rs2029721 in one GAPD pseudogene was also found to be
             associated with risk for LOAD in the unrelated case-control
             data set (p = 0.003). CONCLUSIONS: The GAPD gene and its
             pseudogene may play a role in the development of late-onset
             Alzheimer disease. However, the effect, if any, is likely to
             be limited.},
   Doi = {10.1212/01.wnl.0000223438.90113.4e},
   Key = {fds276950}
}

@article{fds277034,
   Author = {Welsh-Bohmer, KA and Breitner, JCS and Hayden, KM and Lyketsos, C and Zandi, PP and Tschanz, JT and Norton, MC and Munger,
             R},
   Title = {Modifying dementia risk and trajectories of cognitive
             decline in aging: the Cache County Memory
             Study.},
   Journal = {Alzheimers Dement},
   Volume = {2},
   Number = {3},
   Pages = {257-260},
   Year = {2006},
   Month = {July},
   url = {http://www.ncbi.nlm.nih.gov/pubmed/19595895},
   Abstract = {The Cache County Study of Memory, Health, and Aging, more
             commonly referred to as the "Cache County Memory Study
             (CCMS)" is a longitudinal investigation of aging and
             Alzheimer's disease (AD) based in an exceptionally
             long-lived population residing in northern Utah. The study
             begun in 1994 has followed an initial cohort of 5,092 older
             individuals (many over age 84) and has examined the
             development of cognitive impairment and dementia in relation
             to genetic and environmental antecedents. This article
             summarizes the major contributions of the CCMS towards the
             understanding of mild cognitive disorders and AD across the
             lifespan, underscoring the role of common health exposures
             in modifying dementia risk and trajectories of cognitive
             change. The study now in its fourth wave of ascertainment
             illustrates the role of population-based approaches in
             informing testable models of cognitive aging and Alzheimer's
             disease.},
   Doi = {10.1016/j.jalz.2006.04.011},
   Key = {fds277034}
}

@article{fds277065,
   Author = {Sheline, YI and Barch, DM and Garcia, K and Gersing, K and Pieper, C and Welsh-Bohmer, K and Steffens, DC and Doraiswamy,
             PM},
   Title = {Cognitive function in late life depression: relationships to
             depression severity, cerebrovascular risk factors and
             processing speed.},
   Journal = {Biol Psychiatry},
   Volume = {60},
   Number = {1},
   Pages = {58-65},
   Year = {2006},
   Month = {July},
   ISSN = {0006-3223},
   url = {http://www.ncbi.nlm.nih.gov/pubmed/16414031},
   Abstract = {BACKGROUND: A number of studies have examined clinical
             factors linked to worse neuropsychological performance in
             late life depression (LLD). To understand the influence of
             LLD on cognition, it is important to determine if deficits
             in a number of cognitive domains are relatively independent,
             or mediated by depression- related deficits in a basic
             domain such as processing speed. METHODS: Patients who met
             DSM-IV criteria for major depression (n = 155) were
             administered a comprehensive neuropsychological battery of
             tasks grouped into episodic memory, language, working
             memory, executive function, and processing speed domains.
             Multiple regression analyses were conducted to determine
             contributions of predictor variables to cognitive domains.
             RESULTS: Age, depression severity, education, race and
             vascular risk factors all made significant and independent
             contributions to one or more domains of cognitive function,
             with all five making independent contributions to processing
             speed. Age of onset made no independent contribution, after
             accounting for age and vascular risk factors. Of the five
             cognitive domains investigated, changes in processing speed
             were found to most fully mediate the influence of predictor
             variables on all other cognitive domains. CONCLUSIONS: While
             slowed processing speed appears to be the most core
             cognitive deficit in LLD, it was closely followed by
             executive function as a core cognitive deficit. Future
             research is needed to help clarify mechanisms leading to
             LLD- related changes in processing speed, including the
             potential role of white matter abnormalities.},
   Doi = {10.1016/j.biopsych.2005.09.019},
   Key = {fds277065}
}

@article{fds277066,
   Author = {Lin, P-I and Martin, ER and Browning-Large, CA and Schmechel, DE and Welsh-Bohmer, KA and Doraiswamy, PM and Gilbert, JR and Haines, JL and Pericak-Vance, MA},
   Title = {Parsing the genetic heterogeneity of chromosome 12q
             susceptibility genes for Alzheimer disease by family-based
             association analysis.},
   Journal = {Neurogenetics},
   Volume = {7},
   Number = {3},
   Pages = {157-165},
   Year = {2006},
   Month = {July},
   ISSN = {1364-6745},
   url = {http://www.ncbi.nlm.nih.gov/pubmed/16770605},
   Abstract = {Previous linkage studies have suggested that chromosome 12
             may harbor susceptibility genes for late-onset Alzheimer
             disease (LOAD). No risk genes on chromosome 12 have been
             conclusively identified yet. We have reported that the
             linkage evidence for LOAD in a 12q region was significantly
             increased in autopsy-confirmed families particularly for
             those showing no linkage to alpha-T catenin gene, a LOAD
             candidate gene on chromosome 10 [LOD score increased from
             0.1 in the autopsy-confirmed subset to 4.19 in the unlinked
             subset (optimal subset); p<0.0001 for the increase in LOD
             score], indicating a one-LOD support interval spanning 6 Mb.
             To further investigate this finding and to identify
             potential candidate LOAD risk genes for follow-up analysis,
             we analyzed 99 single nucleotide polymorphisms in this
             region, for the overall sample, the autopsy-confirmed
             subset, and the optimal subset, respectively, for
             comparison. We saw no significant association (p<0.01) in
             the overall sample. In the autopsy-confirmed subset, the
             best finding was obtained in the activation transcription
             factor 7 (ATF7) gene (single-locus association, p=0.002;
             haplotype association global, p=0.007). In the optimal
             subset, the best finding was obtained in the hypothetical
             protein FLJ20436 (FLJ20436) gene (single-locus association,
             p=0.0026). These results suggest that subset and covariate
             analyses may be one approach to help identify novel
             susceptibility genes on chromosome 12q for
             LOAD.},
   Doi = {10.1007/s10048-006-0047-z},
   Key = {fds277066}
}

@article{fds277099,
   Author = {Akomolafe, A and Lunetta, KL and Erlich, PM and Cupples, LA and Baldwin,
             CT and Huyck, M and Green, RC and Farrer, LA and MIRAGE Study
             Group},
   Title = {Genetic association between endothelial nitric oxide
             synthase and Alzheimer disease.},
   Journal = {Clin Genet},
   Volume = {70},
   Number = {1},
   Pages = {49-56},
   Year = {2006},
   Month = {July},
   ISSN = {0009-9163},
   url = {http://dx.doi.org/10.1111/j.1399-0004.2006.00638.x},
   Abstract = {Evidence suggests that vascular and inflammatory factors may
             be important in the etiology of Alzheimer disease (AD). The
             Glu/Glu genotype at the Glu298Asp variant of the endothelial
             nitric oxide synthase (NOS3) gene has been tested for
             association with AD in several Caucasian and Asian
             populations, with conflicting results. We tested the
             Glu298Asp variant for association in African American and
             Caucasian AD patients, unaffected siblings, and unrelated
             controls from the MIRAGE Study. To explore whether the
             inconsistent results in previous studies might be due to
             linkage disequilibrium with a polymorphism or haplotype not
             previously tested, we genotyped 10 additional NOS3 single
             nucleotide polymorphisms (SNPs) spanning 25.3 kb. Finally,
             we compiled results of previous studies of Glu298Asp using
             meta-analysis, to determine whether the aggregate studies
             support an association between Glu298Asp and AD. We found
             that the Glu298 allele was associated with higher risk of AD
             in the MIRAGE African American (p = 0.002) but not Caucasian
             (p = 0.9) groups. None of the additional SNPs were
             associated with AD in the Caucasians, whereas two showed
             evidence for association in the African Americans. The
             meta-analysis showed a small effect of the Glu298Asp GG
             genotype on AD risk across all studies (summary odds ratio =
             1.15, 95% confidence interval: 0.97-1.35) and significant
             heterogeneity of this association among studies (p =
             0.02).},
   Doi = {10.1111/j.1399-0004.2006.00638.x},
   Key = {fds277099}
}

@article{fds277011,
   Author = {Lyketsos, CG and Toone, L and Tschanz, J and Corcoran, C and Norton, M and Zandi, P and Munger, R and Breitner, JCS and Welsh-Bohmer, K and Cache
             County Study Group},
   Title = {A population-based study of the association between coronary
             artery bypass graft surgery (CABG) and cognitive decline:
             the Cache County study.},
   Journal = {Int J Geriatr Psychiatry},
   Volume = {21},
   Number = {6},
   Pages = {509-518},
   Year = {2006},
   Month = {June},
   ISSN = {0885-6230},
   url = {http://dx.doi.org/10.1002/gps.1502},
   Abstract = {BACKGROUND: The relationship between coronary artery bypass
             graft (CABG) surgery and cognitive decline remains
             uncertain, in particular with regard to whether there is
             delayed cognitive decline associated with this procedure.
             METHODS: This was a population-based cohort study involving
             participants in the Cache County Study of Memory Health and
             Aging. At baseline the study enrolled 5,092 persons age 65
             and older and followed them up three years later and again
             four years after that. Individuals who reported having
             undergone CABG surgery at study baseline or had this surgery
             in between follow-up waves were compared to individuals who
             never reported having the surgery. The main outcome measure
             was the Modified Mini Mental State (3MS). Multilevel models
             were used to examine the relationship between CABG surgery
             and cognitive decline over time. RESULTS: Study participants
             who had CABG surgery evidenced 0.95 points of greater
             decline relative to baseline on the 3MS at the first
             follow-up interview after CABG, and an average of 1.9 points
             of greater decline at the second follow-up interview, than
             those without CABG (t = -2.51, df = 2,316, p = 0.0121),
             after adjusting for several covariates, including number of
             vascular conditions. This decline was restricted to
             individuals who were more than five years past the procedure
             and was not evident in the early years after the surgery.
             CONCLUSIONS: CABG surgery is associated with accelerated
             cognitive decline more than five years after the procedure
             in a long-lived population. This decline is small and its
             clinical significance is uncertain. We could not find an
             association between CABG and decline in the first five
             post-operative years.},
   Doi = {10.1002/gps.1502},
   Key = {fds277011}
}

@article{fds276977,
   Author = {Norton, MC and Skoog, I and Franklin, LM and Corcoran, C and Tschanz,
             JT and Zandi, PP and Breitner, JCS and Welsh-Bohmer, KA and Steffens,
             DC and Cache County Investigators},
   Title = {Gender differences in the association between religious
             involvement and depression: the Cache County (Utah)
             study.},
   Journal = {J Gerontol B Psychol Sci Soc Sci},
   Volume = {61},
   Number = {3},
   Pages = {P129-P136},
   Year = {2006},
   Month = {May},
   ISSN = {1079-5014},
   url = {http://www.ncbi.nlm.nih.gov/pubmed/16670181},
   Abstract = {We examined the relation between religious involvement,
             membership in the Church of Jesus Christ of Latter-Day
             Saints, and major depression in a population-based study of
             aging and dementia in Cache County, Utah. Participants
             included 4,468 nondemented individuals between the ages of
             65 and 100 years who were interviewed in person. In logistic
             regression models adjusting for demographic and health
             variables, frequent church attendance was associated with a
             reduced prevalence of depression in women but increased
             prevalence in men. Social role loss and the potential impact
             of organizational power differential by sex are discussed.
             Though causality cannot be determined here, these findings
             suggest that the association between religious involvement
             and depression may differ substantially between men and
             women.},
   Doi = {10.1093/geronb/61.3.p129},
   Key = {fds276977}
}

@article{fds277033,
   Author = {Khachaturian, AS and Zandi, PP and Lyketsos, CG and Hayden, KM and Skoog, I and Norton, MC and Tschanz, JT and Mayer, LS and Welsh-Bohmer,
             KA and Breitner, JCS},
   Title = {Antihypertensive medication use and incident Alzheimer
             disease: the Cache County Study.},
   Journal = {Arch Neurol},
   Volume = {63},
   Number = {5},
   Pages = {686-692},
   Year = {2006},
   Month = {May},
   ISSN = {0003-9942},
   url = {http://www.ncbi.nlm.nih.gov/pubmed/16533956},
   Abstract = {BACKGROUND: Recent reports suggest that antihypertensive
             (AH) medications may reduce the risk of dementing illnesses.
             OBJECTIVES: To examine the relationship of AH medication use
             with incidence of Alzheimer disease (AD) among the elderly
             population (aged 65 years and older) of Cache County, Utah,
             and to examine whether the relationship varies with
             different classes of AH medications. METHODS: After an
             initial (wave 1) multistage assessment (1995 through 1997)
             to identify prevalent cases of dementia, we used similar
             methods 3 years later (wave 2) to identify 104 incident
             cases of AD among the 3308 survivors. At the baseline
             assessment, we obtained a detailed drug inventory from the
             study participants. We carried out discrete time survival
             analyses to examine the association between the use of AH
             medications (including angiotensin converting enzyme
             inhibitors, beta-blockers, calcium channel blockers, and
             diuretics) at baseline with subsequent risk of AD. RESULTS:
             Use of any AH medication at baseline was associated with
             lower incidence of AD (adjusted hazard ratio, 0.64; 95%
             confidence interval, 0.41-0.98). Examination of medication
             subclasses showed that use of diuretics (adjusted hazard
             ratio, 0.57; 95% confidence interval, 0.33-0.94), and
             specifically potassium-sparing diuretics (adjusted hazard
             ratio, 0.26; 95% confidence interval, 0.08-0.64), was
             associated with the greatest reduction in risk of AD.
             Corresponding analysis with a fully examined subsample
             controlling for blood pressure measurements did not
             substantially change our findings. CONCLUSIONS: These data
             suggest that AH medications, and specifically
             potassium-sparing diuretics, are associated with reduced
             incidence of AD. Because the latter association is a new
             finding, it requires confirmation in further
             study.},
   Doi = {10.1001/archneur.63.5.noc60013},
   Key = {fds277033}
}

@article{fds277064,
   Author = {Slifer, MA and Martin, ER and Bronson, PG and Browning-Large, C and Doraiswamy, PM and Welsh-Bohmer, KA and Gilbert, JR and Haines, JL and Pericak-Vance, MA},
   Title = {Lack of association between UBQLN1 and Alzheimer
             disease.},
   Journal = {Am J Med Genet B Neuropsychiatr Genet},
   Volume = {141B},
   Number = {3},
   Pages = {208-213},
   Year = {2006},
   Month = {April},
   ISSN = {1552-4841},
   url = {http://www.ncbi.nlm.nih.gov/pubmed/16526030},
   Abstract = {Alzheimer disease (AD) is heterogeneous and complex with a
             strong genetic diathesis. It is the most common cause of
             dementia affecting the elderly. Linkage studies [Kehoe et
             al., 1999; Hum Mol Genet 8: 237-245]; [Pericak-Vance et al.,
             2000; Exp Gerontol 35: 1343-1352]; [Myers et al., 2002; Am J
             Med Genet 114: 235-244]; [Blacker et al., 2003; Hum Mol
             Genet 12: 23-32] identified chromosome 9q as a region
             containing a possible AD candidate gene. Functional protein
             studies [Mah et al., 2000; J Cell Biol 151: 847-862]; [Ko et
             al., 2002; J Biol Chem 277: 35386-35392] identified the
             UBQLN1 gene on chromosome 9q that encodes ubiquilin as a
             likely candidate for a role in late-onset AD pathogenesis. A
             recent family-based study by [Bertram et al., 2005; N Engl J
             352: 884-894] reported genetic association and expression
             evidence for a putative AD risk allele of an intronic single
             nucleotide polymorphism (SNP) within the UBQLN1 gene. In
             this study, we comprehensively assessed whether any of seven
             polymorphisms located across the UBQLN1 gene are associated
             with AD in another large family-based data set and an
             independent case-control data set. We found no significant
             association of AD risk with any of the seven SNPs genotyped
             (including those SNPs previously reported by Bertram et al.)
             in either the family-based or case-control data set.
             Age-specific analyses and analyses conditional on
             Apolipoprotein E (ApoE) genotype and sex also revealed no
             significant associations to AD risk in either data set.
             Using age at onset (AAO) as a quantitative trait revealed a
             modest age modifying association; however, the results were
             inconsistent between the data sets. Our results suggest that
             UBQLN1 variants do not increase risk for AD in these
             data.},
   Doi = {10.1002/ajmg.b.30298},
   Key = {fds277064}
}

@article{fds371181,
   Author = {Fotuhi, M and Zandi, PP and Hayden, KM and Khachaturian, AS and Wengreen, H and Munger, R and Norton, MC and Tschanz, JT and Lyketsos,
             CG and Breitner, JCS and Welsh-Bohmer, KA},
   Title = {Better cognitive performance in elderly taking vitamins E
             and C in combination with NSAIDs: The Cache County
             study},
   Journal = {NEUROLOGY},
   Volume = {66},
   Number = {5},
   Pages = {A217-A217},
   Publisher = {LIPPINCOTT WILLIAMS & WILKINS},
   Year = {2006},
   Month = {March},
   Key = {fds371181}
}

@article{fds276949,
   Author = {Hahs, DW and McCauley, JL and Crunk, AE and McFarland, LL and Gaskell,
             PC and Jiang, L and Slifer, SH and Vance, JM and Scott, WK and Welsh-Bohmer, KA and Johnson, SR and Jackson, CE and Pericak-Vance,
             MA and Haines, JL},
   Title = {A genome-wide linkage analysis of dementia in the
             Amish.},
   Journal = {Am J Med Genet B Neuropsychiatr Genet},
   Volume = {141B},
   Number = {2},
   Pages = {160-166},
   Year = {2006},
   Month = {March},
   ISSN = {1552-4841},
   url = {http://www.ncbi.nlm.nih.gov/pubmed/16389594},
   Abstract = {Susceptibility genes for Alzheimer's disease are proving to
             be highly challenging to detect and verify. Population
             heterogeneity may be a significant confounding factor
             contributing to this difficulty. To increase the power for
             disease susceptibility gene detection, we conducted a
             genome-wide genetic linkage screen using individuals from
             the relatively isolated, genetically homogeneous, Amish
             population. Our genome linkage analysis used a
             407-microsatellite-marker map (average density 7 cM) to
             search for autosomal genes linked to dementia in five Amish
             families from four Midwestern U.S. counties. Our highest
             two-point lod score (3.01) was observed at marker D4S1548 on
             chromosome 4q31. Five other regions (10q22, 3q28, 11p13,
             4q28, 19p13) also demonstrated suggestive linkage with
             markers having two-point lod scores >2.0. While two of these
             regions are novel (4q31 and 11p13), the other regions lie
             close to regions identified in previous genome scans in
             other populations. Our results identify regions of the
             genome that may harbor genes involved in a subset of
             dementia patients, in particular the North American Amish
             community.},
   Doi = {10.1002/ajmg.b.30257},
   Key = {fds276949}
}

@article{fds276948,
   Author = {McCauley, JL and Hahs, DW and Jiang, L and Scott, WK and Welsh-Bohmer,
             KA and Jackson, CE and Vance, JM and Pericak-Vance, MA and Haines,
             JL},
   Title = {Combinatorial Mismatch Scan (CMS) for loci associated with
             dementia in the Amish.},
   Journal = {BMC Med Genet},
   Volume = {7},
   Pages = {19},
   Year = {2006},
   Month = {March},
   url = {http://www.ncbi.nlm.nih.gov/pubmed/16515697},
   Abstract = {BACKGROUND: Population heterogeneity may be a significant
             confounding factor hampering detection and verification of
             late onset Alzheimer's disease (LOAD) susceptibility genes.
             The Amish communities located in Indiana and Ohio are
             relatively isolated populations that may have increased
             power to detect disease susceptibility genes. METHODS: We
             recently performed a genome scan of dementia in this
             population that detected several potential loci. However,
             analyses of these data are complicated by the highly
             consanguineous nature of these Amish pedigrees. Therefore we
             applied the Combinatorial Mismatch Scanning (CMS) method
             that compares identity by state (IBS) (under the presumption
             of identity by descent (IBD)) sharing in distantly related
             individuals from such populations where standard linkage and
             association analyses are difficult to implement. CMS
             compares allele sharing between individuals in affected and
             unaffected groups from founder populations. Comparisons
             between cases and controls were done using two Fisher's
             exact tests, one testing for excess in IBS allele frequency
             and the other testing for excess in IBS genotype frequency
             for 407 microsatellite markers. RESULTS: In all, 13 dementia
             cases and 14 normal controls were identified who were not
             related at least through the grandparental generation. The
             examination of allele frequencies identified 24 markers (6%)
             nominally (p < or = 0.05) associated with dementia; the most
             interesting (empiric p < or = 0.005) markers were D3S1262,
             D5S211, and D19S1165. The examination of genotype
             frequencies identified 21 markers (5%) nominally (p < or =
             0.05) associated with dementia; the most significant markers
             were both located on chromosome 5 (D5S1480 and D5S211).
             Notably, one of these markers (D5S211) demonstrated
             differences (empiric p < or = 0.005) under both tests.
             CONCLUSION: Our results provide the initial groundwork for
             identifying genes involved in late-onset Alzheimer's disease
             within the Amish community. Genes identified within this
             isolated population will likely play a role in a subset of
             late-onset AD cases across more general populations. Regions
             highlighted by markers demonstrating suggestive allelic
             and/or genotypic differences will be the focus of more
             detailed examination to characterize their involvement in
             dementia.},
   Doi = {10.1186/1471-2350-7-19},
   Key = {fds276948}
}

@article{fds276976,
   Author = {Norton, MC and Skoog, I and Toone, L and Corcoran, C and Tschanz, JT and Lisota, RD and Hart, AD and Zandi, PP and Breitner, JCS and Welsh-Bohmer, KA and Steffens, DC and Cache County
             Investigators},
   Title = {Three-year incidence of first-onset depressive syndrome in a
             population sample of older adults: the Cache County
             study.},
   Journal = {Am J Geriatr Psychiatry},
   Volume = {14},
   Number = {3},
   Pages = {237-245},
   Year = {2006},
   Month = {March},
   ISSN = {1064-7481},
   url = {http://www.ncbi.nlm.nih.gov/pubmed/16505128},
   Abstract = {OBJECTIVE: Estimates of incidence of late-life depression
             vary greatly with few studies excluding demented cases
             through in-depth evaluation and most studies failing to
             control for the effect of mortality and interval treatment.
             In a large population-based study, the authors examined the
             effect on incidence of first-onset depressive syndrome to
             determine whether any gender or age differences in incidence
             are attenuated with inclusion of these additional measures.
             METHOD: Incidence rates of depressive syndrome per 1,000
             person-years are presented for 2,877 nondemented elderly
             (ages 65 to 100 years) residents of Cache County, Utah.
             Cases are identified by direct interview methods, by
             inference from prescription antidepressant medicine use, and
             by postmortem informant interview for decedents. RESULTS:
             In-person interviews yielded incidence rates of first-onset
             depressive disorder (any type) of 13.09 for men and 19.44
             for women. Inclusion of antidepressant users increased these
             figures to 15.55 for men and 23.30 for women. Addition of
             postmortem interview data yielded rates of 20.66 for men and
             26.29 for women. Individuals with no history of depression
             had rates for major depression of 7.88 for men and 8.75 for
             women; minor depression rates were 19.23 for men and 24.46
             for women (p = 0.691; effect for minor depression p
             <0.0001). Age did not predict incidence. CONCLUSIONS:
             Incidence of first-onset major depression varies with data
             source and prior lifetime history of depression. Gender
             effects apparent in interview data are attenuated when
             postmortem information and pharmacotherapy were
             considered.},
   Doi = {10.1097/01.JGP.0000196626.34881.42},
   Key = {fds276976}
}

@article{fds276975,
   Author = {Steffens, DC and Otey, E and Alexopoulos, GS and Butters, MA and Cuthbert, B and Ganguli, M and Geda, YE and Hendrie, HC and Krishnan,
             RR and Kumar, A and Lopez, OL and Lyketsos, CG and Mast, BT and Morris, JC and Norton, MC and Peavy, GM and Petersen, RC and Reynolds, CF and Salloway,
             S and Welsh-Bohmer, KA and Yesavage, J},
   Title = {Perspectives on depression, mild cognitive impairment, and
             cognitive decline.},
   Journal = {Arch Gen Psychiatry},
   Volume = {63},
   Number = {2},
   Pages = {130-138},
   Year = {2006},
   Month = {February},
   ISSN = {0003-990X},
   url = {http://www.ncbi.nlm.nih.gov/pubmed/16461855},
   Abstract = {CONTEXT: The public health implications of depression and
             cognitive impairment in late life are enormous. Cognitive
             impairment and late-life depression are associated with
             increased risk for subsequent dementia; however,
             investigations of these phenomena appear to be proceeding
             along separate tracks. OBJECTIVES AND DATA SOURCE: The
             National Institute of Mental Health organized the conference
             "Perspectives on Depression, Mild Cognitive Impairment, and
             Cognitive Decline" to consider how the varied perspectives
             might be better integrated to examine the associations among
             depression, mild cognitive impairment, and cognitive decline
             and to illuminate the common or distinct mechanisms involved
             in these associations. DATA SYNTHESIS: The following 2 broad
             questions were addressed: (1) What gaps in our knowledge
             have the greatest public health significance? (2) Can we
             more efficiently use our research dollars and participant
             resources to fill these gaps? Meeting participants included
             grantees from the National Institute of Mental Health and
             the National Institute on Aging and program staff from the
             National Institute of Mental Health, the National Institute
             on Aging, and the National Institute of Neurological
             Disorders and Stroke. CONCLUSIONS: One of the most important
             recommendations to emerge from the meeting discussions is
             for increased collaboration among clinical and
             epidemiological investigators whose work focuses in the area
             of depression with those working primarily in the area of
             memory disorders. Directions for future research were
             identified.},
   Doi = {10.1001/archpsyc.63.2.130},
   Key = {fds276975}
}

@article{fds277032,
   Author = {Østbye, T and Krause, KM and Norton, MC and Tschanz, J and Sanders, L and Hayden, K and Pieper, C and Welsh-Bohmer, KA and Cache County
             Investigators},
   Title = {Ten dimensions of health and their relationships with
             overall self-reported health and survival in a predominately
             religiously active elderly population: the cache county
             memory study.},
   Journal = {J Am Geriatr Soc},
   Volume = {54},
   Number = {2},
   Pages = {199-209},
   Year = {2006},
   Month = {February},
   ISSN = {0002-8614},
   url = {http://www.ncbi.nlm.nih.gov/pubmed/16460369},
   Abstract = {OBJECTIVES: To document the extent of healthy aging along 10
             different dimensions in a population known for its
             longevity. DESIGN: A cohort study with baseline measures of
             overall self-reported health and health along 10 specific
             dimensions; analyses investigated the 10 dimensions as
             predictors of self-reported health and 10-year mortality.
             SETTING: Cache County, Utah, which is among the areas with
             the highest conditional life expectancy at age 65 in the
             United States. PARTICIPANTS: Inhabitants of Cache County
             aged 65 and older (January 1, 1995). MEASUREMENTS:
             Self-reported overall health and 10 specific dimensions of
             healthy aging: independent living, vision, hearing,
             activities of daily living, instrumental activities of daily
             living, absence of physical illness, cognition, healthy
             mood, social support and participation, and religious
             participation and spirituality. RESULTS: This elderly
             population was healthy overall. With few exceptions, 80% to
             90% of persons aged 65 to 75 were healthy according to each
             measure used. Prevalence of excellent and good self-reported
             health decreased with age, to approximately 60% in those
             aged 85 and older. Even in the oldest old, the majority of
             respondents were independent in activities of daily living.
             Although vision, hearing, and mood were significant
             predictors of overall self-reported health in the final
             models, age, sex, and cognition were significant only in the
             final survival models. CONCLUSION: This population has a
             high prevalence of most factors representing healthy aging.
             The predictors of overall self-reported health are distinct
             from the predictors of survival in this age group and, being
             potentially modifiable, are amenable to clinical and public
             health efforts.},
   Doi = {10.1111/j.1532-5415.2005.00583.x},
   Key = {fds277032}
}

@article{fds277081,
   Author = {Plassman, BL and Khachaturian, AS and Townsend, JJ and Ball, MJ and Steffens, DC and Leslie, CE and Tschanz, JT and Norton, MC and Burke,
             JR and Welsh-Bohmer, KA and Hulette, CM and Nixon, RR and Tyrey, M and Breitner, JCS},
   Title = {Comparison of clinical and neuropathologic diagnoses of
             Alzheimer's disease in 3 epidemiologic samples.},
   Journal = {Alzheimers Dement},
   Volume = {2},
   Number = {1},
   Pages = {2-11},
   Year = {2006},
   Month = {January},
   url = {http://www.ncbi.nlm.nih.gov/pubmed/19595851},
   Abstract = {BACKGROUND: Studies of dementia in populations avoid many of
             the selection biases in clinical samples but require special
             evaluation and diagnostic methods to obtain high
             participation rates. To address this issue, we developed a
             unique in-home dementia assessment. We assessed validity of
             these assessments using neuropathologic confirmation of the
             clinical diagnosis in 3 epidemiologic samples. METHODS:
             Subjects were 175 participants in 3 ongoing studies of
             dementia. Two were population based and identified dementia
             by cognitive screening. The third study sought volunteers
             via advertisements. Dementia evaluations were then conducted
             at the participants' residences by specially trained nurses
             and psychometricians. Evaluation results were interpreted,
             and preliminary diagnoses were assigned by a
             geropsychiatrist or neurologist and a psychologist. Final
             diagnoses were assigned by a consensus panel of
             neurologists, geropsychiatrists, and psychologists. We
             compared the clinical diagnoses with the gold-standard
             neuropathologic diagnoses for those participants who
             subsequently underwent autopsy. RESULTS: Among the demented,
             the sensitivity of a clinical diagnosis of probable or
             possible Alzheimer's disease (AD) was 93% across the 3
             studies. The rate of overall diagnostic agreement was 81%.
             Measures of agreement did not differ meaningfully across
             varying levels of dementia severity. CONCLUSIONS: Rates of
             neuropathologic confirmation for clinical AD diagnoses in
             these studies were similar to those reported from
             clinic-based samples. These results support the validity of
             clinical diagnoses of AD from a structured in-home
             assessment of community dwelling and institutionalized
             individuals using relatively economical methods of dementia
             screening and assessment.},
   Doi = {10.1016/j.jalz.2005.11.001},
   Key = {fds277081}
}

@article{fds277098,
   Author = {Erlich, PM and Lunetta, KL and Cupples, LA and Huyck, M and Green, RC and Baldwin, CT and Farrer, LA and MIRAGE Study Group},
   Title = {Polymorphisms in the PON gene cluster are associated with
             Alzheimer disease.},
   Journal = {Hum Mol Genet},
   Volume = {15},
   Number = {1},
   Pages = {77-85},
   Year = {2006},
   Month = {January},
   ISSN = {0964-6906},
   url = {http://dx.doi.org/10.1093/hmg/ddi428},
   Abstract = {Paraoxonase is an arylesterase enzyme that is expressed in
             the liver and found in the circulation in association with
             apoA1 and the high-density lipoprotein, and prevents the
             accumulation of oxidized lipids in low-density lipoproteins
             in vitro. Common polymorphisms in genes encoding paraoxonase
             are established risk factors in a variety of vascular
             disorders including coronary artery disease and carotid
             artery stenosis, but their association with Alzheimer
             disease (AD) is controversial. We tested the association of
             29 SNPs in PON1, PON2 and PON3 with AD in 730 Caucasian and
             467 African American participants of the MIRAGE Study, an
             ongoing multi-center family-based genetic epidemiology study
             of AD. Eight SNPs were associated with AD in the African
             American families (0.0001< or =P< or =0.04) and two SNPs
             were associated with AD in Caucasian families (0.01< or =P<
             or =0.04). Of note, the pattern of association for the PON1
             promoter SNP -161[C/T] was the same in both ethnic groups
             (P=0.006). Haplotype analysis using sliding windows revealed
             11 contiguous SNP combinations spanning the three PON genes
             with significant global test scores (0.006< or =P< or =0.04)
             in the two ethnic groups combined. The most significantly
             associated haplotype comprised SNPs in the region spanning
             the -161[C/T] SNP (P=0.00009). Our results demonstrate
             association between AD and variants in the PON gene cluster
             in Caucasians and African Americans.},
   Doi = {10.1093/hmg/ddi428},
   Key = {fds277098}
}

@article{fds276979,
   Author = {Sair, HI and Welsh-Bohmer, KA and Wagner, HR and Steffens,
             DC},
   Title = {Ascending digits task as a measure of executive function in
             geriatric depression.},
   Journal = {J Neuropsychiatry Clin Neurosci},
   Volume = {18},
   Number = {1},
   Pages = {117-120},
   Year = {2006},
   ISSN = {0895-0172},
   url = {http://www.ncbi.nlm.nih.gov/pubmed/16525080},
   Abstract = {The authors hypothesized that older depressed patients would
             perform more poorly on the Ascending Digits Task (ADT) when
             matched against a nondepressed elderly comparison group. In
             a novel measure, the ADT, 129 older depressives scored more
             poorly than 129 comparison subjects in bivariate analyses
             and models controlling for demographic variables. The ADT
             may be a good measure of executive function in older
             adults.},
   Doi = {10.1176/jnp.18.1.117},
   Key = {fds276979}
}

@article{fds277030,
   Author = {Hayden, KM and Zandi, PP and Lyketsos, CG and Khachaturian, AS and Bastian, LA and Charoonruk, G and Tschanz, JT and Norton, MC and Pieper,
             CF and Munger, RG and Breitner, JCS and Welsh-Bohmer, KA and Cache
             County Investigators},
   Title = {Vascular risk factors for incident Alzheimer disease and
             vascular dementia: the Cache County study.},
   Journal = {Alzheimer Dis Assoc Disord},
   Volume = {20},
   Number = {2},
   Pages = {93-100},
   Year = {2006},
   ISSN = {0893-0341},
   url = {http://www.ncbi.nlm.nih.gov/pubmed/16772744},
   Abstract = {Vascular risk factors for Alzheimer disease (AD) and
             vascular dementia (VaD) have been evaluated; however, few
             studies have compared risks by dementia subtypes and sex. We
             evaluated relationships between cardiovascular risk factors
             (hypertension, high cholesterol, diabetes mellitus, and
             obesity), events (stroke, coronary artery bypass graft
             surgery, and myocardial infarction), and subsequent risk of
             AD and VaD by sex in a community-based cohort of 3264 Cache
             County residents aged 65 or older. Cardiovascular history
             was ascertained by self-report or proxy-report in detailed
             interviews. AD and VaD were diagnosed using standard
             criteria. Estimates from discrete-time survival models
             showed no association between self-reported history of
             hypertension and high cholesterol and AD after adjustments.
             Hypertension increased the risk of VaD [adjusted hazard
             ratio (aHR) 2.42, 95% confidence interval (CI) 0.95-7.44].
             Obesity increased the risk of AD in females (aHR 2.23, 95%
             CI 1.09-4.30) but not males. Diabetes increased the risk of
             VaD in females after adjustments (aHR 3.33, 95% CI
             1.03-9.78) but not males. The risk of VaD after stroke was
             increased in females (aHR 16.90, 95% CI 5.58-49.03) and
             males (aHR 10.95, 95% CI 2.48-44.78). The results indicate
             that vascular factors increase risks for AD and VaD
             differentially by sex. Future studies should focus on
             specific causal pathways for each of these factors with
             regard to sex to determine if sex differences in the
             prevalence of vascular factors have an influence on sex
             differences in dementia risk.},
   Doi = {10.1097/01.wad.0000213814.43047.86},
   Key = {fds277030}
}

@article{fds371182,
   Author = {Lin, PI and Martin, ER and Bronson, PG and Browning-Large, CA and Small,
             GW and Schmechel, DE and Welsh-Bohmer, KA and Haines, JL and Gilbert,
             JR and Pericak-Vance, MA},
   Title = {Exploring the effect of interactions of GAPD genes on
             late-onset al.zheimer disease},
   Journal = {GENETIC EPIDEMIOLOGY},
   Volume = {29},
   Number = {3},
   Pages = {264-265},
   Publisher = {WILEY-LISS},
   Year = {2005},
   Month = {November},
   Key = {fds371182}
}

@article{fds277145,
   Author = {van der Walt, JM and Scott, WK and Slifer, S and Gaskell, PC and Martin,
             ER and Welsh-Bohmer, K and Creason, M and Crunk, A and Fuzzell, D and McFarland, L and Kroner, CC and Jackson, CE and Haines, JL and Pericak-Vance, MA},
   Title = {Maternal lineages and Alzheimer disease risk in the Old
             Order Amish.},
   Journal = {Hum Genet},
   Volume = {118},
   Number = {1},
   Pages = {115-122},
   Year = {2005},
   Month = {October},
   ISSN = {0340-6717},
   url = {http://www.ncbi.nlm.nih.gov/pubmed/16078048},
   Abstract = {Old Order Amish, founded by a small number of Swiss
             immigrants, exist in culturally isolated communities across
             rural North America. The consequences of genetic isolation
             and inbreeding within this group are evident by increased
             frequencies of many monogenic diseases and several complex
             disorders. Conversely, the prevalence of Alzheimer disease
             (AD), the most common form of dementia, is lower in the
             Amish than in the general American population. Since
             mitochondrial dysfunction has been proposed as an underlying
             cause of AD and a specific haplogroup was found to affect AD
             susceptibility in Caucasians, we investigated whether
             inherited mitochondrial haplogroups affect risk of
             developing AD dementia in Ohio and Indiana Amish
             communities. Ninety-five independent matrilines were
             observed across six large pedigrees and three small
             pedigrees then classified into seven major European
             haplogroups. Haplogroup T is the most frequent haplogroup
             represented overall in these maternal lines (35.4%) while
             observed in only 10.6% in outbred American and European
             populations. Furthermore, haplogroups J and K are less
             frequent (1.0%) than in the outbred data set (9.4-11.2%).
             Affected case matrilines and unaffected control lines were
             chosen from pedigrees to test whether specific haplogroups
             and their defining SNPs confer risk of AD. We did not
             observe frequency differences between AD cases compared to
             controls overall or when stratified by sex. Therefore, we
             suggest that the genetic effect responsible for AD dementia
             in the affected Amish pedigrees is unlikely to be of
             mitochondrial origin and may be caused by nuclear genetic
             factors.},
   Doi = {10.1007/s00439-005-0032-x},
   Key = {fds277145}
}

@article{fds277146,
   Author = {Lyketsos, CG and Toone, L and Tschanz, J and Rabins, PV and Steinberg,
             M and Onyike, CU and Corcoran, C and Norton, M and Zandi, P and Breitner,
             JCS and Welsh-Bohmer, K and Anthony, J and Østbye, T and Bigler, E and Pieper, C and Burke, J and Plassman, B and Green, RC and Steffens, DC and Klein, L and Leslie, C and Townsend, JJ and Wyse, BW and Munger, R and Williams, M and Cache County Study Group},
   Title = {Population-based study of medical comorbidity in early
             dementia and "cognitive impairment, no dementia (CIND)":
             association with functional and cognitive impairment: The
             Cache County Study.},
   Journal = {Am J Geriatr Psychiatry},
   Volume = {13},
   Number = {8},
   Pages = {656-664},
   Year = {2005},
   Month = {August},
   ISSN = {1064-7481},
   url = {http://www.ncbi.nlm.nih.gov/pubmed/16085781},
   Abstract = {OBJECTIVE: Authors investigated medical comorbidity in
             persons with dementia and "Cognitive Impairment, No
             Dementia" (CIND). METHODS: The Cache County Study is an
             ongoing population-based study of the epidemiology of
             dementia, the risk factors for conversion from CIND to
             dementia, and the progression of dementia. As part of the
             study's first incidence wave, persons with dementia (N=149),
             CIND (N=225), or without cognitive impairment (N=321) were
             identified and studied. Participants received comprehensive
             clinical evaluations and were rated on the General Medical
             Health Rating (GMHR), a global measure of seriousness of
             medical comorbidity. Participants and informants also
             completed the Mini-Mental State Exam and provided
             self-report information about comorbid medical conditions
             and functioning in activities of daily living. RESULTS:
             There were few differences in number or type of comorbid
             medical conditions between persons with CIND and dementia,
             but persons with dementia were prescribed more medications.
             Stroke was more common in dementia participants, but other
             illnesses common in old age were not significantly different
             across cognitive groups. Medical comorbidity was more
             serious in both dementia and CIND, such that both groups
             were less likely to have "little to no" comorbidity.
             Seriousness of medical comorbidity was significantly
             associated with worse day-to-day functioning and cognition.
             CONCLUSIONS: Persons with CIND and dementia have more
             serious medical comorbidity than comparable persons without
             cognitive impairment. This comorbidity may play a role in
             the progression of CIND and dementia. Future studies should
             investigate the role of medical comorbidity and its
             treatment on dementia onset or progression, as well as the
             mechanisms mediating its neuropathologic
             effects.},
   Doi = {10.1176/appi.ajgp.13.8.656},
   Key = {fds277146}
}

@article{fds277031,
   Author = {Hayden, KM and Warren, LH and Pieper, CF and Østbye, T and Tschanz, JT and Norton, MC and Breitner, JCS and Welsh-Bohmer,
             KA},
   Title = {Identification of VaD and AD prodromes: the Cache County
             Study.},
   Journal = {Alzheimers Dement},
   Volume = {1},
   Number = {1},
   Pages = {19-29},
   Year = {2005},
   Month = {July},
   url = {http://www.ncbi.nlm.nih.gov/pubmed/19595812},
   Abstract = {BACKGROUND: It is unclear whether vascular dementia (VaD)
             has a cognitive prodrome, akin to the mild cognitive
             impairment (MCI) prodrome to Alzheimer's dementia (AD). To
             evaluate whether VaD has a cognitive prodrome, and if it can
             be differentiated from prodromal AD, we examined
             neuropsychological test performance of participants in a
             nested case-control study within a population-based cohort
             aged 65 or older. METHODS: Participants (n = 485) were
             identified from the Cache County Study, a large
             population-based study of aging and dementia. After an
             average of 3 years of follow-up, a total of 62 incident
             dementia cases were identified (14 VaD, 48 AD). We
             identified a number of neuropsychological tests (executive
             and memory) that discriminated between diagnosed VaD and AD
             cases. Multivariate analyses sought to differentiate between
             these same groups 3 years before clinical diagnosis.
             RESULTS: The Consortium to Establish a Registry for
             Alzheimer's Disease Word List Recognition Test correct
             recognition of foils (mean difference, 1.25; 95% confidence
             interval [CI], 0.42 to 2.07; p < 0.01), Logical Memory I
             (mean difference, 7.16; 95% CI, 0.78 to 13.55, p < 0.05),
             Logical Memory II delayed recall (mean difference, 8.67; 95%
             CI, 1.59 to 15.74, p < 0.05), and percent savings (mean
             difference, 51.07; 95% CI, 32.58 to 69.56, p < 0.0001)
             differentiated VaD from AD cases after adjustment for age,
             sex, education, and dementia severity. Three years before
             dementia diagnosis, word list recognition ("no" responses
             mean difference, 1.40; 95% CI, 0.64 to 2.17; p < 0.001, and
             "yes" responses mean difference, -1.14; 95% CI, -2.14 to
             -0.13; p < 0.03) discriminated between prodromal VaD and AD.
             CONCLUSION: These results suggest that VaD has a prodromal
             syndrome, the cognitive features of which are
             distinguishable from the cognitive prodrome of
             AD.},
   Doi = {10.1016/j.jalz.2005.06.002},
   Key = {fds277031}
}

@article{fds277028,
   Author = {Hayden, KM and Zandi, PP and Lyketsos, CG and Tschanz, JT and Norton,
             MC and Khachaturian, AS and Pieper, CF and Welsh-Bohmer, KA and Breitner, JCS and Cache County Investigators},
   Title = {Apolipoprotein E genotype and mortality: findings from the
             Cache County Study.},
   Journal = {J Am Geriatr Soc},
   Volume = {53},
   Number = {6},
   Pages = {935-942},
   Year = {2005},
   Month = {June},
   ISSN = {0002-8614},
   url = {http://www.ncbi.nlm.nih.gov/pubmed/15935014},
   Abstract = {OBJECTIVES: To evaluate the association between
             apolipoprotein E (apo E) epsilon4 and mortality, the
             population attributable risk for mortality with epsilon4,
             and relative contributions of cardiovascular disease (CVD)
             and Alzheimer's disease (AD). DESIGN: Population-based
             cohort study. SETTING: Community-based. PARTICIPANTS:
             Permanent residents of Cache County, Utah, aged 65 and older
             as of January 1, 1995. MEASUREMENTS: Participants were
             genotyped at the apo E locus using buccal-swab
             deoxyribonucleic acid. Cardiovascular health was ascertained
             using self- or proxy-report interviews at participants'
             residences. AD was diagnosed according to Diagnostic and
             Statistical Manual of Mental Disorders, Third Edition,
             Revised, and National Institute of Neurological and
             Communicative Disorders and Stroke-Alzheimer's Disease and
             Related Disorders criteria. Utah Department of Vital
             Statistics quarterly reports were reviewed to identify
             participants who died. RESULTS: Crude evaluations showed
             nonsignificantly greater risk of death for epsilon2/2
             (hazard ratio (HR)=1.66, 95% confidence interval
             (CI)=0.92-2.76) and epsilon3/4 (HR=1.11, 95% CI=0.97-1.26)
             genotypes and significantly greater risk for epsilon4/4
             (HR=1.48, 95% CI=1.09-1.96). After adjustment for age,
             age(2), sex, and education, risks increased to 1.98 (95%
             CI=1.08-3.35), 1.28 (95% CI=1.12-1.46), and 2.02 (95%
             CI=1.47-2.71), respectively, compared with epsilon3/3
             genotypes. Adjustment for presence of any CVD did not change
             the risk of death for epsilon3/4 and epsilon4/4. Adjustment
             for AD reduced the risk of death for epsilon3/4 (HR=1.13,
             95% CI=0.99-1.30) and epsilon4/4 (HR=1.59, 95%
             CI=1.15-2.14). The population attributable risk of death for
             epsilon3/4 and epsilon4/4 genotypes combined is estimated at
             9.6%. CONCLUSION: These findings suggested that the
             epsilon2/2, epsilon3/4, and epsilon4/4 genotypes have
             greater early mortality risks. Further analyses showed that
             AD partially mediates the association between epsilon3/4,
             epsilon4/4, and death.},
   Doi = {10.1111/j.1532-5415.2005.53301.x},
   Key = {fds277028}
}

@article{fds277119,
   Author = {Ashley-Koch, AE and Shao, Y and Rimmler, JB and Gaskell, PC and Welsh-Bohmer, KA and Jackson, CE and Scott, WK and Haines, JL and Pericak-Vance, MA},
   Title = {An autosomal genomic screen for dementia in an extended
             Amish family.},
   Journal = {Neurosci Lett},
   Volume = {379},
   Number = {3},
   Pages = {199-204},
   Year = {2005},
   Month = {May},
   ISSN = {0304-3940},
   url = {http://www.ncbi.nlm.nih.gov/pubmed/15843063},
   Abstract = {Apolipoprotein E (APOE) is the only universally confirmed
             susceptibility gene for late-onset Alzheimer disease (LOAD),
             although many loci are believed to modulate LOAD risk. The
             genetic homogeneity of isolated populations, such as the
             Amish, potentially provide increased power to identify LOAD
             susceptibility genes. Population homogeneity in these
             special populations may reduce the total number of
             susceptibility genes contributing to the complex disorder,
             thereby increasing the ability to identify any one
             susceptibility gene. Dementia in the Amish is clinically
             indistinguishable from LOAD in the general population.
             Previous studies in the Amish demonstrated a significantly
             decreased frequency of the APOE-4 susceptibility allele, but
             significant familial clustering of dementia [M.A.
             Pericak-Vance, C.C. Johnson, J.B. Rimmler, A.M. Saunders,
             L.C. Robinson, E.G. D'Hondt, C.E. Jackson, J.L. Haines,
             Alzheimer's disease and apolipoprotein E-4 allele in an
             Amish population, Ann. Neurol. 39 (1996) 700-704]. These
             data suggested that a genetic etiology independent of APOE
             may underlie the dementia observed in this population. In
             the present analysis, we focused on a large, multiplex,
             inbred Amish family (24 sampled individuals; 10 of whom are
             affected). We completed a genomic screen to identify novel
             LOAD loci (n=316 genetic markers), using both
             model-dependent "affecteds-only" analysis (dominant and
             recessive) and model-independent affected relative pair
             analysis. Interesting results (lod>1.5 or p<0.01) were
             obtained for markers on eight chromosomes (2q, 5q, 6q, 7p,
             8p, 8q, 11p, 18p, 18q, and 19q). The highest overall score
             was a multipoint lod score of 3.1 on chromosome 11p. Most
             regions we identified were not previously detected by
             genomic screens of outbred populations and may represent
             population-specific susceptibilities to LOAD. These loci are
             currently under further investigation in a study of LOAD
             including additional Amish families.},
   Doi = {10.1016/j.neulet.2004.12.065},
   Key = {fds277119}
}

@article{fds371183,
   Author = {Hulette, CM and Ervin, JF and Edmonds, Y and Stewart, N and Welsh-Bohmer, K and Pieper, CF and Szymanski, MH and Schmechel,
             DE},
   Title = {Cerebrovascular smooth muscle actin is increased in possible
             Azheimer disease brain tissue: A controlled autopsy
             study.},
   Journal = {JOURNAL OF NEUROPATHOLOGY AND EXPERIMENTAL
             NEUROLOGY},
   Volume = {64},
   Number = {5},
   Pages = {467-467},
   Publisher = {LIPPINCOTT WILLIAMS & WILKINS},
   Year = {2005},
   Month = {May},
   Key = {fds371183}
}

@article{fds277144,
   Author = {Zandi, PP and Sparks, DL and Khachaturian, AS and Tschanz, J and Norton,
             M and Steinberg, M and Welsh-Bohmer, KA and Breitner, JCS and Cache
             County Study investigators},
   Title = {Do statins reduce risk of incident dementia and Alzheimer
             disease? The Cache County Study.},
   Journal = {Arch Gen Psychiatry},
   Volume = {62},
   Number = {2},
   Pages = {217-224},
   Year = {2005},
   Month = {February},
   ISSN = {0003-990X},
   url = {http://www.ncbi.nlm.nih.gov/pubmed/15699299},
   Abstract = {BACKGROUND: Prior reports suggest reduced occurrence of
             dementia and Alzheimer disease (AD) in statin users, but, to
             our knowledge, no prospective studies relate statin use and
             dementia incidence. OBJECTIVE: To examine the association of
             statin use with both prevalence and incidence of dementia
             and AD. DESIGN: Cross-sectional studies of prevalence and
             incidence and a prospective study of incidence of dementia
             and AD among 5092 elderly residents (aged 65 years or older)
             of a single county. Participants were assessed at home in
             1995-1997 and again in 1998-2000. A detailed visual
             inventory of medicines, including statins and other
             lipid-lowering agents, was collected at both assessments.
             MAIN OUTCOME MEASURES: Diagnosis of dementia and of AD.
             RESULTS: From 4895 participants with data sufficient to
             determine cognitive status, we identified 355 cases of
             prevalent dementia (200 with AD) at initial assessment.
             Statin use was inversely associated with prevalence of
             dementia (adjusted odds ratio, 0.44; 95% confidence
             interval, 0.17-0.94). Three years later, we identified 185
             cases of incident dementia (104 with AD) among 3308
             survivors at risk. Statin use at baseline did not predict
             incidence of dementia or AD (adjusted hazard ratio for
             dementia, 1.19; 95% confidence interval, 0.53-2.34; adjusted
             hazard ratio for AD, 1.19; 95% confidence interval,
             0.35-2.96), nor did statin use at follow-up (adjusted odds
             ratio for dementia, 1.04; 95% confidence interval,
             0.56-1.81; adjusted odds ratio for AD, 0.85; 95% confidence
             interval, 0.32-1.88). CONCLUSIONS: Although statin use might
             be less frequent in those with prevalent dementia, we found
             no association between statin use and subsequent onset of
             dementia or AD. Further research is warranted before costly
             dementia prevention trials with statins are
             undertaken.},
   Doi = {10.1001/archpsyc.62.2.217},
   Key = {fds277144}
}

@article{fds277079,
   Author = {LaBar, KS and Torpey, DC and Cook, CA and Johnson, SR and Warren, LH and Burke, JR and Welsh-Bohmer, KA},
   Title = {Emotional enhancement of perceptual priming is preserved in
             aging and early-stage Alzheimer's disease.},
   Journal = {Neuropsychologia},
   Volume = {43},
   Number = {12},
   Pages = {1824-1837},
   Year = {2005},
   ISSN = {0028-3932},
   url = {http://www.ncbi.nlm.nih.gov/pubmed/16154458},
   Abstract = {Perceptual priming for emotionally-negative and neutral
             scenes was tested in early-stage Alzheimer's disease (AD)
             patients and healthy younger, middle-aged and older adults.
             In the study phase, participants rated the scenes for their
             arousal properties. In the test phase, studied and novel
             scenes were initially presented subliminally, and the
             exposure duration was gradually increased until a valence
             categorization was made. The difference in exposure duration
             required to categorize novel versus studied items was the
             dependent measure of priming. Aversive content increased the
             magnitude of priming, an effect that was preserved in
             healthy aging and AD. Results from an immediate recognition
             memory test showed that the priming effects could not be
             attributable to enhanced explicit memory for the aversive
             scenes. These findings implicate a dissociation between the
             modulatory effect of emotion across implicit and explicit
             forms of memory in aging and early-stage
             AD.},
   Doi = {10.1016/j.neuropsychologia.2005.01.018},
   Key = {fds277079}
}

@article{fds277080,
   Author = {Tschanz, JT and Treiber, K and Norton, MC and Welsh-Bohmer, KA and Toone, L and Zandi, PP and Szekely, CA and Lyketsos, C and Breitner,
             JCS and Cache County Study Group},
   Title = {A population study of Alzheimer's disease: findings from the
             Cache County Study on Memory, Health, and
             Aging.},
   Journal = {Care Manag J},
   Volume = {6},
   Number = {2},
   Pages = {107-114},
   Year = {2005},
   ISSN = {1521-0987},
   url = {http://www.ncbi.nlm.nih.gov/pubmed/16544872},
   Abstract = {There are several population-based studies of aging, memory,
             and dementia being conducted worldwide. Of these, the Cache
             County Study on Memory, Health and Aging is noteworthy for
             its large number of "oldest-old" members. This study, which
             has been following an initial cohort of 5,092 seniors since
             1995, has reported among its major findings the role of the
             Apolipoprotein E gene on modifying the risk for Alzheimer's
             disease (AD) in males and females and identifying
             pharmacologic compounds that may act to reduce AD risk. This
             article summarizes the major findings of the Cache County
             study to date, describes ongoing investigations, and reports
             preliminary analyses on the outcome of the oldest-old in
             this population, the subgroup of participants who were over
             age 84 at the study's inception.},
   Doi = {10.1891/cmaj.6.2.107},
   Key = {fds277080}
}

@article{fds277010,
   Author = {Potter, GG and Plassman, BL and Helms, MJ and Steffens, DC and Welsh-Bohmer, KA},
   Title = {Age effects of coronary artery bypass graft on cognitive
             status change among elderly male twins.},
   Journal = {Neurology},
   Volume = {63},
   Number = {12},
   Pages = {2245-2249},
   Year = {2004},
   Month = {December},
   url = {http://www.ncbi.nlm.nih.gov/pubmed/15623681},
   Abstract = {BACKGROUND: Research regarding long-term cognitive outcome
             following coronary artery bypass graft (CABG) is
             inconsistent, which may be due in part to differential
             genetic and environmental influences within most study
             samples. METHODS: The authors examined the effect of CABG on
             cognitive status change scores in members of the National
             Academy of Sciences-National Research Council Twins Registry
             of World War II veterans. Subjects were administered the
             modified Telephone Interview for Cognitive Status (TICS-m)
             at approximately 3-year intervals between 1990 and 2002 as
             part of an epidemiologic study of dementia. RESULTS: Based
             on co-twin control analyses using a repeated-measures
             analysis of variance matching twins discordant for CABG
             within the pair (n = 464 individuals) across three age
             categories (63 to 70, 71 to 73, 74 to 83), the authors found
             at follow-up that men who had CABG between ages 63 and 70
             showed an increase in TICS-m scores and performed better
             than their co-twin who did not have the procedure. No
             significant differences were found within twin pairs for the
             older two age groups following CABG surgery. This age effect
             was replicated when comparing individuals positive for CABG
             surgery with nonfamilial, age- and education-matched
             controls who were negative for CABG. CONCLUSIONS: In this
             study of twin pairs who share many genetic and environmental
             risks for cerebrovascular problems, the results suggest that
             timing of the CABG procedure may be important to predicting
             positive cognitive outcomes.},
   Doi = {10.1212/01.wnl.0000147291.49404.0a},
   Key = {fds277010}
}

@article{fds277113,
   Author = {Steffens, DC and Welsh-Bohmer, KA and Burke, JR and Plassman, BL and Beyer, JL and Gersing, KR and Potter, GG},
   Title = {Methodology and preliminary results from the neurocognitive
             outcomes of depression in the elderly study.},
   Journal = {J Geriatr Psychiatry Neurol},
   Volume = {17},
   Number = {4},
   Pages = {202-211},
   Year = {2004},
   Month = {December},
   ISSN = {0891-9887},
   url = {http://www.ncbi.nlm.nih.gov/pubmed/15533991},
   Abstract = {A methodology is presented for following a cohort of older
             depressed patients to examine neurocognitive outcomes of
             depression. A total of 265 depressed individuals and 138
             healthy, nondepressed controls age 60 and older who
             completed at least 1 year of follow-up data underwent
             periodic clinical evaluation by a geriatric psychiatrist. A
             subset of 141 patients and 137 controls had
             neuropsychological testing. A consensus panel of experts
             reviewed 63 depressed subjects with suspected cognitive
             impairment. Twenty-seven individuals in the depressed group
             were assigned diagnoses of dementia, including 11 with
             Alzheimer's disease, 8 with vascular dementia, and 8 with
             dementia of undetermined etiology. In addition, 25
             individuals had other forms of cognitive impairment, and 11
             were considered cognitively normal. Among elderly controls,
             2 developed substantial cognitive impairment with clinical
             diagnoses of dementia. Among the depressed group, the
             incidence rates for dementia for this age are much higher
             than would be expected. These results are consistent with
             prior evidence linking depression and later dementia. Future
             studies are needed to examine neuroimaging and genetic,
             clinical, and social predictors of neurocognitive decline in
             depression.},
   Doi = {10.1177/0891988704269819},
   Key = {fds277113}
}

@article{fds277143,
   Author = {Nicodemus, KK and Stenger, JE and Schmechel, DE and Welsh-Bohmer, KA and Saunders, AM and Roses, AD and Gilbert, JR and Vance, JM and Haines, JL and Pericak-Vance, MA and Martin, ER},
   Title = {Comprehensive association analysis of APOE regulatory region
             polymorphisms in Alzheimer disease.},
   Journal = {Neurogenetics},
   Volume = {5},
   Number = {4},
   Pages = {201-208},
   Year = {2004},
   Month = {December},
   ISSN = {1364-6745},
   url = {http://www.ncbi.nlm.nih.gov/pubmed/15455263},
   Abstract = {Several recent case-control studies have examined the
             association between single nucleotide polymorphisms (SNPs)
             in the promoter region of the apolipoprotein E gene (APOE)
             and risk of Alzheimer disease (AD), with conflicting
             results. We assessed the relation between five APOE region
             SNPs and risk of AD in both case-control and family-based
             analyses. We observed a statistically significant
             association with the +5361T allele in the overall
             case-control analysis (P value=0.04) after adjusting for the
             known effect of the APOE-4 allele. Further analysis revealed
             this association to be limited to carriers of the APOE-4
             allele. Age-stratified analyses in the patients with age at
             onset of 80 years or greater and age-matched controls showed
             that the -219T allele (P value=0.009) and the +113C allele
             (P value=0.03) are associated with increased risk of AD.
             Despite these findings, haplotype and family-based
             association analyses were unable to provide evidence that
             the APOE region SNPs influenced risk of AD independent of
             the APOE-4 allele. In addition to risk, we tested for
             association between the SNPs and age at onset of AD, but
             found no association in the case-control or family based
             analyses. In conclusion, SNPs +5361, or a SNP in strong
             linkage disequilibrium, may confer some additional risk for
             developing AD beyond the risk due to APOE-4; however, the
             effect independent of APOE-4 is likely to be
             small.},
   Doi = {10.1007/s10048-004-0189-9},
   Key = {fds277143}
}

@article{fds277017,
   Author = {Labar, KS and Cook, CA and Torpey, DC and Welsh-Bohmer,
             KA},
   Title = {Impact of healthy aging on awareness and fear
             conditioning.},
   Journal = {Behav Neurosci},
   Volume = {118},
   Number = {5},
   Pages = {905-915},
   Year = {2004},
   Month = {October},
   ISSN = {0735-7044},
   url = {http://www.ncbi.nlm.nih.gov/pubmed/15506873},
   Abstract = {Fear conditioning has provided a useful model system for
             studying associative emotional learning, but the impact of
             healthy aging has gone relatively unexplored. The present
             study investigated fear conditioning across the adult life
             span in humans. A delay discrimination task was employed
             using visual conditioned stimuli and an auditory
             unconditioned stimulus. Awareness of the reinforcement
             contingencies was assessed in a postexperimental interview.
             Compared with young adult participants, middle-aged and
             older adults displayed reductions in unconditioned
             responding, discriminant conditioning, and contingency
             awareness. When awareness and overall arousability were
             taken into consideration, there were no residual effects of
             aging on conditioning. These results highlight the
             importance of considering the influence of declarative
             knowledge when interpreting age-associated changes in
             discriminative conditioned learning.},
   Doi = {10.1037/0735-7044.118.5.905},
   Key = {fds277017}
}

@article{fds277063,
   Author = {van der Walt, JM and Dementieva, YA and Martin, ER and Scott, WK and Nicodemus, KK and Kroner, CC and Welsh-Bohmer, KA and Saunders, AM and Roses, AD and Small, GW and Schmechel, DE and Murali Doraiswamy and P and Gilbert, JR and Haines, JL and Vance, JM and Pericak-Vance,
             MA},
   Title = {Analysis of European mitochondrial haplogroups with
             Alzheimer disease risk.},
   Journal = {Neurosci Lett},
   Volume = {365},
   Number = {1},
   Pages = {28-32},
   Year = {2004},
   Month = {July},
   ISSN = {0304-3940},
   url = {http://www.ncbi.nlm.nih.gov/pubmed/15234467},
   Abstract = {We examined the association of mtDNA variation with
             Alzheimer disease (AD) risk in Caucasians (989 cases and 328
             controls) testing the effect of individual haplogroups and
             single nucleotide polymorphisms (SNPs). Logistic regression
             analyses were used to assess risk of haplogroups and SNPs
             with AD in both main effects and interaction models. Males
             classified as haplogroup U showed an increase in risk (OR =
             2.30; 95% CI, 1.03-5.11; P = 0.04) of AD relative to the
             most common haplogroup H, while females demonstrated a
             significant decrease in risk with haplogroup U (OR = 0.44 ;
             95% CI, 0.24-0.80; P = 0.007). Our results were independent
             of APOE genotype, demonstrating that the effect of mt
             variation is not confounded by APOE4 carrier status. We
             suggest that variations within haplogroup U may be involved
             in AD expression in combination with environmental exposures
             or nuclear proteins other than APOE.},
   Doi = {10.1016/j.neulet.2004.04.051},
   Key = {fds277063}
}

@article{fds371184,
   Author = {Martin, E and Bronson, P and Jiang, L and Welsh-Bohmer, K and Schmechel,
             D and Small, G and Haines, J and Gilbert, J and Pericak-Vance, M and Moore,
             J},
   Title = {P4-134 A multilocus association analysis of APOE, LRP1 and
             A2M in Alzheimer disease},
   Journal = {Neurobiology of Aging},
   Volume = {25},
   Pages = {S513-S513},
   Publisher = {Elsevier BV},
   Year = {2004},
   Month = {July},
   url = {http://dx.doi.org/10.1016/s0197-4580(04)81692-0},
   Doi = {10.1016/s0197-4580(04)81692-0},
   Key = {fds371184}
}

@article{fds371185,
   Author = {Steinberg, M and Corcoran, C and Huber, C and Welsh-Bohmer, K and Norton, M and Zandi, P and Breitner, J and Tschanz, J and Lyketsos,
             C},
   Title = {P2-333 A longitudinal model for neuropsychiatric symptoms in
             dementia: the Cache County Study},
   Journal = {Neurobiology of Aging},
   Volume = {25},
   Pages = {S327-S328},
   Publisher = {Elsevier BV},
   Year = {2004},
   Month = {July},
   url = {http://dx.doi.org/10.1016/s0197-4580(04)81076-5},
   Doi = {10.1016/s0197-4580(04)81076-5},
   Key = {fds371185}
}

@article{fds371186,
   Author = {Tschanz, J and Klein, E and Treiber, K and Corcoran, C and Norton, M and Toone, L and Welsh-Bohmer, K and Steinberg, M and Munger, R and Pieper,
             C and Breitner, J and Zandi, P and Lyketsos, C},
   Title = {P2-288 Neuropsychiatric symptoms in mild cognitive
             impairment and dementia: prevalence and relationship to
             cognitive and functional impairment. The Cache County
             Study},
   Journal = {Neurobiology of Aging},
   Volume = {25},
   Pages = {S314-S314},
   Publisher = {Elsevier BV},
   Year = {2004},
   Month = {July},
   url = {http://dx.doi.org/10.1016/s0197-4580(04)81033-9},
   Doi = {10.1016/s0197-4580(04)81033-9},
   Key = {fds371186}
}

@article{fds371187,
   Author = {Toone, L and Tschanz, J and Rabins, PV and Steinberg, M and Onyike, C and Corcoran, C and Norton, M and Welsh-Bohmer, K and Breitner, J and Zandi,
             P and Lyketsos, CG},
   Title = {P2-247 A population based study of medical co-morbidity in
             early dementia and mild cognitive syndrome: association with
             functional and cognitive impairment},
   Journal = {Neurobiology of Aging},
   Volume = {25},
   Pages = {S302-S302},
   Publisher = {Elsevier BV},
   Year = {2004},
   Month = {July},
   url = {http://dx.doi.org/10.1016/s0197-4580(04)80993-x},
   Doi = {10.1016/s0197-4580(04)80993-x},
   Key = {fds371187}
}

@article{fds371188,
   Author = {Klein, E and Corcoran, C and Tschanz, J and Norton, M and Welsh-Bohmer,
             K and Breitner, J and Zandi, P and Lyketsos, C},
   Title = {P1-051 Survival from memory symptom onset: a comparison of
             individuals with dementia and cognitive impairment. The
             cache county study},
   Journal = {Neurobiology of Aging},
   Volume = {25},
   Pages = {S109-S109},
   Publisher = {Elsevier BV},
   Year = {2004},
   Month = {July},
   url = {http://dx.doi.org/10.1016/s0197-4580(04)80365-8},
   Doi = {10.1016/s0197-4580(04)80365-8},
   Key = {fds371188}
}

@article{fds371189,
   Author = {Plassman, BL and Steffens, DC and Burke, JR and Welsh-Bohmer, KA and Helms, MJ and Breitner, JC},
   Title = {P4-137 Alzheimer's disease in the NAS-NRC twin registry of
             WWII veterans},
   Journal = {Neurobiology of Aging},
   Volume = {25},
   Pages = {S514-S514},
   Publisher = {Elsevier BV},
   Year = {2004},
   Month = {July},
   url = {http://dx.doi.org/10.1016/s0197-4580(04)81695-6},
   Doi = {10.1016/s0197-4580(04)81695-6},
   Key = {fds371189}
}

@article{fds371190,
   Author = {Pericak-Vance, MA and Bronson, P and Martin, ER and Browning, C and Rayner, M and Xu, P and Small, GW and Roses, AD and Schmechel, DE and Doraiswamy, PM and Welsh-Bohmer, KA and Haines, JL and Gilbert,
             JR},
   Title = {P4-080 Genetic studies of Alzheimer disease on chromosome
             9P},
   Journal = {Neurobiology of Aging},
   Volume = {25},
   Pages = {S497-S497},
   Publisher = {Elsevier BV},
   Year = {2004},
   Month = {July},
   url = {http://dx.doi.org/10.1016/s0197-4580(04)81638-5},
   Doi = {10.1016/s0197-4580(04)81638-5},
   Key = {fds371190}
}

@article{fds371191,
   Author = {Hayden, KM and Norton, MC and Tschanz, JT and Zandi, PP and Lyketsos,
             CG and Khachaturian, AS and Pieper, CF and Welsh-Bohmer, KA and Breitner, JC},
   Title = {P3-116 The complex role of cardiovascular risk factors,
             cerebrovascular disease, and gender, with the incidence of
             Alzheimer's disease: findings from the Cache County
             Study},
   Journal = {Neurobiology of Aging},
   Volume = {25},
   Pages = {S388-S388},
   Publisher = {Elsevier BV},
   Year = {2004},
   Month = {July},
   url = {http://dx.doi.org/10.1016/s0197-4580(04)81268-5},
   Doi = {10.1016/s0197-4580(04)81268-5},
   Key = {fds371191}
}

@article{fds371192,
   Author = {Khachaturian, AS and Zandi, PP and Hayden, KM and Lyketsos, CG and Mayer, LS and Skoog, I and Tschanz, JT and Norton, MC and Welsh-Bohmer,
             KA and Breitner, JCS},
   Title = {O3-01-07 Anti-hypertensive medication use may reduce risk of
             incident Alzheimer's disease. The Cache County
             Study},
   Journal = {Neurobiology of Aging},
   Volume = {25},
   Pages = {S53-S53},
   Publisher = {Elsevier BV},
   Year = {2004},
   Month = {July},
   url = {http://dx.doi.org/10.1016/s0197-4580(04)80181-7},
   Doi = {10.1016/s0197-4580(04)80181-7},
   Key = {fds371192}
}

@article{fds371193,
   Author = {Browning, C and Vance, DD and Bronson, PG and Schmechel, D and Welsh-Bohmer, K and Scott, W and Haines, JL and Vance, JM and Pericak-Vance, MA},
   Title = {P4-072 Follow-up analysis of chromosome 2 linkage in
             early-onset Alzheimer disease},
   Journal = {Neurobiology of Aging},
   Volume = {25},
   Pages = {S494-S494},
   Publisher = {Elsevier BV},
   Year = {2004},
   Month = {July},
   url = {http://dx.doi.org/10.1016/s0197-4580(04)81630-0},
   Doi = {10.1016/s0197-4580(04)81630-0},
   Key = {fds371193}
}

@article{fds371194,
   Author = {Quan, H and Haines, JL and Sanchez-Boutard, N and Small, G and Roses, A and Schmechel, D and Welsh-Bohmer, K and Xu, P-T and Li, Y-J and Gilbert,
             JR and Vance, JM and Pericak-Vance, MA},
   Title = {P4-138 Genomic convergence on chromosome 12 in Alzheimer's
             disease},
   Journal = {Neurobiology of Aging},
   Volume = {25},
   Pages = {S514-S514},
   Publisher = {Elsevier BV},
   Year = {2004},
   Month = {July},
   url = {http://dx.doi.org/10.1016/s0197-4580(04)81696-8},
   Doi = {10.1016/s0197-4580(04)81696-8},
   Key = {fds371194}
}

@article{fds371195,
   Author = {Li, Y-J and Oliveira, SA and Qin, X-J and Roses, AD and Schmechel, D and Welsh-Bohmer, KA and Haines, JL and Pericak-Vance, MA and Vance,
             JM},
   Title = {P4-088 Association analysis of interleukin 1 (IL-1)
             polymorphisms with age-at-onset and risk of Alzheimer
             disease},
   Journal = {Neurobiology of Aging},
   Volume = {25},
   Pages = {S499-S499},
   Publisher = {Elsevier BV},
   Year = {2004},
   Month = {July},
   url = {http://dx.doi.org/10.1016/s0197-4580(04)81646-4},
   Doi = {10.1016/s0197-4580(04)81646-4},
   Key = {fds371195}
}

@article{fds371196,
   Author = {Haines, JL and Crunk, A and McFarland, L and Gaskell, P and Fuzzell, D and Creason, M and Jiang, L and Jackson, CE and Scott, WK and A.
             Welsh-Bohmer, K and Pericak-Vance, MA},
   Title = {S2-01-01 Dementia in mid-Western U.S. amish
             families},
   Journal = {Neurobiology of Aging},
   Volume = {25},
   Pages = {S24-S25},
   Publisher = {Elsevier BV},
   Year = {2004},
   Month = {July},
   url = {http://dx.doi.org/10.1016/s0197-4580(04)80079-4},
   Doi = {10.1016/s0197-4580(04)80079-4},
   Key = {fds371196}
}

@article{fds371197,
   Author = {Norton, MC and Steffens, DC and Toone, L and Tschanz, JT and Hayden, K and Corcoran, C and Klein, L and Zandi, P and Breitner, JCS and Welsh-Bohmer, KA},
   Title = {P3-108 Late-life depression, mild cognitive impairment, APOE
             and their interactive effects on conversion to dementia in a
             population-based study. The Cache County
             Study},
   Journal = {Neurobiology of Aging},
   Volume = {25},
   Pages = {S385-S386},
   Publisher = {Elsevier BV},
   Year = {2004},
   Month = {July},
   url = {http://dx.doi.org/10.1016/s0197-4580(04)81260-0},
   Doi = {10.1016/s0197-4580(04)81260-0},
   Key = {fds371197}
}

@article{fds371198,
   Author = {Zandi, PP and Szekely, CA and Green, RC and Breitner, JC and Welsh-Bohmer, KA},
   Title = {S1-03-03 Pooled analysis of the association between
             different NSAIDS and AD: preliminary findings},
   Journal = {Neurobiology of Aging},
   Volume = {25},
   Pages = {S5-S6},
   Publisher = {Elsevier BV},
   Year = {2004},
   Month = {July},
   url = {http://dx.doi.org/10.1016/s0197-4580(04)80018-6},
   Doi = {10.1016/s0197-4580(04)80018-6},
   Key = {fds371198}
}

@article{fds277021,
   Author = {Ervin, JF and Pannell, C and Szymanski, M and Welsh-Bohmer, K and Schmechel, DE and Hulette, CM},
   Title = {Vascular smooth muscle actin is reduced in Alzheimer disease
             brain: a quantitative analysis.},
   Journal = {J Neuropathol Exp Neurol},
   Volume = {63},
   Number = {7},
   Pages = {735-741},
   Year = {2004},
   Month = {July},
   ISSN = {0022-3069},
   url = {http://www.ncbi.nlm.nih.gov/pubmed/15290898},
   Abstract = {We analyzed smooth muscle actin (SMA) immunoreactivity in
             brain blood vessels of 10 ApoE 4,4 Alzheimer disease (AD)
             patients and 10 ApoE 3,3 AD patients matched for age, sex,
             and duration of dementia. We also examined 10 cognitively
             and neuropathologically normal controls matched for age and
             sex. Vascular SMA immunoreactivity in the arachnoid, grey
             matter, and white matter was quantified by image analysis.
             There was less SMA immunoreactivity in blood vessels of all
             AD patients when compared to cognitively and
             neuropathologically normal controls (p < 0.001). In
             addition, arachnoidal vessels of ApoE 4,4 AD patients had
             less SMA immunoreactivity than ApoE 3,3 AD patients (p <
             0.05). There is decreased vascular SMA density in arachnoid,
             grey matter, and white matter blood vessels in patients with
             AD when compared to age matched, cognitively and
             neuropathologically normal controls. The severity of the
             loss of SMA within the AD group may depend on ApoE
             type.},
   Doi = {10.1093/jnen/63.7.735},
   Key = {fds277021}
}

@article{fds277029,
   Author = {Hayden, KM and Pieper, CF and Welsh-Bohmer, KA and Breitner, JCS and Norton, MC and Munger, R},
   Title = {Self- or proxy-reported stroke and the risk of Alzheimer
             disease.},
   Journal = {Arch Neurol},
   Volume = {61},
   Number = {6},
   Pages = {982},
   Year = {2004},
   Month = {June},
   ISSN = {0003-9942},
   url = {http://www.ncbi.nlm.nih.gov/pubmed/15210546},
   Doi = {10.1001/archneur.61.6.982},
   Key = {fds277029}
}

@article{fds277142,
   Author = {Bigler, ED and Neeley, ES and Miller, MJ and Tate, DF and Rice, SA and Cleavinger, H and Wolfson, L and Tschanz, J and Welsh-Bohmer,
             K},
   Title = {Cerebral volume loss, cognitive deficit and
             neuropsychological performance: comparative measures of
             brain atrophy: I. Dementia.},
   Journal = {J Int Neuropsychol Soc},
   Volume = {10},
   Number = {3},
   Pages = {442-452},
   Year = {2004},
   Month = {May},
   url = {http://dx.doi.org/10.1017/S1355617704103111},
   Abstract = {There are several magnetic resonance (MR) imaging methods to
             measure brain volume and cerebral atrophy; however, the best
             measure for examining potential relationships between such
             measures and neuropsychological performance has not been
             established. Relationships between seven measures of MR
             derived brain volume or indices of atrophy and
             neuropsychological performance in the elderly subjects of
             the population-based Cache County, Utah Study of Aging and
             Memory (n = 195) were evaluated. The seven MR measures
             included uncorrected total brain volume (TBV), TBV corrected
             by total intracranial volume (TICV), TBV corrected by the
             ratio of the individuals TICV by group TICV (TBVC), a
             ventricle-to-brain ratio (VBR), total ventricular volume
             (TVV), TVV corrected by TICV, and a measure of parenchymal
             volume loss. The cases from the Cache County Study were
             comprised of elderly individuals classified into one of four
             subject groups based on a consensus diagnostic process,
             independent of quantitative MR imaging findings. The groups
             included subjects with Alzheimer's disease (AD, n = 85), no
             dementia but mild/ambiguous (M/A) deficits (n = 30), a group
             of subjects with non-AD dementia or neuropsychiatric
             disorder including vascular dementia (n = 60), and control
             subjects (n = 20). Neuropsychological performance was based
             on the Mini-Mental Status Exam (MMSE) and an expanded
             neuropsychological test battery (consortium to establish a
             registry for Alzheimer's disease (CERAD). The results
             demonstrated that the various quantitative MR measures were
             highly interrelated and no single measure was statistically
             superior. However, TBVC, TBV/TICV and VBR consistently
             exhibited the more robust relationships with
             neuropsychological performance. These results suggest that a
             single corrected brain volume measure or index is sufficient
             in studies examining global MR indicators of cerebral
             atrophy in relation to cognitive function and recommends use
             of either TBVC, TBV/TICV, or VBR.},
   Doi = {10.1017/S1355617704103111},
   Key = {fds277142}
}

@article{fds277027,
   Author = {Tschanz, JT and Corcoran, C and Skoog, I and Khachaturian, AS and Herrick, J and Hayden, KM and Welsh-Bohmer, KA and Calvert, T and Norton, MC and Zandi, P and Breitner, JCS and Cache County Study
             Group},
   Title = {Dementia: the leading predictor of death in a defined
             elderly population: the Cache County Study.},
   Journal = {Neurology},
   Volume = {62},
   Number = {7},
   Pages = {1156-1162},
   Year = {2004},
   Month = {April},
   url = {http://www.ncbi.nlm.nih.gov/pubmed/15079016},
   Abstract = {OBJECTIVE: To examine the relative risk and population
             attributable risk (PAR) of death with dementia of varying
             type and severity and other risk factors in a population of
             exceptional longevity. METHODS: Deaths were monitored over 5
             years using vital statistics records and newspaper
             obituaries in 355 individuals with prevalent dementia and
             4,328 without in Cache County, UT. Mean age was 83.3 (SD
             7.0) years with dementia and 73.7 (SD 6.8) years without.
             History of coronary artery disease, hypertension, diabetes,
             and other life-shortening illness was ascertained from
             interviews. RESULTS: Death certificates implicated dementia
             as an important cause of death, but other data suggested a
             stronger association. Adjusted Cox relative hazard and PAR
             of death were higher with dementia than with any other
             illness studied. Relative hazard of death with dementia was
             highest at ages 65 to 74, but the high prevalence of
             dementia after age 85 resulted in 27% PAR among the oldest
             old. Mortality increased substantially with severity of
             dementia. Alzheimer disease shortened survival time most
             dramatically in younger participants, but vascular dementia
             posed a greater mortality risk among the oldest old.
             CONCLUSION: In this population, dementia was the strongest
             predictor of mortality, with a risk two to three times those
             of other life-shortening illnesses.},
   Doi = {10.1212/01.wnl.0000118210.12660.c2},
   Key = {fds277027}
}

@article{fds277051,
   Author = {Li, YL and Oliveira, SA and Xu, P and Martin, ER and Stenger, JE and Hulette, C and Scherzer, CR and Hauser, MA and Scott, WK and Small, GW and Nance, MA and Watts, RL and Hubble, JP and Koller, WC and Pahwa, R and Stern, MB and Hiner, BC and Jankovic, J and Goetz, CG and Mastaglia, F and Middleton, LT and Roses, AD and Saunders, AM and Welsh-Bohmer, KA and Schmechel, DE and Gullans, SR and Haines, JL and Gilbert, JR and Vance,
             JM and Pericak-Vance, MA},
   Title = {Erratum: Glutathione S-transferase omega-1 modifies
             age-at-onset of Alzheimer disease and Parkinson disease
             (Human Molecular Genetics (2003) vol. 12
             (3259-3267))},
   Journal = {Human Molecular Genetics},
   Volume = {13},
   Number = {5},
   Pages = {573},
   Publisher = {Oxford University Press (OUP)},
   Year = {2004},
   Month = {March},
   url = {http://dx.doi.org/10.1093/hmg/ddh059},
   Doi = {10.1093/hmg/ddh059},
   Key = {fds277051}
}

@article{fds277141,
   Author = {Zandi, PP and Anthony, JC and Khachaturian, AS and Stone, SV and Gustafson, D and Tschanz, JT and Norton, MC and Welsh-Bohmer, KA and Breitner, JCS and Cache County Study Group},
   Title = {Reduced risk of Alzheimer disease in users of antioxidant
             vitamin supplements: the Cache County Study.},
   Journal = {Arch Neurol},
   Volume = {61},
   Number = {1},
   Pages = {82-88},
   Year = {2004},
   Month = {January},
   ISSN = {0003-9942},
   url = {http://www.ncbi.nlm.nih.gov/pubmed/14732624},
   Abstract = {BACKGROUND: Antioxidants may protect the aging brain against
             oxidative damage associated with pathological changes of
             Alzheimer disease (AD). OBJECTIVE: To examine the
             relationship between antioxidant supplement use and risk of
             AD. DESIGN: Cross-sectional and prospective study of
             dementia. Elderly (65 years or older) county residents were
             assessed in 1995 to 1997 for prevalent dementia and AD, and
             again in 1998 to 2000 for incident illness. Supplement use
             was ascertained at the first contact. SETTING: Cache County,
             Utah. PARTICIPANTS: Among 4740 respondents (93%) with data
             sufficient to determine cognitive status at the initial
             assessment, we identified 200 prevalent cases of AD. Among
             3227 survivors at risk, we identified 104 incident AD cases
             at follow-up. MAIN OUTCOME MEASURE: Diagnosis of AD by means
             of multistage assessment procedures. RESULTS: Analyses of
             prevalent and incident AD yielded similar results. Use of
             vitamin E and C (ascorbic acid) supplements in combination
             was associated with reduced AD prevalence (adjusted odds
             ratio, 0.22; 95% confidence interval, 0.05-0.60) and
             incidence (adjusted hazard ratio, 0.36; 95% confidence
             interval, 0.09-0.99). A trend toward lower AD risk was also
             evident in users of vitamin E and multivitamins containing
             vitamin C, but we saw no evidence of a protective effect
             with use of vitamin E or vitamin C supplements alone, with
             multivitamins alone, or with vitamin B-complex supplements.
             CONCLUSIONS: Use of vitamin E and vitamin C supplements in
             combination is associated with reduced prevalence and
             incidence of AD. Antioxidant supplements merit further study
             as agents for the primary prevention of AD.},
   Doi = {10.1001/archneur.61.1.82},
   Key = {fds277141}
}

@article{fds371199,
   Author = {Norton, M and Klein, L and Welsh-Bohmer, K and Tschanz, J and Toone, L and Steffens, D and Zandi, P and Corcoran, C and Hayden, K and Breitner,
             J},
   Title = {Late-life depression, mild cognitive impairment, APOE and
             their interactive effects on 3-year conversion to
             dementia},
   Journal = {GERONTOLOGIST},
   Volume = {44},
   Pages = {219-219},
   Year = {2004},
   Key = {fds371199}
}

@article{fds371200,
   Author = {Klein, E and Tschanz, J and Corcoran, C and Norton, M and Welsh-Bohmer,
             K and Breitner, J and Zandi, P and Lyketsos, CC},
   Title = {Estimating survival duration from memory symptom onset: A
             comparison of methods. The Cache County Study.},
   Journal = {AMERICAN JOURNAL OF EPIDEMIOLOGY},
   Volume = {159},
   Number = {11},
   Pages = {S3-S3},
   Year = {2004},
   Key = {fds371200}
}

@article{fds371201,
   Author = {Norton, M and Skoog, I and Toone, L and Tschanz, J and Corcoran, C and Zandi, P and Hart, A and Breitner, J and Welsh-Bohmer, K and Steffens,
             D},
   Title = {Improving assessment of incidence of first-onset geriatric
             depression in population-based studies.},
   Journal = {AMERICAN JOURNAL OF EPIDEMIOLOGY},
   Volume = {159},
   Number = {11},
   Pages = {S2-S2},
   Year = {2004},
   Key = {fds371201}
}

@article{fds276974,
   Author = {Steffens, DC and McQuoid, DR and Welsh-Bohmer, KA and Krishnan,
             KRR},
   Title = {Left orbital frontal cortex volume and performance on the
             benton visual retention test in older depressives and
             controls.},
   Journal = {Neuropsychopharmacology},
   Volume = {28},
   Number = {12},
   Pages = {2179-2183},
   Year = {2003},
   Month = {December},
   ISSN = {0893-133X},
   url = {http://www.ncbi.nlm.nih.gov/pubmed/14532909},
   Abstract = {Changes within the prefrontal cortex (PFC) have been
             associated with both mood disorders and with specific
             impairments in cognitive testing. The left PFC has been
             implicated in relational processing, that is, binding
             different pieces of information. We hypothesized that among
             older depressives and elderly controls, lower performance on
             one test of relational processing would be associated with
             smaller volume of the orbital frontal cortex (OFC). A total
             of 30 depressed and 40 control subjects were included in the
             study. All subjects were administered the Benton Visual
             Retention Test (BVRT). Subjects received a standardized
             magnetic resonance imaging, for which volumes of the OFC and
             total brain were calculated. We found that, controlling for
             age and education, total correct on BVRT was associated with
             left OFC volume normalized for total brain volume among the
             entire sample. For the depressed sample only, the number of
             perseverative errors was negatively associated with left OFC
             volume normalized for total brain volume after controlling
             for age and education. These results add to the literature
             linking mood and cognitive disturbances to the PFC. Future
             studies with a larger sample employing functional measures
             are warranted.},
   Doi = {10.1038/sj.npp.1300285},
   Key = {fds276974}
}

@article{fds371202,
   Author = {Vance, JM and Li, Y and Oliveira, AS and Hauser, MA and Martin, ER and Scott, WK and Stenger, JE and Scherzer, C and Gullans, SR and Small, GW and Roses, AD and Saunders, AM and Schmechel, DE and Welsh-Bohmer, KA and Hulette, C and Haines, JL and Gilbert, JR and Pericak-Vance,
             MA},
   Title = {Glutathione S-transferase, omega-1 (GSTO1) modifies age at
             onset of Alzheimer disease and Parkinson
             disease.},
   Journal = {AMERICAN JOURNAL OF HUMAN GENETICS},
   Volume = {73},
   Number = {5},
   Pages = {182-182},
   Publisher = {UNIV CHICAGO PRESS},
   Year = {2003},
   Month = {November},
   Key = {fds371202}
}

@article{fds277052,
   Author = {Scott, WK and Hauser, ER and Schmechel, DE and Welsh-Bohmer, KA and Small, GW and Roses, AD and Saunders, AM and Gilbert, JR and Vance, JM and Haines, JL and Pericak-Vance, MA},
   Title = {Ordered-subsets linkage analysis detects novel Alzheimer
             disease loci on chromosomes 2q34 and 15q22.},
   Journal = {Am J Hum Genet},
   Volume = {73},
   Number = {5},
   Pages = {1041-1051},
   Year = {2003},
   Month = {November},
   ISSN = {0002-9297},
   url = {http://www.ncbi.nlm.nih.gov/pubmed/14564669},
   Abstract = {Alzheimer disease (AD) is a complex disorder characterized
             by a wide range, within and between families, of ages at
             onset of symptoms. Consideration of age at onset as a
             covariate in genetic-linkage studies may reduce genetic
             heterogeneity and increase statistical power.
             Ordered-subsets analysis includes continuous covariates in
             linkage analysis by rank ordering families by a covariate
             and summing LOD scores to find a subset giving a
             significantly increased LOD score relative to the overall
             sample. We have analyzed data from 336 markers in 437
             multiplex (>/=2 sampled individuals with AD) families
             included in a recent genomic screen for AD loci. To identify
             genetic heterogeneity by age at onset, families were ordered
             by increasing and decreasing mean and minimum ages at onset.
             Chromosomewide significance of increases in the LOD score in
             subsets relative to the overall sample was assessed by
             permutation. A statistically significant increase in the
             nonparametric multipoint LOD score was observed on
             chromosome 2q34, with a peak LOD score of 3.2 at D2S2944
             (P=.008) in 31 families with a minimum age at onset between
             50 and 60 years. The LOD score in the chromosome 9p region
             previously linked to AD increased to 4.6 at D9S741 (P=.01)
             in 334 families with minimum age at onset between 60 and 75
             years. LOD scores were also significantly increased on
             chromosome 15q22: a peak LOD score of 2.8 (P=.0004) was
             detected at D15S1507 (60 cM) in 38 families with minimum age
             at onset >/=79 years, and a peak LOD score of 3.1 (P=.0006)
             was obtained at D15S153 (62 cM) in 43 families with mean age
             at onset >80 years. Thirty-one families were contained in
             both 15q22 subsets, indicating that these results are likely
             detecting the same locus. There is little overlap in these
             subsets, underscoring the utility of age at onset as a
             marker of genetic heterogeneity. These results indicate that
             linkage to chromosome 9p is strongest in late-onset AD and
             that regions on chromosome 2q34 and 15q22 are linked to
             early-onset AD and very-late-onset AD, respectively.},
   Doi = {10.1086/379083},
   Key = {fds277052}
}

@article{fds277122,
   Author = {Mackensen, GB and Ti, LK and Phillips-Bute, BG and Mathew, JP and Newman, MF and Grocott, HP and Neurologic Outcome Research Group
             (NORG)},
   Title = {Cerebral embolization during cardiac surgery: impact of
             aortic atheroma burden.},
   Journal = {Br J Anaesth},
   Volume = {91},
   Number = {5},
   Pages = {656-661},
   Year = {2003},
   Month = {November},
   ISSN = {0007-0912},
   url = {http://www.ncbi.nlm.nih.gov/pubmed/14570786},
   Abstract = {BACKGROUND: Aortic atheromatous disease is known to be
             associated with an increased risk of perioperative stroke in
             the setting of cardiac surgery. In this study, we sought to
             determine the relationship between cerebral microemboli and
             aortic atheroma burden in patients undergoing cardiac
             surgery. METHODS: Transoesophageal echocardiographic images
             of the ascending, arch and descending aorta were evaluated
             in 128 patients to determine the aortic atheroma burden.
             Transcranial Doppler (TCD) of the right middle cerebral
             artery was performed in order to measure cerebral embolic
             load during surgery. Using multivariate linear regression,
             the numbers of emboli were compared with the atheroma
             burden. RESULTS: After controlling for age, cardiopulmonary
             bypass time and the number of bypass grafts, cerebral emboli
             were significantly associated with atheroma in the ascending
             aorta (R2=0.11, P=0.02) and aortic arch (P=0.013). However,
             there was no association between emboli and descending
             aortic atheroma burden (R2=0.05, P=0.20). CONCLUSIONS: We
             demonstrate a positive relationship between TCD-detected
             cerebral emboli and the atheromatous burden of the ascending
             aorta and aortic arch. Previously demonstrated associations
             between TCD-detectable cerebral emboli and adverse cerebral
             outcome may be related to the presence of significant aortic
             atheromatous disease.},
   Doi = {10.1093/bja/aeg234},
   Key = {fds277122}
}

@article{fds277009,
   Author = {Bigler, ED and Lowry, CM and Kerr, B and Tate, DF and Hessel, CD and Earl,
             HD and Miller, MJ and Rice, SA and Smith, KH and Tschanz, JT and Welsh-Bohmer, K and Plassman, B and Victoroff,
             J},
   Title = {Role of white matter lesions, cerebral atrophy, and APOE on
             cognition in older persons with and without dementia: the
             Cache County, Utah, study of memory and aging.},
   Journal = {Neuropsychology},
   Volume = {17},
   Number = {3},
   Pages = {339-352},
   Year = {2003},
   Month = {July},
   url = {http://dx.doi.org/10.1037/0894-4105.17.3.339},
   Abstract = {Neuropsychological, qualitative, and quantitative magnetic
             resonance imaging findings were examined in subjects with
             Alzheimer's disease (AD), non-AD dementia or mixed
             neuropsychiatric disorder, subjects characterized as
             mild/ambiguous, and controls, all with known apolipoprotein
             E (APOE) genotype. Neuropsychological tasks included an
             expanded Consortium to Establish a Registery for Alzheimer's
             Disease (J. T. Tschanz et al., 2000; K. A. Welsh, J. M.
             Hoffman, N. L. Earl, & M. W. Hanson 1994) battery and the
             Mini-Mental Status Examination (M. F. Folstein, S. E.
             Folstein, & P. R. McHugh, 1975). Periventricular white
             matter lesions were the most clinically salient, and
             generalized measures of cerebral atrophy were the most
             significant quantitative indicators. APOE genotype was
             unrelated to imaging or neuropsychological performance.
             Neuropsychological relationships with neuroimaging findings
             depend on the qualitative or quantitative method
             used.},
   Doi = {10.1037/0894-4105.17.3.339},
   Key = {fds277009}
}

@article{fds277078,
   Author = {Steffens, DC and Norton, MC and Hart, AD and Skoog, I and Corcoran, C and Breitner, JCS and Cache County Study Group},
   Title = {Apolipoprotein E genotype and major depression in a
             community of older adults. The Cache County
             Study.},
   Journal = {Psychol Med},
   Volume = {33},
   Number = {3},
   Pages = {541-547},
   Year = {2003},
   Month = {April},
   ISSN = {0033-2917},
   url = {http://www.ncbi.nlm.nih.gov/pubmed/12701674},
   Abstract = {BACKGROUND: The role of allelic variation in APOE, the
             genetic locus for apolipoprotein E, in geriatric depression
             is poorly understood. There are conflicting reports as to an
             association between the epsilon4 allele and depression in
             late life. METHOD: Using a community based study of
             non-demented elders in Cache County, Utah, that included
             many very old individuals, we examined the relationship
             between APOE and late-onset (age > 60) depression, with
             particular attention to possible age effects. RESULTS: There
             was no overall association between APOE and depression.
             However, there was a significant interaction effect of APOE
             and age such that the relationship of late-onset depression
             with respect to presence of the epsilon4 allele was larger
             among those 80 and older compared with those below age 80.
             Consistent with previous studies, women were more likely to
             experience late-onset depression than men. CONCLUSIONS:
             Because we excluded prevalent cases of dementia, this
             pattern of relative risk with age may reflect the appearance
             of depressive symptoms as a prodrome of Alzheimer's disease
             or vascular dementia. Longitudinal studies should help to
             confirm or refute this explanation of the
             data.},
   Doi = {10.1017/s0033291702007201},
   Key = {fds277078}
}

@article{fds371203,
   Author = {Welsh-Bohmer, KA},
   Title = {Defining "prodromal" Alzheimer's disease and the factors
             that modify its trajectory to dementia: Clues from
             population based studies},
   Journal = {CLINICAL NEUROPSYCHOLOGIST},
   Volume = {17},
   Number = {1},
   Pages = {97-97},
   Publisher = {SWETS ZEITLINGER PUBLISHERS},
   Year = {2003},
   Month = {February},
   Key = {fds371203}
}

@article{fds371204,
   Author = {Broadbent, D and Norton, M and Tschanz, J and Welsh-Bohmer, K and Breitner, J},
   Title = {Effect of prior and recent health problems on performance of
             activities of daily living by the elderly: The cache county
             study},
   Journal = {GERONTOLOGIST},
   Volume = {43},
   Pages = {458-459},
   Year = {2003},
   Key = {fds371204}
}

@article{fds371205,
   Author = {Franklin, L and Norton, M and Steffens, D and Tschanz, J and Welsh-Bohmer, K and Breitner, J},
   Title = {Couples' religious parity, their depression history and
             social support, as predictors of new-onset major depression
             in elderly wives. The Cache County Study.},
   Journal = {GERONTOLOGIST},
   Volume = {43},
   Pages = {392-392},
   Year = {2003},
   Key = {fds371205}
}

@article{fds371206,
   Author = {Hayden, K and Zandi, P and Tschanz, J and Khachaturian, A and Garrett,
             K and Welsh-Bohmer, K and Breitner, J},
   Title = {APOE genotype and mortality: The Cache County
             study},
   Journal = {GERONTOLOGIST},
   Volume = {43},
   Pages = {8-8},
   Year = {2003},
   Key = {fds371206}
}

@article{fds371207,
   Author = {Yoash-Gantz, RE and DiPersio, DA and Fahey, F and Welsh-Bohmer,
             K},
   Title = {Asymmetry of hemispheric function in MCI as measured by
             FDG-PET},
   Journal = {ARCHIVES OF CLINICAL NEUROPSYCHOLOGY},
   Volume = {18},
   Number = {7},
   Pages = {694-694},
   Year = {2003},
   Key = {fds371207}
}

@article{fds371208,
   Author = {Johnson, S and Welsh-Bohmer, K and Wagner, RH and Steffens,
             DC},
   Title = {Expression of geriatric depression in African-Americans and
             Caucasians},
   Journal = {CLINICAL NEUROPSYCHOLOGIST},
   Volume = {17},
   Number = {1},
   Pages = {113-113},
   Year = {2003},
   Key = {fds371208}
}

@article{fds277110,
   Author = {Phillips Bute and B and Mathew, J and Blumenthal, JA and Welsh-Bohmer,
             K and White, WD and Mark, D and Landolfo, K and Newman,
             MF},
   Title = {Female gender is associated with impaired quality of life 1
             year after coronary artery bypass surgery.},
   Journal = {Psychosom Med},
   Volume = {65},
   Number = {6},
   Pages = {944-951},
   Year = {2003},
   url = {http://www.ncbi.nlm.nih.gov/pubmed/14645771},
   Abstract = {OBJECTIVE: To evaluate gender-related differences in quality
             of life (QOL) and cognitive function 1 year after coronary
             artery bypass surgery (CABG) after adjusting for known
             baseline differences. MATERIALS AND METHODS: Two hundred
             eighty patients (96 women and 184 men) underwent
             neurocognitive and QOL evaluation at baseline
             (preoperatively) and at 1 year after CABG. Multivariable
             linear regression was used to assess the relationship of
             gender to follow-up QOL and cognitive function. Measures
             used to evaluate QOL were IADL, DASI, work activities
             (SF-36), social activities, social support, general health
             perception (SF-36), CESD, STAI, and symptom limitations.
             Cognitive function was measured with a battery of
             performance-based neuropsychological tests, reduced to a
             four-cognitive domain scores with factor analysis, and a
             self-report measure of cognitive difficulties. Covariates in
             multiple regression models included age, years of education,
             marital status, Charlson Comorbidity Index, hypertension,
             diabetes, race, and baseline QOL/cognitive status. RESULTS:
             Female patients showed significantly worse outcome than male
             patients at 1 year follow-up in several key areas of QOL.
             After adjusting for baseline differences, women are at
             greater risk for increased cognitive difficulties (p= 0.04)
             and anxiety (p= 0.03), as well as impaired DASI (p= 0.02),
             IADL (p= 0.03), and work activities (p= 0.02). Cognitive
             sequelae attributable to bypass surgery were similar between
             men and women. CONCLUSIONS: Even after adjusting for known
             risk factors for compromised QOL and cognitive functioning,
             women do not show the same long-term quality benefits of
             CABG surgery that men do.},
   Doi = {10.1097/01.psy.0000097342.24933.a2},
   Key = {fds277110}
}

@article{fds277008,
   Author = {Zandi, PP and Carlson, MC and Plassman, BL and Welsh-Bohmer, KA and Mayer, LS and Steffens, DC and Breitner, JCS and Cache County Memory
             Study Investigators},
   Title = {Hormone replacement therapy and incidence of Alzheimer
             disease in older women: the Cache County
             Study.},
   Journal = {JAMA},
   Volume = {288},
   Number = {17},
   Pages = {2123-2129},
   Year = {2002},
   Month = {November},
   ISSN = {0098-7484},
   url = {http://www.ncbi.nlm.nih.gov/pubmed/12413371},
   Abstract = {CONTEXT: Previous studies have shown a sex-specific
             increased risk of Alzheimer disease (AD) in women older than
             80 years. Basic neuroscience findings suggest that hormone
             replacement therapy (HRT) could reduce a woman's risk of AD.
             Epidemiologic findings on AD and HRT are mixed. OBJECTIVE:
             To examine the relationship between use of HRT and risk of
             AD among elderly women. DESIGN, SETTING, AND PARTICIPANTS:
             Prospective study of incident dementia among 1357 men (mean
             age, 73.2 years) and 1889 women (mean age, 74.5 years)
             residing in a single county in Utah. Participants were first
             assessed in 1995-1997, with follow-up conducted in
             1998-2000. History of women's current and former use of HRT,
             as well as of calcium and multivitamin supplements, was
             ascertained at the initial contact. MAIN OUTCOME MEASURE:
             Diagnosis of incident AD. RESULTS: Thirty-five men (2.6%)
             and 88 women (4.7%) developed AD between the initial
             interview and time of the follow-up (3 years). Incidence
             among women increased after age 80 years and exceeded the
             risk among men of similar age (adjusted hazard ratio [HR],
             2.11; 95% confidence interval [CI], 1.22-3.86). Women who
             used HRT had a reduced risk of AD (26 cases among 1066
             women) compared with non-HRT users (58 cases among 800
             women) (adjusted HR, 0.59; 95% CI, 0.36-0.96). Risk varied
             with duration of HRT use, so that a woman's sex-specific
             increase in risk disappeared entirely with more than 10
             years of treatment (7 cases among 427 women). Adjusted HRs
             were 0.41 (95% CI, 0.17-0.86) for HRT users compared with
             nonusers and 0.77 (95% CI, 0.31-1.67) compared with men. No
             similar effect was seen with calcium or multivitamin use.
             Almost all of the HRT-related reduction in incidence
             reflected former use of HRT (9 cases among 490 women;
             adjusted HR, 0.33 [95% CI, 0.15-0.65]). There was no effect
             with current HRT use (17 cases among 576 women; adjusted HR,
             1.08 [95% CI, 0.59-1.91]) unless duration of treatment
             exceeded 10 years (6 cases among 344 women; adjusted HR,
             0.55 [95% CI, 0.21-1.23]). CONCLUSIONS: Prior HRT use is
             associated with reduced risk of AD, but there is no apparent
             benefit with current HRT use unless such use has exceeded 10
             years.},
   Doi = {10.1001/jama.288.17.2123},
   Key = {fds277008}
}

@article{fds277007,
   Author = {Bigler, ED and Tate, DF and Miller, MJ and Rice, SA and Hessel, CD and Earl, HD and Tschanz, JT and Plassman, B and Welsh-Bohmer,
             KA},
   Title = {Dementia, asymmetry of temporal lobe structures, and
             apolipoprotein E genotype: relationships to cerebral atrophy
             and neuropsychological impairment.},
   Journal = {J Int Neuropsychol Soc},
   Volume = {8},
   Number = {7},
   Pages = {925-933},
   Year = {2002},
   Month = {November},
   ISSN = {1355-6177},
   url = {http://www.ncbi.nlm.nih.gov/pubmed/12405544},
   Abstract = {We examined asymmetry of hippocampal volume as well as other
             temporal lobe structures (temporal lobe, temporal horn of
             the lateral ventricular system, parahippocampal and fusiform
             gyri) in 194 subjects from the Cache County, Utah study,
             with varying disorders [85 with Alzheimer's disease (AD), 59
             with some cognitive or neuropsychiatric disorder-referenced
             as a Mixed Neuropsychiatric group, 30 with mild
             ambiguous/mild cognitive impairment (MA/MCI) and 20
             controls] and APOE genotypes. Asymmetry was determined by
             subtracting left-side volume from the right corrected by
             total intracranial volume. No significant asymmetry was
             observed to be associated with presence of the epsilon4
             allele. Since the AD-epsilon4 allele risk effect may be
             expressed early in the course of the disorder, we also
             examined asymmetry indices in AD, MA/MCI and Mixed
             Neuropsychiatric subjects early in the course of their
             disorder (2 years or less) to those with longer duration
             illness (greater than 2 years). We observed a leftward
             asymmetry (i.e., left side larger) regardless of APOE
             genotype in hippocampal volume where both AD and MCI
             subjects demonstrated a leftward shift in hippocampal size
             when length of disease (LOD) was less but not more than 2
             years. Leftward asymmetry was not associated with LOD in the
             Mixed Neuropsychiatric group. These findings do not support
             an association between epsilon4 and hippocampal asymmetry in
             dementia. We also examined whether asymmetry influenced
             neuropsychological performance, but minimal effects were
             observed. Where significance or strong trends were observed,
             better neuropsychological performance was associated with
             larger parenchymal volume of temporal lobe structures. These
             findings were interpreted as representing cognitive reserve
             effects where larger volume was protective against
             impairment. The role of asymmetry research in understanding
             neuropsychological performance in dementia is
             discussed.},
   Doi = {10.1017/s1355617702870072},
   Key = {fds277007}
}

@article{fds277121,
   Author = {Swaminathan, M and McCreath, BJ and Phillips-Bute, BG and Newman, MF and Mathew, JP and Smith, PK and Blumenthal, JA and Stafford-Smith, M and Perioperative Outcomes Research Group},
   Title = {Serum creatinine patterns in coronary bypass surgery
             patients with and without postoperative cognitive
             dysfunction.},
   Journal = {Anesth Analg},
   Volume = {95},
   Number = {1},
   Pages = {1-8},
   Year = {2002},
   Month = {July},
   ISSN = {0003-2999},
   url = {http://www.ncbi.nlm.nih.gov/pubmed/12088934},
   Abstract = {UNLABELLED: Renal dysfunction is common after coronary
             artery bypass graft (CABG) surgery. We have previously shown
             that CABG procedures complicated by stroke have a threefold
             greater peak serum creatinine level relative to
             uncomplicated surgery. However, postoperative creatinine
             patterns for procedures complicated by cognitive dysfunction
             are unknown. Therefore, we tested the hypothesis that
             postoperative cognitive dysfunction is associated with acute
             perioperative renal injury after CABG surgery. Data were
             prospectively gathered for 282 elective CABG surgery
             patients. Psychometric tests were performed at baseline and
             6 wk after surgery. Cognitive dysfunction was defined both
             as a dichotomous variable (cognitive deficit [CD]) and as a
             continuous variable (cognitive index). Forty percent of
             patients had CD at 6 wk. However, the association between
             peak percentage change in postoperative creatinine and CD
             (parameter estimate = -0.41; P = 0.91) or cognitive index
             (parameter estimate = -1.29; P = 0.46) was not significant.
             These data indicate that postcardiac surgery cognitive
             dysfunction, unlike stroke, is not associated with major
             increases in postoperative renal dysfunction. IMPLICATIONS:
             We previously noted that patients with postcardiac surgery
             stroke also have greater acute renal injury than unaffected
             patients. However, in the same setting, we found no
             difference in renal injury between patients with and without
             cognitive dysfunction. Factors responsible for subtle
             postoperative cognitive dysfunction do not appear to be
             associated with clinically important renal
             effects.},
   Doi = {10.1097/00000539-200207000-00001},
   Key = {fds277121}
}

@article{fds371209,
   Author = {Li, YJ and Scott, WK and Hedges, DK and Zhang, FY and Gaskell, PC and Stajich, JM and Saunders, AM and Scott, BL and Schmechel, DE and Welsh-Bohmer, KA and Conneally, PM and Gilbert, JR and Vance, JM and Pericak-Vance, MA and Nance, MA and Watts, RL and Hubble, JP and Koller,
             WC and Pahwa, R and Stern, MB and Hiner, BC and Jankovic, J and Allen, FH and Goetz, CG and Mastaglia, F and Gibson, RA and Middelton, LT and Small,
             GW and Roses, AD and Haines, JL},
   Title = {Location of common age-at onset genes in Alzheimer and
             Parkinson diseases},
   Journal = {NEUROBIOLOGY OF AGING},
   Volume = {23},
   Number = {1},
   Pages = {S317-S317},
   Publisher = {ELSEVIER SCIENCE INC},
   Year = {2002},
   Month = {July},
   Key = {fds371209}
}

@article{fds276986,
   Author = {Li, Y-J and Scott, WK and Hedges, DJ and Zhang, F and Gaskell, PC and Nance, MA and Watts, RL and Hubble, JP and Koller, WC and Pahwa, R and Stern, MB and Hiner, BC and Jankovic, J and Allen, FA and Goetz, CG and Mastaglia, F and Stajich, JM and Gibson, RA and Middleton, LT and Saunders, AM and Scott, BL and Small, GW and Nicodemus, KK and Reed, AD and Schmechel, DE and Welsh-Bohmer, KA and Conneally, PM and Roses, AD and Gilbert, JR and Vance, JM and Haines, JL and Pericak-Vance,
             MA},
   Title = {Age at onset in two common neurodegenerative diseases is
             genetically controlled.},
   Journal = {Am J Hum Genet},
   Volume = {70},
   Number = {4},
   Pages = {985-993},
   Year = {2002},
   Month = {April},
   ISSN = {0002-9297},
   url = {http://www.ncbi.nlm.nih.gov/pubmed/11875758},
   Abstract = {To identify genes influencing age at onset (AAO) in two
             common neurodegenerative diseases, a genomic screen was
             performed for AAO in families with Alzheimer disease (AD;
             n=449) and Parkinson disease (PD; n=174). Heritabilities
             between 40%--60% were found in both the AD and PD data sets.
             For PD, significant evidence for linkage to AAO was found on
             chromosome 1p (LOD = 3.41). For AD, the AAO effect of APOE
             (LOD = 3.28) was confirmed. In addition, evidence for AAO
             linkage on chromosomes 6 and 10 was identified independently
             in both the AD and PD data sets. Subsequent unified analyses
             of these regions identified a single peak on chromosome 10q
             between D10S1239 and D10S1237, with a maximum LOD score of
             2.62. These data suggest that a common gene affects AAO in
             these two common complex neurodegenerative
             diseases.},
   Doi = {10.1086/339815},
   Key = {fds276986}
}

@article{fds371210,
   Author = {Norton, MC and Steffens, DC and Skoog, I and Welsh-Bohmer, KA and Wyse,
             BW and Breitner, JC},
   Title = {Relation of functional impairment, social support, and
             religious involvement to incident depression over a
             three-year interval.},
   Journal = {AMERICAN JOURNAL OF GERIATRIC PSYCHIATRY},
   Volume = {10},
   Number = {2},
   Pages = {76-76},
   Publisher = {AMER PSYCHIATRIC PRESS, INC},
   Year = {2002},
   Month = {March},
   Key = {fds371210}
}

@article{fds277077,
   Author = {Tschanz, JT and Welsh-Bohmer, KA and Plassman, BL and Norton, MC and Wyse, BW and Breitner, JCS and Cache County Study
             Group},
   Title = {An adaptation of the modified mini-mental state examination:
             analysis of demographic influences and normative data: the
             cache county study.},
   Journal = {Neuropsychiatry Neuropsychol Behav Neurol},
   Volume = {15},
   Number = {1},
   Pages = {28-38},
   Year = {2002},
   Month = {March},
   ISSN = {0894-878X},
   url = {http://www.ncbi.nlm.nih.gov/pubmed/11877549},
   Abstract = {OBJECTIVES: To present a new version of the Modified
             Mini-Mental State Examination (3MS-R), provide normative
             information extending to individuals in the 10th decade, and
             examine the effects of demographic variables on test
             performance. BACKGROUND: The Modified Mini-Mental State
             Examination, based originally on the Mini-Mental State
             Examination, has been used to screen populations for
             dementia. Providing normative information and an analysis of
             demographic variables on test performance for this version
             would support broader use in clinical and other settings.
             METHODS: Two thousand, nine hundred thirteen elderly
             individuals determined to be free of dementia and other
             neurologic and psychiatric conditions served as subjects. An
             analysis of variance was conducted to examine the effects of
             age, gender, and education on test performance. Descriptive
             statistics (means, standard deviations, and percentile
             ranks) were calculated to summarize the range of normal
             performance. To examine the sensitivity/specificity of the
             suggested cut-off points at the 7th and 10th percentiles,
             two subsamples of elderly individuals, on whom clinical
             dementia assessments were available, were used to classify
             individuals with regard to dementia status. RESULTS: Lower
             age, higher education, and female gender were associated
             with higher 3MS-R scores. Gender effects were among the
             weakest, but most important at lower levels of education.
             Education effects were most prominent in the youngest age
             groups. Selection of a cut-off point at the 7th percentile
             revealed 69%-70% sensitivity for detecting dementia, and
             higher sensitivity for individuals in the youngest age
             groups. Specificity at this cut-off point was 89%. Raising
             the cut-off point to the 10th percentile improved
             sensitivity to 73%-76%, but reduced specificity to 85%-86%.
             CONCLUSION: We present a version of the Modified Mini-Mental
             State Examination that has demonstrated utility in screening
             a population for dementia. An analysis of normative
             information and the effects of demographic influences
             suggest that the 7th percentile cut-off point performs very
             well in detecting dementia in 65-79-year-old individuals but
             less well for individuals in their 80s and 90s. To increase
             the sensitivity of the 3MS-R to detect dementia or other
             forms of cognitive impairment, particularly among the
             "old-old," the test user may wish to raise the cut-off point
             for impairment in some demographic groups or to supplement
             the test with additional cognitive measures.},
   Key = {fds277077}
}

@article{fds371211,
   Author = {Norton, M and Tschanz, J and Corcoran, C and Mumford, S and Welsh-Bohmer, K and Breitner, J},
   Title = {Apolipoprotein ε4 interacts with mild cognitive deficit to
             shorten time to dementia onset},
   Journal = {NEUROBIOLOGY OF AGING},
   Volume = {23},
   Number = {1},
   Pages = {S299-S299},
   Year = {2002},
   Key = {fds371211}
}

@article{fds371212,
   Author = {Tschanz, J and Norton, M and Corcoran, C and LaCaille, R and Welsh-Bohmer, K and Breitner, J},
   Title = {Cognitive screening and self-perception of memory problems
             predict mild cognitive impairment and dementia},
   Journal = {NEUROBIOLOGY OF AGING},
   Volume = {23},
   Number = {1},
   Pages = {S262-S262},
   Year = {2002},
   Key = {fds371212}
}

@article{fds371213,
   Author = {Lisota, R and Tschanz, J and Norton, M and Corcoran, C and Leslie, C and Lyketsos, C and Steinberg, M and Breitner, J and Welsh-Bohmer,
             K},
   Title = {Differential impact of genetic and demographic variables on
             clinical course of dementia and Alzheimer's
             disease},
   Journal = {NEUROBIOLOGY OF AGING},
   Volume = {23},
   Number = {1},
   Pages = {S152-S153},
   Year = {2002},
   Key = {fds371213}
}

@article{fds371214,
   Author = {Gaskell, PC and Hulette, C and Welsh-Bohmer, K and Schmechel, D and Pericak-Vance, M and Roses, A},
   Title = {Followup of multiplex Alzheimer disease (AD) families:
             Presentation of data and discussion of their importance in
             genetic analysis},
   Journal = {NEUROBIOLOGY OF AGING},
   Volume = {23},
   Number = {1},
   Pages = {S323-S323},
   Year = {2002},
   Key = {fds371214}
}

@article{fds276946,
   Author = {Carlson, MC and Tschanz, JT and Norton, MC and Welsh-Bohmer, K and Martin, BK and Breitner, JCS},
   Title = {H2 histamine receptor blockade in the treatment of Alzheimer
             disease: a randomized, double-blind, placebo-controlled
             trial of nizatidine.},
   Journal = {Alzheimer Dis Assoc Disord},
   Volume = {16},
   Number = {1},
   Pages = {24-30},
   Year = {2002},
   ISSN = {0893-0341},
   url = {http://dx.doi.org/10.1097/00002093-200201000-00004},
   Abstract = {OBJECTIVES: To evaluate the efficacy of nizatidine, a
             histamine H2-blocking drug, in delaying the progression of
             cognitive impairment in older adults with Alzheimer disease
             (AD). DESIGN: A one-year, randomized, double-blind,
             placebo-controlled trial. PARTICIPANTS: Fifty-one older men
             and women aged 67 to 96 years with AD were recruited from
             the Cache County Study on Memory in Aging. METHODS: Patients
             were stratified by age and by the presence of one or more
             epsilon 4 alleles at the APOE locus, then randomized to
             receive nizatidine 75 mg (Axid ARTM, Whitehall Robins) or a
             matching placebo tablet twice daily. Cognitive outcomes were
             assessed at baseline, six, and twelve months after
             enrollment using tests from the CERAD battery and additional
             measures of visuospatial memory, verbal memory, and verbal
             fluency. RESULTS: Subjects showed significant declines in
             language, fluency, and praxis but most measures of memory
             had already "bottomed out." Intention-to-treat and
             compliance-based analyses showed no effect of nizatidine on
             any of the cognitive outcome measures over the one-year
             study interval. CONCLUSIONS: These results do not support
             claims for the efficacy of nizatidine in over-the-counter
             dosages as a means of preventing symptom progression in
             AD.},
   Doi = {10.1097/00002093-200201000-00004},
   Key = {fds276946}
}

@article{fds276947,
   Author = {Schiffman, SS and Graham, BG and Sattely-Miller, EA and Zervakis, J and Welsh-Bohmer, K},
   Title = {Taste, smell and neuropsychological performance of
             individuals at familial risk for Alzheimer's
             disease.},
   Journal = {Neurobiol Aging},
   Volume = {23},
   Number = {3},
   Pages = {397-404},
   Year = {2002},
   ISSN = {0197-4580},
   url = {http://dx.doi.org/10.1016/s0197-4580(01)00337-2},
   Abstract = {The purpose of the study was to determine whether there are
             chemosensory and neuropsychological changes that predate the
             onset of Alzheimer's disease in individuals at enhanced risk
             of developing the condition. To study this question, a
             unique sample of individuals (n = 33) was studied who were
             genetically at-risk for AD by virtue of documented
             multigenerational evidence of the disease (so-called
             multiplex families). The performance of at-risk individuals
             was evaluated on various smell, taste, and
             neuropsychological measures at baseline and 18 months later.
             Their performance was compared to a control group (n = 32)
             that was matched in age, gender, education, and race. At
             baseline the at-risk group performed worse than the control
             group on the chemosensory measures of phenethyl alcohol
             smell detection, smell memory, and taste memory, and on a
             memory measure involving recall of narrative information
             (Logical Memory I from the Wechsler Memory Scale- Revised).
             Across both sessions, the at-risk group had lower smell
             memory scores than the control group. At-risk status was not
             significantly associated with APOE status. The results of
             this and other studies suggest that individuals who are
             genetically at risk for developing AD may perform more
             poorly on memory and smell measures compared to those not at
             risk. This effect may be separate from one known genetic
             risk factor of AD, APOE, and supports that multiple genes
             are likely responsible for the disease and its associated
             memory and other neurocognitive symptoms.},
   Doi = {10.1016/s0197-4580(01)00337-2},
   Key = {fds276947}
}

@article{fds277016,
   Author = {Steffens, DC and Payne, ME and Greenberg, DL and Byrum, CE and Welsh-Bohmer, KA and Wagner, HR and MacFall, JR},
   Title = {Hippocampal volume and incident dementia in geriatric
             depression.},
   Journal = {Am J Geriatr Psychiatry},
   Volume = {10},
   Number = {1},
   Pages = {62-71},
   Year = {2002},
   ISSN = {1064-7481},
   url = {http://www.ncbi.nlm.nih.gov/pubmed/11790636},
   Abstract = {The authors investigated the role of baseline hippocampal
             volume on later clinical emergence of dementia in a group of
             older, non-demented depressed individuals. Subjects were 115
             depressed, non-demented participants in a mental health
             clinical research center. All subjects were screened for
             dementia and agreed to have a magnetic resonance imaging
             (MRI) brain scan at baseline. Subjects were clinically
             evaluated by geriatric psychiatrists quarterly for up to 5
             years and received annual neuropsychological testing.
             Bivariate analyses examined age, gender, race, educational
             level, baseline depression severity, age at depression
             onset, baseline Mini-Mental State Exam (MMSE), left and
             right hippocampal volume, and total cerebral volume. Age,
             baseline MMSE, total cerebral volume, and having a small
             left hippocampal volume were associated with later dementia
             and were included in subsequent survival analysis. Small
             left hippocampal volume was significantly associated with
             later dementia (hazard ratio=2.762). Small left hippocampal
             size on neuroimaging may be a marker for dementia in
             depressed patients who have not yet met criteria for a
             clinical diagnosis of a dementing disorder.},
   Doi = {10.1176/appi.ajgp.10.1.62},
   Key = {fds277016}
}

@article{fds277006,
   Author = {Carlson, MC and Zandi, PP and Plassman, BL and Tschanz, JT and Welsh-Bohmer, KA and Steffens, DC and Bastian, LA and Mehta, KM and Breitner, JC and Cache County Study Group},
   Title = {Hormone replacement therapy and reduced cognitive decline in
             older women: the Cache County Study.},
   Journal = {Neurology},
   Volume = {57},
   Number = {12},
   Pages = {2210-2216},
   Year = {2001},
   Month = {December},
   ISSN = {0028-3878},
   url = {http://www.ncbi.nlm.nih.gov/pubmed/11756599},
   Abstract = {OBJECTIVE: To examine the association between postmenopausal
             hormone replacement therapy (HRT) and the trajectory of
             global cognitive change with age. METHODS: The Modified
             Mini-Mental State Examination (MMSE) was administered to a
             population sample of 2,073 nondemented, community-dwelling
             female residents of Cache County, UT, aged 65 and older.
             Current and past HRT and other medications at a baseline
             interview and at follow-up 3 years later were assessed.
             Between interviews, a telephone Women's Health Questionnaire
             was administered to assess initial exposure, duration, and
             recency of HRT. Generalized estimating equation marginal
             models were used to evaluate the cross-sectional and
             longitudinal relations of HRT and modified MMSE score. Also
             assessed were effects with multivitamins and calcium
             supplements as exposures likely to reflect a "healthy
             lifestyle" among HRT users. Model covariates included the
             presence of APOE epsilon4 alleles, age, education,
             concurrent depression, several chronic diseases, and
             self-perceived general health. RESULTS: Age, lower
             education, depression, and APOE epsilon4 were all associated
             with lower baseline modified MMSE scores. With these
             covariates in the model, lifetime HRT use was associated
             with better baseline modified MMSE scores and a slower rate
             of decline. Stratification by APOE genotype did not alter
             these effects. Apparent benefits with HRT were attenuated
             but remained significant after elimination of scores from
             participants with incident dementia. A significant
             interaction between age and HRT indicated the strongest
             effects in women aged 85 and older. Measures of age at
             initial use of HRT, duration, and recency of exposure did
             not improve the models. No effects were seen with the
             "healthy lifestyle" control exposures. CONCLUSIONS: In a
             population cohort of older women, lifetime HRT exposure was
             associated with improved global cognition and attenuated
             decline over a 3-year interval. Improvements were greatest
             in the oldest old.},
   Doi = {10.1212/wnl.57.12.2210},
   Key = {fds277006}
}

@article{fds276945,
   Author = {Silverman, DH and Small, GW and Chang, CY and Lu, CS and Kung De Aburto,
             MA and Chen, W and Czernin, J and Rapoport, SI and Pietrini, P and Alexander, GE and Schapiro, MB and Jagust, WJ and Hoffman, JM and Welsh-Bohmer, KA and Alavi, A and Clark, CM and Salmon, E and de Leon,
             MJ and Mielke, R and Cummings, JL and Kowell, AP and Gambhir, SS and Hoh,
             CK and Phelps, ME},
   Title = {Positron emission tomography in evaluation of dementia:
             Regional brain metabolism and long-term outcome.},
   Journal = {JAMA},
   Volume = {286},
   Number = {17},
   Pages = {2120-2127},
   Year = {2001},
   Month = {November},
   ISSN = {0098-7484},
   url = {http://www.ncbi.nlm.nih.gov/pubmed/11694153},
   Abstract = {CONTEXT: Deficits in cerebral glucose utilization have been
             identified in patients with cognitive dysfunction attributed
             to various disease processes, but their prognostic and
             diagnostic value remains to be defined. OBJECTIVE: To assess
             the sensitivity and specificity with which cerebral
             metabolic patterns at a single point in time forecast
             subsequent documentation of progressive dementia. DESIGN,
             SETTING, AND PATIENTS: Positron emission tomography (PET)
             studies of [(18)F]fluorodeoxyglucose in 146 patients
             undergoing evaluation for dementia with at least 2 years'
             follow-up for disease progression at the University of
             California, Los Angeles, from 1991 to 2000, and PET studies
             in 138 patients undergoing evaluation for dementia at an
             international consortium of facilities, with
             histopathological diagnoses an average of 2.9 years later,
             conducted from 1984 to 2000. MAIN OUTCOME MEASURES: Regional
             distribution of [(18)F]fluorodeoxyglucose in each patient,
             classified by criteria established a priori as positive or
             negative for presence of a progressive neurodegenerative
             disease in general and of Alzheimer disease (AD)
             specifically, compared with results of longitudinal or
             neuropathologic analyses. RESULTS: Progressive dementia was
             detected by PET with a sensitivity of 93% (191/206) and a
             specificity of 76% (59/78). Among patients with
             neuropathologically based diagnoses, PET identified patients
             with AD and patients with any neurodegenerative disease with
             a sensitivity of 94% and specificities of 73% and 78%,
             respectively. The negative likelihood ratio of experiencing
             a progressive vs nonprogressive course over the several
             years following a single negative brain PET scan was 0.10
             (95% confidence interval, 0.06-0.16), and the initial
             pattern of cerebral metabolism was significantly associated
             with the subsequent course of progression overall (P<.001).
             CONCLUSION: In patients presenting with cognitive symptoms
             of dementia, regional brain metabolism was a sensitive
             indicator of AD and of neurodegenerative disease in general.
             A negative PET scan indicated that pathologic progression of
             cognitive impairment during the mean 3-year follow-up was
             unlikely to occur.},
   Doi = {10.1001/jama.286.17.2120},
   Key = {fds276945}
}

@article{fds276944,
   Author = {Sugarman, J and Cain, C and Wallace, R and Welsh-Bohmer,
             KA},
   Title = {How proxies make decisions about research for patients with
             Alzheimer's disease.},
   Journal = {J Am Geriatr Soc},
   Volume = {49},
   Number = {8},
   Pages = {1110-1119},
   Year = {2001},
   Month = {August},
   ISSN = {0002-8614},
   url = {http://www.ncbi.nlm.nih.gov/pubmed/11555076},
   Abstract = {We examined the proxy decision-making and informed consent
             processes for clinical research involving 49
             patient-subjects with dementia in an outpatient setting by
             performing serial in-depth, structured, open-ended telephone
             interviews. Interviews were tape recorded and transcribed.
             Transcripts were then coded and analyzed. Although in all
             cases proxy consent was obtained for research from a legally
             authorized representative, proxies reported considerable
             ambiguity regarding who made the decision to participate in
             research, or to what degree the decision was that of the
             proxy or of the patient. Reasons proxies gave for
             participating in research included: hope of direct or
             indirect benefits to the patient, caregiver, or patient's
             descendents; desperation; trust in the investigator; belief
             in the goodness of research; and altruism. These reasons
             varied according to the type of research. For instance, in
             drug trials hope of direct benefit prevailed; in studies not
             evaluating a potential therapy more altruistic concerns
             predominated. Being a proxy decision maker for research can
             be burdensome. The degree of burden related to making a
             decision to participate in research seems influenced by a
             number of intersecting factors, most importantly, the risk
             and nature of the study, the extent to which patients were
             able to participate in the decision, and the duration and
             severity of dementia. Proxy decision-making concerning
             participation in research for patients with dementia can be
             a difficult task. The process might be improved by
             emphasizing that proxy consent is being sought because the
             nature of the patient's underlying medical condition can
             preclude the ability to make meaningful decisions. In
             addition, clinical researchers should recognize that giving
             proxy consent might place additional burdens on caregivers
             and discuss this explicitly when proxy consent is
             solicited.},
   Doi = {10.1046/j.1532-5415.2001.49218.x},
   Key = {fds276944}
}

@article{fds277116,
   Author = {Koltai, DC and Welsh-Bohmer, KA and Schmechel,
             DE},
   Title = {Influence of anosognosia on treatment outcome among dementia
             patients},
   Journal = {Neuropsychological Rehabilitation},
   Volume = {11},
   Number = {3-4},
   Pages = {455-475},
   Publisher = {Informa UK Limited},
   Year = {2001},
   Month = {July},
   ISSN = {0960-2011},
   url = {http://dx.doi.org/10.1080/09602010042000097},
   Abstract = {This study was a preliminary investigation of the effects of
             a Memory and Coping Program among mild to moderate dementia
             patients. A total of 24 elderly participants were randomly
             assigned to treatment and waiting-list control conditions. A
             pre-test, post-test design was used, with group comparisons
             of change scores on objective cognitive tests and subjective
             ratings of mood and memory. While encouraging trends emerged
             suggesting improvement among those who received treatment,
             group differences did not reach statistical significance.
             However, when outcomes were compared by treatment subgroups
             with and without anosognosia, significant differences
             emerged. Participants with insight made significantly
             greater gains in perceived memory functioning that those
             without insight. In contrast, informants perceived greater
             gains among treatment subjects relative to controls
             independent of insight status. The need for additional
             research to further delineate the influences of anosognia,
             baseline cognition, and affective status, and other
             potential intervention outcome modifiers is discussed, with
             attention to instrumentation and methodological
             considerations.},
   Doi = {10.1080/09602010042000097},
   Key = {fds277116}
}

@article{fds372627,
   Author = {Grundman, M and Kim, HT and Salmon, D and Storandt, M and Smith, G and Ferris, S and Mohs, R and Brandt, J and Doody, R and Welsh‐Bohmer, K and Saxton, J and Saine, K and Schmitt, F and Ogrocki, P and Johnson, N and Howieson, D and Papka, M and Green, J and Gamst, A and Kukull, W and Thal,
             LJ},
   Title = {The Alzheimer's Disease Centers' Neuropsychological Database
             Initiative: A Resource for Alzheimer's Disease Prevention
             Trials},
   Pages = {129-140},
   Publisher = {Wiley},
   Year = {2001},
   Month = {March},
   ISBN = {9780471521761},
   url = {http://dx.doi.org/10.1002/0470846453.ch13},
   Doi = {10.1002/0470846453.ch13},
   Key = {fds372627}
}

@article{fds372476,
   Author = {Welsh‐Bohmer, KA and Hulette, C and Schmechel, D and Burke, J and Saunders, A},
   Title = {Neuropsychological Detection of Preclinical Alzheimer's
             Disease: Results of a Neuropathological Series of
             ‘Normal’ Controls},
   Pages = {111-122},
   Publisher = {Wiley},
   Year = {2001},
   Month = {March},
   ISBN = {9780471521761},
   url = {http://dx.doi.org/10.1002/0470846453.ch11},
   Doi = {10.1002/0470846453.ch11},
   Key = {fds372476}
}

@article{fds277015,
   Author = {Steffens, DC and MacFall, JR and Payne, ME and Welsh-Bohmer, KA and Krishnan, KR},
   Title = {Grey-matter lesions and dementia.},
   Journal = {Lancet},
   Volume = {356},
   Number = {9242},
   Pages = {1686-1687},
   Year = {2000},
   Month = {November},
   ISSN = {0140-6736},
   url = {http://www.ncbi.nlm.nih.gov/pubmed/11089848},
   Doi = {10.1016/S0140-6736(05)70393-7},
   Key = {fds277015}
}

@article{fds277055,
   Author = {Hoffman, JM and Welsh-Bohmer, KA and Hanson, M and Crain, B and Hulette,
             C and Earl, N and Coleman, RE},
   Title = {FDG PET imaging in patients with pathologically verified
             dementia.},
   Journal = {J Nucl Med},
   Volume = {41},
   Number = {11},
   Pages = {1920-1928},
   Year = {2000},
   Month = {November},
   ISSN = {0161-5505},
   url = {http://www.ncbi.nlm.nih.gov/pubmed/11079505},
   Abstract = {UNLABELLED: The purpose of this study was to confirm with
             pathologic verification 2 beliefs related to Alzheimer's
             disease (AD): (a) the long-standing impression that
             bilateral temporo-parietal hypometabolism, as noted on FDG
             PET imaging, is the metabolic abnormality associated with
             Alzheimer's disease (AD) and (b) that the sensitivity,
             specificity, and diagnostic accuracy of the metabolic
             pattern of bilateral temporo-parietal hypometabolism allows
             differentiation between other degenerative causes of
             dementia. METHODS: Twenty two individuals (8 women, 14 men)
             with difficult-to-characterize memory loss or dementia
             (using standard clinical criteria), and who eventually
             received pathologic confirmation of diagnosis, were
             evaluated. FDG PET brain scans were obtained and visually
             graded by an experienced nuclear medicine physician as to
             the presence of classic bilateral temporo-parietal
             hypometabolism as seen in Alzheimer's type dementia.
             Sensitivity, specificity, positive predictive value,
             negative predictive value, and diagnostic accuracy of the
             metabolic pattern of bilateral temporo-parietal
             hypometabolism were determined using pathologic diagnosis as
             the gold standard. RESULTS: The clinical diagnosis of
             possible or probable AD was determined as the primary cause
             of dementia in 12 patients. The sensitivity and specificity
             of the clinical diagnosis for probable AD were 63% and 100%,
             respectively. The sensitivity and specificity of the
             clinical diagnosis for possible and probable AD were 75% and
             100%, respectively. The sensitivity, specificity, and
             diagnostic accuracy of bilateral temporo-parietal
             hypometabolism being associated with AD were 93%, 63%, and
             82%, respectively. CONCLUSION: This study confirms that
             bilateral temporo-parietal hypometabolism is indeed the
             classic metabolic abnormality associated with AD.
             Furthermore, in individuals with dementia whose FDG PET
             scans indicated a metabolic pattern other than bilateral
             temporo-parietal hypometabolism, a cause of dementia other
             than AD should be suspected. These observations may be of
             clinical importance in differentiating dementia syndromes.
             The sensitivity, specificity, and diagnostic accuracy of FDG
             PET are acceptable as tests to be used in the evaluation of
             dementia and particularly to confirm the clinical suspicion
             of AD.},
   Key = {fds277055}
}

@article{fds277076,
   Author = {Plassman, BL and Havlik, RJ and Steffens, DC and Helms, MJ and Newman,
             TN and Drosdick, D and Phillips, C and Gau, BA and Welsh-Bohmer, KA and Burke, JR and Guralnik, JM and Breitner, JC},
   Title = {Documented head injury in early adulthood and risk of
             Alzheimer's disease and other dementias.},
   Journal = {Neurology},
   Volume = {55},
   Number = {8},
   Pages = {1158-1166},
   Year = {2000},
   Month = {October},
   ISSN = {0028-3878},
   url = {http://www.ncbi.nlm.nih.gov/pubmed/11071494},
   Abstract = {BACKGROUND: The association between antecedent head injury
             and AD is inconsistent. OBJECTIVE: To examine the
             association between early adult head injury, as documented
             by military hospital records, and dementia in late life; and
             to evaluate the interaction between head injury and APOE
             epsilon4 as risk factors for dementia. METHODS: The study
             had a population-based prospective historical cohort design.
             It included men who were World War II Navy and Marine
             veterans, and were hospitalized during their military
             service with a diagnosis of either a nonpenetrating head
             injury or another unrelated condition. In 1996 to 1997,
             military medical records were abstracted to document the
             occurrence and details of closed head injury. The entire
             sample was then evaluated for dementia and AD using a
             multistage procedure. There were 548 veterans with head
             injury and 1228 without head injury who completed all
             assigned stages of the study. The authors estimated risk of
             dementia, specifically AD, using proportional hazards
             models. RESULTS: Both moderate head injury (hazard ratio
             [HR] = 2.32; CI = 1.04 to 5.17) and severe head injury (HR =
             4.51; CI = 1.77 to 11.47) were associated with increased
             risk of AD. Results were similar for dementia in general.
             The results for mild head injury were inconclusive. When the
             authors stratified by the number of APOE epsilon4 alleles,
             they observed a nonsignificant trend toward a stronger
             association between AD and head injury in men with more
             epsilon4 alleles. CONCLUSIONS: Moderate and severe head
             injuries in young men may be associated with increased risk
             of AD and other dementias in late life. However, the authors
             cannot exclude the possibility that other unmeasured factors
             may be influencing this association.},
   Doi = {10.1212/wnl.55.8.1158},
   Key = {fds277076}
}

@article{fds277004,
   Author = {Steffens, DC and Skoog, I and Norton, MC and Hart, AD and Tschanz, JT and Plassman, BL and Wyse, BW and Welsh-Bohmer, KA and Breitner,
             JC},
   Title = {Prevalence of depression and its treatment in an elderly
             population: the Cache County study.},
   Journal = {Arch Gen Psychiatry},
   Volume = {57},
   Number = {6},
   Pages = {601-607},
   Year = {2000},
   Month = {June},
   ISSN = {0003-990X},
   url = {http://www.ncbi.nlm.nih.gov/pubmed/10839339},
   Abstract = {BACKGROUND: Previous estimates of the prevalence of
             geriatric depression have varied. There are few large
             population-based studies; most of these focused on
             individuals younger than 80 years. No US studies have been
             published since the advent of the newer antidepressant
             agents. METHODS: In 1995 through 1996, as part of a large
             population study, we examined the current and lifetime
             prevalence of depressive disorders in 4,559 nondemented
             individuals aged 65 to 100 years. This sample represented
             90% of the elderly population of Cache County, Utah. Using a
             modified version of the Diagnostic Interview Schedule, we
             ascertained past and present DSM-IV major depression,
             dysthymia, and subclinical depressive disorders. Medication
             use was determined through a structured interview and a
             "medicine chest inventory." RESULTS: Point prevalence of
             major depression was estimated at 4.4% in women and 2.7% in
             men (P= .003). Other depressive syndromes were surprisingly
             uncommon (combined point prevalence, 1.6%). Among subjects
             with current major depression, 35.7% were taking an
             antidepressant (mostly selective serotonin reuptake
             inhibitors) and 27.4% a sedative/hypnotic. The current
             prevalence of major depression did not change appreciably
             with age. Estimated lifetime prevalence of major depression
             was 20.4% in women and 9.6% in men (P<.001), decreasing with
             age. CONCLUSIONS: These estimates for prevalence of major
             depression are higher than those reported previously in
             North American studies. Treatment with antidepressants was
             more common than reported previously, but was still lacking
             in most individuals with major depression. The prevalence of
             subsyndromal depressive symptoms was low, possibly because
             of unusual characteristics of the population.},
   Doi = {10.1001/archpsyc.57.6.601},
   Key = {fds277004}
}

@article{fds276943,
   Author = {Tschanz, JT and Welsh-Bohmer, KA and Skoog, I and West, N and Norton,
             MC and Wyse, BW and Nickles, R and Breitner, JC},
   Title = {Dementia diagnoses from clinical and neuropsychological data
             compared: the Cache County study.},
   Journal = {Neurology},
   Volume = {54},
   Number = {6},
   Pages = {1290-1296},
   Year = {2000},
   Month = {March},
   ISSN = {0028-3878},
   url = {http://www.ncbi.nlm.nih.gov/pubmed/10746600},
   Abstract = {OBJECTIVE: To validate a neuropsychological algorithm for
             dementia diagnosis. METHODS: We developed a
             neuropsychological algorithm in a sample of 1,023 elderly
             residents of Cache County, UT. We compared algorithmic and
             clinical dementia diagnoses both based on DSM-III-R
             criteria. The algorithm diagnosed dementia when there was
             impairment in memory and at least one other cognitive
             domain. We also tested a variant of the algorithm that
             incorporated functional measures that were based on
             structured informant reports. RESULTS: Of 1,023
             participants, 87% could be classified by the basic
             algorithm, 94% when functional measures were considered.
             There was good concordance between basic psychometric and
             clinical diagnoses (79% agreement, kappa = 0.57). This
             improved after incorporating functional measures (90%
             agreement, kappa = 0.76). CONCLUSIONS: Neuropsychological
             algorithms may reasonably classify individuals on dementia
             status across a range of severity levels and ages and may
             provide a useful adjunct to clinical diagnoses in population
             studies.},
   Doi = {10.1212/wnl.54.6.1290},
   Key = {fds276943}
}

@article{fds277005,
   Author = {Steffens, DC and Plassman, BL and Helms, MJ and Welsh-Bohmer, KA and Newman, TT and Breitner, JC},
   Title = {APOE and AD concordance in twin pairs as predictors of AD in
             first-degree relatives.},
   Journal = {Neurology},
   Volume = {54},
   Number = {3},
   Pages = {593-598},
   Year = {2000},
   Month = {February},
   ISSN = {0028-3878},
   url = {http://www.ncbi.nlm.nih.gov/pubmed/10680788},
   Abstract = {OBJECTIVE: To examine the independent effects of the APOE
             genotype (APOE) and concordance for AD in twin pairs on the
             occurrence of AD in first-degree relatives. BACKGROUND:
             Studies of twins have been undertaken to investigate the
             influence of genes in a variety of conditions, including AD.
             A previous study, performed before reports linking APOE to
             AD, demonstrated an increase in AD among first-degree
             relatives of twins concordant for AD compared with relatives
             of discordant twins. METHODS: In a sample of 94 twin pairs
             the authors examined the association between concordance for
             AD within the twin pair and family history of AD among
             first-degree relatives of twins. They then examined the
             extent to which the presence of the APOE epsilon4 allele in
             the twin pair explains the association between concordance
             for AD within the twin pair and family history of AD.
             RESULTS: Concordance among twins was associated with
             increased risk of AD among relatives (logrank test, chi2 =
             12.558; p = 0.0004), and the presence of at least one APOE
             epsilon4 allele in each member of the twin pair is also
             associated with increased risk of AD among family members
             (logrank test, chi2 = 7.712; p = 0.0055). CONCLUSIONS: APOE
             genotype explains much but not all of the association
             between concordance among twins and increased familial risk
             of AD.},
   Doi = {10.1212/wnl.54.3.593},
   Key = {fds277005}
}

@article{fds277109,
   Author = {Clark, LM and Bosworth, HB and Welsh-Bohmer, KA and Dawson, DV and Siegler, IC},
   Title = {Relation between informant-rated personality and
             clinician-rated depression in patients with memory
             disorders.},
   Journal = {Neuropsychiatry Neuropsychol Behav Neurol},
   Volume = {13},
   Number = {1},
   Pages = {39-47},
   Year = {2000},
   Month = {January},
   ISSN = {0894-878X},
   url = {http://www.ncbi.nlm.nih.gov/pubmed/10645735},
   Abstract = {OBJECTIVE: The goal of this study was to examine the
             convergent validity of informant-rated changes in depressive
             and related personality traits with clinician-assessed
             depression in memory-disordered patients. BACKGROUND:
             Depressive symptoms are frequent complications in persons
             with dementias such as Alzheimer disease, and caregiver
             informants consistently report changes in depression and
             related neurotic traits on the NEO Personality Inventory
             (NEO-PI) in dementia patients. METHODS: In 78 patients
             undergoing evaluation of memory complaints at an Alzheimer
             disease clinic, depression was characterized by clinical
             diagnosis, a clinician-rated scale, and informant ratings of
             premorbid versus current depression, anxiety, vulnerability,
             and neuroticism on the NEO-PI. RESULTS: The diagnostic
             groups differed in meaningful patterns on the NEO-PI
             measures. Those with a diagnosis of major depression
             differed from never-depressed patients in all personality
             areas, although those with depressed mood differed only on
             NEO-PI depression. The clinician-rated depression scale
             correlated modestly with current personality and change from
             baseline personality. CONCLUSIONS: The NEO-PI provides a
             useful measure of informants' perspectives on depressive
             personality changes in patients with memory disorders but
             does not correspond fully with a clinical syndrome of
             depression.},
   Key = {fds277109}
}

@article{fds276985,
   Author = {Dawson, DV and Welsh-Bohmer, KA and Siegler, IC},
   Title = {Premorbid personality predicts level of rated personality
             change in patients with Alzheimer disease.},
   Journal = {Alzheimer Dis Assoc Disord},
   Volume = {14},
   Number = {1},
   Pages = {11-19},
   Year = {2000},
   ISSN = {0893-0341},
   url = {http://www.ncbi.nlm.nih.gov/pubmed/10718200},
   Abstract = {Multiple studies of individuals with Alzheimer disease have
             substantiated significant levels of informant-rated change
             in several domains and facets of the Neuroticism-Extraversion-Openness
             Personality Inventory, including increases in Neuroticism
             and decreases in Extraversion and Conscientiousness relative
             to premorbid personality traits. Decline in Openness was
             cited in some reports, and replicable changes were
             identified in several facets. Current and premorbid
             personality of 50 patients with Alzheimer disease were rated
             by informants using the Neuroticism-Extraversion-Openness
             Personality Inventory. Multiple regression analysis was used
             to assess possible relationships of levels of reported
             change with covariates, including premorbid rating,
             education, duration of dementia, age, gender, and
             Mini-Mental State Examination score. Premorbid rating was
             the only significant predictor of reported change for
             Neuroticism, Extraversion, Conscientiousness, and the facets
             Anxiety (N1), Assertiveness (E3), and Activity (E4). Rated
             change in Depression was also found to be related to
             duration of dementia, change in Vulnerability was influenced
             by gender, and reported change in both Openness and Ideas
             showed a relationship to level of education.},
   Doi = {10.1097/00002093-200001000-00002},
   Key = {fds276985}
}

@article{fds371215,
   Author = {Yuspeh, RL and Koltai, DC and Welsh-Bohmer, KA},
   Title = {Learning over trials: Normative data for an index of verbal
             learning for the CERAD Word List Learning
             Task},
   Journal = {Archives of Clinical Neuropsychology},
   Volume = {14},
   Number = {8},
   Pages = {670-670},
   Publisher = {Oxford University Press (OUP)},
   Year = {1999},
   Month = {November},
   url = {http://dx.doi.org/10.1093/arclin/14.8.670},
   Doi = {10.1093/arclin/14.8.670},
   Key = {fds371215}
}

@article{fds277013,
   Author = {Welsh-Bohmer, KA and Morgenlander, JC},
   Title = {Determining the cause of memory loss in the elderly. From
             in-office screening to neuropsychological
             referral.},
   Journal = {Postgrad Med},
   Volume = {106},
   Number = {5},
   Pages = {99-passim},
   Year = {1999},
   Month = {October},
   ISSN = {0032-5481},
   url = {http://www.ncbi.nlm.nih.gov/pubmed/10560471},
   Abstract = {Improved understanding of neurobehavior in normal aging,
             Alzheimer's disease, and late-life depression makes early
             detection of neurodegenerative conditions possible. Primary
             care physicians can screen patients' mental status and mood
             states with simple in-office tests. When screening results
             or the clinical picture is ambiguous or complex,
             neuropsychological evaluation is useful in making an early,
             reliable differentiation between dementia and normal aging.
             Early identification of neurologic problems provides an
             opportunity to enhance quality of life and long-term care.
             Medical interventions, such as a trial of donepezil
             hydrochloride (Aricept) or other memory-enhancing
             medications as they become available, can be started when
             results are likely to be optimal. Common coexisting problems
             (e.g., depression, falls) can be sought and managed.
             Additional important medical decisions (e.g., elective
             surgeries) may be considered differently when dementia is
             diagnosed early.},
   Doi = {10.3810/pgm.1999.10.15.747},
   Key = {fds277013}
}

@article{fds277002,
   Author = {Norton, MC and Tschanz, JA and Fan, X and Plassman, BL and Welsh-Bohmer,
             KA and West, N and Wyse, BW and Breitner, JC},
   Title = {Telephone adaptation of the Modified Mini-Mental State Exam
             (3MS). The Cache County Study.},
   Journal = {Neuropsychiatry Neuropsychol Behav Neurol},
   Volume = {12},
   Number = {4},
   Pages = {270-276},
   Year = {1999},
   Month = {October},
   ISSN = {0894-878X},
   url = {http://www.ncbi.nlm.nih.gov/pubmed/10527112},
   Abstract = {OBJECTIVE: To examine the concurrent validity of a newly
             developed telephone adaptation of the Modified Mini-Mental
             State Exam. BACKGROUND: Longitudinal studies of cognition
             may be advantaged by availability of assessment instruments
             that can be used over the telephone, as well as in person.
             METHOD: Subjects were 263 noninstitutionalized elderly
             residents of a rural community in southern Idaho, aged 65 to
             93, who had little or no cognitive difficulty. At an average
             interval of four weeks, we administered the Modified
             Mini-Mental State Exam (3MS) and the newly adapted Telephone
             Modified Mini-Mental State Exam (T3MS). Order of
             administration was randomly assigned. RESULTS: Agreement
             between scores on the two instruments was good (r = 0.82, p
             < 0.001). When we applied various cutoff scores to the
             instruments, thereby generating assignments of individuals
             to "screen positive" and "screen negative" groups, the
             percent agreement in screening results ranged from 80% to
             96% as we reduced the cutoff scores from 90 to 74 (100
             points possible). CONCLUSIONS: At least among subjects
             without major cognitive syndromes, the Telephone Modified
             Mini-Mental State Exam provides a reasonable substitute for
             the more costly in-person 3MS. The telephone instrument
             should now be tested over a broader range of cognitive
             abilities in order to assess its validity in more impaired
             subjects, e.g., by studying an institutionalized
             sample.},
   Key = {fds277002}
}

@article{fds277003,
   Author = {Steffens, DC and Norton, MC and Plassman, BL and Tschanz, JT and Wyse,
             BW and Welsh-Bohmer, KA and Anthony, JC and Breitner,
             JC},
   Title = {Enhanced cognitive performance with estrogen use in
             nondemented community-dwelling older women.},
   Journal = {J Am Geriatr Soc},
   Volume = {47},
   Number = {10},
   Pages = {1171-1175},
   Year = {1999},
   Month = {October},
   ISSN = {0002-8614},
   url = {http://www.ncbi.nlm.nih.gov/pubmed/10522948},
   Abstract = {OBJECTIVE: To examine the association between history of
             postmenopausal estrogen use and cognitive function in a
             large sample of nondemented community-dwelling older women.
             SETTING: A community of older residents in Cache County,
             Utah. PARTICIPANTS: A total of 2338 nondemented women aged
             65 and older. MEASUREMENTS: All subjects were administered
             the Modified Mini-Mental State Examination (3MSE).
             Self-reported information on current and past use of
             estrogen after menopause was also obtained using a
             structured interview. Estrogen use was trichotomized as: no
             use, past use, and current use. Apolipoprotein E (APOE)
             genotype was determined and was dichotomized by the presence
             of an epsilon4 allele. A series of variance/covariance
             models was conducted with the 3MSE score as the dependent
             variable, first considering estrogen use alone, then adding,
             sequentially as covariates, education, age, health status,
             APOE genotype, current depression status, and history of
             head injury. RESULTS: In the simplest bivariate model, the
             3MSE means (and confidence intervals) were 92.1 (91.7-92.4),
             93.5 (93.1-93.9), and 94.4 (94.0-94.7) for never-, past-,
             and current users, respectively. In the final model (R2 =
             0.28), no use of estrogen replacement therapy (P = .006),
             lower education (P < .001), poorer perceived health status
             (P = .035), current depression (P = .014), and presence of
             at least one APOE epsilon4 allele (P < .001) each
             independently predicted lower 3MSE score. Both current and
             past estrogen users had significantly higher 3MSE scores
             than never-users (P = .0063 and P = .0096, respectively).
             CONCLUSIONS: In this large community study, women who had
             used estrogen after menopause scored higher on the 3MSE.
             This finding remained, even after controlling for the
             effects of age, education, APOE genotype, and other
             variables that may affect cognition. These data support
             studies reporting a beneficial role of estrogen on cognition
             in postmenopausal women, particularly among current estrogen
             users.},
   Doi = {10.1111/j.1532-5415.1999.tb05195.x},
   Key = {fds277003}
}

@article{fds277020,
   Author = {Hulette, CM and Pericak-Vance, MA and Roses, AD and Schmechel, DE and Yamaoka, LH and Gaskell, PC and Welsh-Bohmer, KA and Crowther, RA and Spillantini, MG},
   Title = {Neuropathological features of frontotemporal dementia and
             parkinsonism linked to chromosome 17q21-22 (FTDP-17): Duke
             Family 1684.},
   Journal = {J Neuropathol Exp Neurol},
   Volume = {58},
   Number = {8},
   Pages = {859-866},
   Year = {1999},
   Month = {August},
   ISSN = {0022-3069},
   url = {http://www.ncbi.nlm.nih.gov/pubmed/10446810},
   Abstract = {Frontotemporal dementia with parkinsonism (FTDP-17) is an
             autosomal dominant disorder that presents clinically with
             dementia, extrapyramidal signs, and behavioral disturbances
             in mid-life and progresses to death within 5 to 10 years.
             Pathologically, the disorder is characterized by variable
             neuronal loss and gliosis in the frontal and temporal lobes,
             limbic structures, and the midbrain. Autopsied individuals
             from some kindreds display abundant neurofibrillary change
             while others, including a single affected individual from
             Duke Family 1684, lack distinctive histological features and
             exhibit only mild neuronal loss and gliosis in limbic
             structures and subcortical nuclei when examined by routine
             silver stain. Recently, mutations in the microtubule
             associated protein tau have been shown to segregate with the
             disease in this family and in many other affected kindreds.
             In order to examine the distribution of tau deposits, we
             performed tau immunohistochemistry, immunoblotting, and
             immunoelectron microscopy of tau-containing filaments.
             Immunohistochemistry revealed numerous tau deposits within
             glial cells and within neurons. Twisted ribbon-like
             filaments observed by immunoelectron microscopy were
             immunodecorated with tau AT8 antibody. Sarkosyl-insoluble
             tau extracted from the hippocampus and cortex migrated as 2
             major bands at 64 and 68 kilodaltons and a minor band at 72
             kilodaltons, which after alkaline phosphatase treatment
             appeared to contain mainly tau isoforms with 4 repeats.
             Furthermore, the ratio of soluble tau with 4 to 3
             microtubule-binding repeats was increased. The role of tau
             mutations in this disorder is discussed in this
             paper.},
   Doi = {10.1097/00005072-199908000-00008},
   Key = {fds277020}
}

@article{fds277000,
   Author = {Breitner, JC and Wyse, BW and Anthony, JC and Welsh-Bohmer, KA and Steffens, DC and Norton, MC and Tschanz, JT and Plassman, BL and Meyer,
             MR and Skoog, I and Khachaturian, A},
   Title = {APOE-epsilon4 count predicts age when prevalence of AD
             increases, then declines: the Cache County
             Study.},
   Journal = {Neurology},
   Volume = {53},
   Number = {2},
   Pages = {321-331},
   Year = {1999},
   Month = {July},
   ISSN = {0028-3878},
   url = {http://dx.doi.org/10.1212/wnl.53.2.321},
   Abstract = {OBJECTIVE: To examine the prevalence of Alzheimer's disease
             (AD) and other dementias in relation to age, education, sex,
             and genotype at APOE. Recent studies suggest age
             heterogeneity in the risk of AD associated with the APOE
             genotype and a possible interaction between APOE-epsilon4
             and female sex as risk factors. We studied these topics in
             the 5,677 elderly residents of Cache County, Utah, a
             population known for long life expectancy and high
             participation rates. METHODS: We screened for dementia with
             a brief cognitive test and structured telephone Dementia
             Questionnaire, then examined all individuals with apparent
             cognitive symptoms and a sample of others. We estimated
             age-specific prevalence of AD and other dementias and used
             multiple logistic regression models to describe relation of
             AD prevalence to age, sex, education, and APOE genotype.
             RESULTS: We found 335 demented individuals, 230 (69%) with
             definite, probable, or possible AD (positive predictive
             value versus autopsy confirmation 85%). The adjusted
             prevalence estimate for AD was 6.5% and for all dementias
             9.6%. After age 90, the adjusted prevalence estimate for AD
             was 28% and for all dementias 38%. Regression models showed
             strong variation in AD prevalence with age, sex, education,
             and number of epsilon4 alleles (effect of epsilon2 not
             significant). Models were improved by a term for age-squared
             (negative coefficient) and by separate terms for interaction
             of age with presence of one or two epsilon4 alleles. An
             association of AD with female sex was ascribable entirely to
             individuals with epsilon4. CONCLUSIONS: In participants with
             no epsilon4 alleles, the age-specific prevalence of AD
             reached a maximum and then declined after age 95. In
             epsilon4 heterozygotes a similar maximum was noted earlier
             at age 87, in homozygotes at age 73. Female sex was a risk
             factor for AD only in those with epsilon4. The epsilon4
             allele accounted for 70% of the population attributable risk
             for AD.},
   Doi = {10.1212/wnl.53.2.321},
   Key = {fds277000}
}

@article{fds277061,
   Author = {Heyman, A and Fillenbaum, GG and Gearing, M and Mirra, SS and Welsh-Bohmer, KA and Peterson, B and Pieper, C},
   Title = {Comparison of Lewy body variant of Alzheimer's disease with
             pure Alzheimer's disease: Consortium to Establish a Registry
             for Alzheimer's Disease, Part XIX.},
   Journal = {Neurology},
   Volume = {52},
   Number = {9},
   Pages = {1839-1844},
   Year = {1999},
   Month = {June},
   ISSN = {0028-3878},
   url = {http://www.ncbi.nlm.nih.gov/pubmed/10371532},
   Abstract = {OBJECTIVE: To compare the clinical, neuropsychological, and
             neuropathologic findings in patients with AD alone with
             those in patients with the Lewy body variant of AD (LBV).
             BACKGROUND: Prior studies indicate that patients with LBV
             not only have distinct clinical and neuropsychological
             differences from those with AD alone, but have a poorer
             prognosis with shorter survival time. METHODS: The authors
             evaluated 74 patients with autopsy-confirmed AD alone and 27
             patients with LBV, and compared demographic characteristics
             and clinical, neuropsychological, and neuropathologic
             findings. RESULTS: The two groups of patients were
             equivalent with respect to age at time of entry into the
             study, years of education, and sex. Two or more
             extrapyramidal clinical manifestations were found in 44% of
             patients with LBV, compared with 16% of patients with AD
             alone (p = 0.02). Duration of survival after entry into the
             study was similar in both groups, with a mean survival of
             3.6 (+/-2.1) years for AD alone versus 3.8 (+/-1.9) years
             for LBV. Of the various neuropsychological tests
             administered at the last Consortium to Establish a Registry
             for Alzheimer's Disease evaluation, only delayed recall of a
             learned word list was significantly different in the two
             groups, with 32% of patients with LBV versus 15% of patients
             with AD alone recalling any items (p = 0.04).
             Neuropathologic findings confirmed those of previous studies
             and showed that neurofibrillary tangles were significantly
             less frequent in the neocortex of patients with LBV than in
             those with AD alone. CONCLUSION: Compared with patients with
             AD alone, those with LBV had a greater frequency of
             extrapyramidal manifestations, somewhat better recall on a
             selected memory task at their final evaluation, and a
             significantly lower frequency of neocortical neurofibrillary
             tangles at autopsy. There were no differences between the
             two groups, however, in survival time from entry into the
             study.},
   Doi = {10.1212/wnl.52.9.1839},
   Key = {fds277061}
}

@article{fds276987,
   Author = {Watson, ME and Welsh-Bohmer, KA and Hoffman, JM and Lowe, V and Rubin,
             DC},
   Title = {The neural basis of naming impairments in Alzheimer's
             disease revealed through positron emission
             tomography.},
   Journal = {Arch Clin Neuropsychol},
   Volume = {14},
   Number = {4},
   Pages = {347-357},
   Year = {1999},
   Month = {May},
   ISSN = {0887-6177},
   url = {http://www.ncbi.nlm.nih.gov/pubmed/14590589},
   Abstract = {The naming impairments in Alzheimer's disease (AD) have been
             attributed to a variety of cognitive processing deficits,
             including impairments in semantic memory, visual perception,
             and lexical access. To further understand the underlying
             biological basis of the naming failures in AD, the present
             investigation examined the relationship of various classes
             of naming errors to regional brain measures of cerebral
             glucose metabolism as measured with 18 F-Fluoro-2-deoxyglucose
             (FDG) and positron emission tomography (PET). Errors
             committed on a visual naming test were categorized according
             to a cognitive processing schema and then examined in
             relationship to metabolism within specific brain regions.
             The results revealed an association of semantic errors with
             glucose metabolism in the frontal and temporal regions.
             Language access errors, such as circumlocutions, and word
             blocking nonresponses were associated with decreased
             metabolism in areas within the left hemisphere.
             Visuoperceptive errors were related to right inferior
             parietal metabolic function. The findings suggest that
             specific brain areas mediate the perceptual, semantic, and
             lexical processing demands of visual naming and that visual
             naming problems in dementia are related to dysfunction in
             specific neural circuits.},
   Doi = {10.1016/s0887-6177(98)00027-4},
   Key = {fds276987}
}

@article{fds277140,
   Author = {McKinney, CJ and MacCormac, ER and Welsh-Bohmer,
             KA},
   Title = {Hardware and software for tachistoscopy: how to make
             accurate measurements on any PC utilizing the Microsoft
             Windows operating system.},
   Journal = {Behav Res Methods Instrum Comput},
   Volume = {31},
   Number = {1},
   Pages = {129-136},
   Year = {1999},
   Month = {February},
   ISSN = {0743-3808},
   url = {http://www.ncbi.nlm.nih.gov/pubmed/10495844},
   Abstract = {Methods for automated stimulus display and accurate response
             time measurement with IBM-compatible PCs are of great
             importance in cognitive research designs. Accurate
             measurements of reaction times are required to interpolate
             other measures, such as speed of mental processing. We
             present a description of hardware and software that displays
             stimulus images and performs reaction timing that is not
             dependent on PC performance characteristics. This is
             accomplished by electronically bypassing timing errors
             normally inherent to the operation of the computer. A
             high-precision external timer measures the time between
             stimulus onset and a subject's push-button response while a
             video blanking circuit controls the video presented to the
             monitor screen. Two options for accurately detecting
             stimulus onset are presented: (1) A photodetector can be
             used to sense the actual onset of the stimulus on a
             secondary video monitor screen; (2) the video blanking
             circuit can provide a signal coincident with the initiation
             of video to the monitor. Both methods result in a system
             timing accuracy of 100 microseconds.},
   Doi = {10.3758/bf03207703},
   Key = {fds277140}
}

@article{fds277014,
   Author = {Madden, DJ and Welsh-Bohmer, KA and Tupler, LA},
   Title = {Task complexity and signal detection analyses of lexical
             decision performance in Alzheimer's disease},
   Journal = {Developmental Neuropsychology},
   Volume = {16},
   Number = {1},
   Pages = {1-18},
   Publisher = {Informa UK Limited},
   Year = {1999},
   Month = {January},
   url = {http://dx.doi.org/10.1207/S15326942DN160101},
   Abstract = {This experiment addressed the issue of whether the changes
             in semantic memory performance associated with Alzheimer's
             disease (AD) could be distinguished from a generalized
             cognitive slowing. Young adults, healthy older adults, and
             AD patients performed 3 different reaction time (RT) tasks
             involving yes-no responses to visually presented letter
             strings. Task complexity analyses indicated that performance
             in the semantic task (lexical decision) was consistent with
             a generalized slowing of cognitive function that was greater
             in magnitude for AD than for normal aging. Signal detection
             analyses of the lexical decision data demonstrated
             AD-related changes in word-nonword discrimination, response
             bias, and the relation between discrimination and RT. The
             general cognitive slowing associated with AD was accompanied
             by additional changes specific to the performance of this
             semantic memory task.},
   Doi = {10.1207/S15326942DN160101},
   Key = {fds277014}
}

@article{fds371216,
   Author = {Koltai, DC and Welsh-Bohmer, KA and Schmechel,
             DE},
   Title = {Subjective appraisals of memory functioning in dementia:
             Validity and cautions},
   Journal = {Archives of Clinical Neuropsychology},
   Volume = {14},
   Number = {1},
   Pages = {82-82},
   Publisher = {Oxford University Press (OUP)},
   Year = {1999},
   Month = {January},
   url = {http://dx.doi.org/10.1093/arclin/14.1.82},
   Doi = {10.1093/arclin/14.1.82},
   Key = {fds371216}
}

@article{fds277019,
   Author = {Hulette, CM and Welsh-Bohmer, KA and Murray, MG and Saunders, AM and Mash, DC and McIntyre, LM},
   Title = {Neuropathological and neuropsychological changes in "normal"
             aging: evidence for preclinical Alzheimer disease in
             cognitively normal individuals.},
   Journal = {J Neuropathol Exp Neurol},
   Volume = {57},
   Number = {12},
   Pages = {1168-1174},
   Year = {1998},
   Month = {December},
   ISSN = {0022-3069},
   url = {http://www.ncbi.nlm.nih.gov/pubmed/9862640},
   Abstract = {The presence of diffuse or primitive senile plaques in the
             neocortex of cognitively normal elderly at autopsy has been
             presumed to represent normal aging. Alternatively, these
             patients may have developed dementia and clinical Alzheimer
             disease (AD) if they had survived. In this setting, these
             patients could be subjects for cognitive or pharmacologic
             intervention to delay disease onset. We have thus followed a
             cohort of cognitively normal elderly subjects with a
             Clinical Dementia Rating (CDR) of 0 at autopsy. Thirty-one
             brains were examined at postmortem according to Consortium
             to Establish a Registry for Alzheimer Disease (CERAD)
             criteria and staged according to Braak. Ten patients were
             pathologically normal according to CERAD criteria (1a). Two
             of these patients were Braak Stage II. Seven very elderly
             subjects exhibited a few primitive neuritic plaques in the
             cortex and thus represented CERAD 1b. These individuals
             ranged in age from 85 to 105 years and were thus older than
             the CERAD la group that ranged in age from 72 to 93.
             Fourteen patients displayed Possible AD according to CERAD
             with ages ranging from 66 to 95. Three of these were Braak
             Stage I, 4 were Braak Stage II, and 7 were Braak Stage III.
             The Apolipoprotein E4 allele was over-represented in this
             possible AD group. Neuropsychological data were available on
             12 individuals. In these 12 individuals, Possible AD at
             autopsy could be predicted by cognitive deficits in 1 or
             more areas including savings scores on memory testing and
             overall performance on some measures of frontal executive
             function.},
   Doi = {10.1097/00005072-199812000-00009},
   Key = {fds277019}
}

@article{fds276984,
   Author = {Clark, LM and McDonald, WM and Welsh-Bohmer, KA and Siegler, IC and Dawson, DV and Tupler, LA and Krishnan, KR},
   Title = {Magnetic resonance imaging correlates of depression in
             early- and late-onset Alzheimer's disease.},
   Journal = {Biol Psychiatry},
   Volume = {44},
   Number = {7},
   Pages = {592-599},
   Year = {1998},
   Month = {October},
   ISSN = {0006-3223},
   url = {http://www.ncbi.nlm.nih.gov/pubmed/9787883},
   Abstract = {BACKGROUND: Depressive symptoms are frequent complications
             of Alzheimer's disease (AD). We hypothesized that AD
             patients with depression would be more likely than
             nondepressed AD patients to show deep white-matter,
             subcortical gray-matter, and periventricular
             hyperintensities on magnetic resonance imaging (MRI).
             METHODS: In a retrospective study of 31 AD patients,
             depression was characterized by clinical diagnosis
             (DSM-III-R major depression, depressive symptoms, or no
             depression), a clinician-rated depression scale, and
             informant ratings of premorbid (before memory disorder) as
             well as current depression using the NEO Personality
             Inventory (NEO-PI), and related to qualitative and
             quantitative ratings of MRI hyperintensities. RESULTS: In
             contrast to reports in nondemented elderly patients, there
             was no relationship between clinical diagnosis of major
             depressive episode and hyperintensities; however,
             clinician-rated depressive symptoms were higher in subjects
             with large anterior hyperintensities. In the early-onset AD
             group only, MRI abnormalities were related to greater
             premorbid depression, and less increase in depression after
             the onset of dementia, as rated by informants on the NEO-PI.
             CONCLUSIONS: Results highlight the need to consider early-
             and late-onset AD separately when assessing relationships
             between personality and MRI abnormalities, and to consider
             premorbid personality style when drawing conclusions about
             the etiology of depressive features seen in
             AD.},
   Doi = {10.1016/s0006-3223(98)00106-1},
   Key = {fds276984}
}

@article{fds277147,
   Author = {Meyer, MR and Tschanz, JT and Norton, MC and Welsh-Bohmer, KA and Steffens, DC and Wyse, BW and Breitner, JC},
   Title = {APOE genotype predicts when--not whether--one is predisposed
             to develop Alzheimer disease.},
   Journal = {Nat Genet},
   Volume = {19},
   Number = {4},
   Pages = {321-322},
   Year = {1998},
   Month = {August},
   ISSN = {1061-4036},
   url = {http://www.ncbi.nlm.nih.gov/pubmed/9697689},
   Doi = {10.1038/1206},
   Key = {fds277147}
}

@article{fds277139,
   Author = {Bass, MP and Yamaoka, LH and Scott, WK and Gaskell, PC and Welsh-Bohmer,
             KA and Roses, AD and Saunders, AM and Haines, JL and Pericak-Vance,
             MA},
   Title = {No association of alpha1-antichymotrypsin flanking region
             polymorphism and Alzheimer disease risk in early- and
             late-onset Alzheimer disease patients.},
   Journal = {Neurosci Lett},
   Volume = {250},
   Number = {2},
   Pages = {79-82},
   Year = {1998},
   Month = {July},
   ISSN = {0304-3940},
   url = {http://www.ncbi.nlm.nih.gov/pubmed/9697923},
   Abstract = {The alpha1-antichymotrypsin (AACT)-155 allele was found
             elsewhere to have a significant effect on Alzheimer disease
             (AD) risk in individuals with at least one APOE-4 allele. We
             compared AACT genotypes of 284 cases of sporadic AD and 172
             controls. The frequency of the AACT-155 allele did not
             differ significantly between cases and controls, either
             overall or when restricted to subjects with at least one
             APOE-4 allele. Logistic regression controlling for age and
             sex failed to show an effect due to AACT either alone or
             acting with APOE. There was no evidence of an interaction
             between APOE-4 and the AACT-155 allele to reduce age at
             onset. Thus, our data do not support an association of
             AACT-155 with risk or age at onset in AD.},
   Doi = {10.1016/s0304-3940(98)00398-x},
   Key = {fds277139}
}

@article{fds277059,
   Author = {Fillenbaum, GG and Peterson, B and Welsh-Bohmer, KA and Kukull, WA and Heyman, A},
   Title = {Progression of Alzheimer's disease in black and white
             patients: the CERAD experience, part XVI. Consortium to
             Establish a Registry for Alzheimer's Disease.},
   Journal = {Neurology},
   Volume = {51},
   Number = {1},
   Pages = {154-158},
   Publisher = {Ovid Technologies (Wolters Kluwer Health)},
   Year = {1998},
   Month = {July},
   ISSN = {0028-3878},
   url = {http://dx.doi.org/10.1212/wnl.51.1.154},
   Abstract = {We compared the progression of Alzheimer's disease (AD) in
             CERAD-enrolled black and white patients, as indicated by
             changes in selected clinical and neuropsychology measures,
             over a 1-year time interval. Of 225 black and 935 white AD
             patients who were enrolled, 148 (66%) black and 770 (82%)
             white patients remained in the study. Of these, 82 black and
             532 white patients provided complete in-person information
             on first annual re-evaluation. Overall, with age, education,
             initial level of performance on each measure, and stage of
             disease at entry controlled, race had a very mild effect on
             change in disease (8 df multivariate analysis of variance
             [MANOVA], p < 0.047). Black patients showed less decline
             than white patients, most notably for the CERAD Boston
             Naming test (p < 0.02) and the third and final trial of the
             10-item Word List Learning task (p < 0.003). Although
             unadjusted data indicate that black and white patients
             appear to differ notably at entry, our findings indicated
             that differences in progression of the dementing process are
             minor, suggesting that course of AD is comparable in these
             racial groups. Examination over a longer period is difficult
             because of the high attrition rate of black
             patients.},
   Doi = {10.1212/wnl.51.1.154},
   Key = {fds277059}
}

@article{fds277060,
   Author = {Heyman, A and Fillenbaum, GG and Welsh-Bohmer, KA and Gearing, M and Mirra, SS and Mohs, RC and Peterson, BL and Pieper,
             CF},
   Title = {Cerebral infarcts in patients with autopsy-proven
             Alzheimer's disease: CERAD, part XVIII. Consortium to
             Establish a Registry for Alzheimer's Disease.},
   Journal = {Neurology},
   Volume = {51},
   Number = {1},
   Pages = {159-162},
   Publisher = {Ovid Technologies (Wolters Kluwer Health)},
   Year = {1998},
   Month = {July},
   ISSN = {0028-3878},
   url = {http://dx.doi.org/10.1212/wnl.51.1.159},
   Abstract = {OBJECTIVE: To study the relation between cerebral infarction
             and clinical and neuropsychologic manifestations in patients
             with autopsy-proven Alzheimer's disease (AD) enrolled in the
             Consortium to Establish a Registry for Alzheimer's Disease
             (CERAD). BACKGROUND: Prior studies report that subjects with
             neuropathologic evidence of AD and concomitant brain
             infarcts had poorer cognitive function and higher frequency
             of dementia than those with AD alone. METHODS: Clinical and
             neuropsychologic manifestations of dementia were studied in
             74 subjects with neuropathologic findings of AD alone and 32
             with AD and concomitant cerebral infarcts or lacunar
             lesions. RESULTS: The 32 patients with both AD and vascular
             lesions were significantly older at time of death (median
             age, 81 years) than the 74 patients with AD alone (76 years;
             p = 0.02). At the final follow-up visit, the severity of the
             dementia was greater in AD patients with vascular lesions
             (median Clinical Dementia Rating [CDR] = 3) than in those
             with AD alone (CDR = 2; p = 0.03). Patients with AD and
             vascular lesions performed significantly worse on verbal
             fluency, Boston Naming, and Mini-Mental State Examination
             (MMSE) tests. No differences between the groups were
             observed, however, in the semiquantitative measures of
             frequency of neuritic plaques or neurofibrillary tangles.
             CONCLUSIONS: The clinical-neuropathologic correlations in
             CERAD patients generally confirm those in prior studies,
             indicating that the presence of cerebral infarction in
             patients with AD is associated with greater overall severity
             of clinical dementia and poorer performance on specific
             tests of language and cognitive function.},
   Doi = {10.1212/wnl.51.1.159},
   Key = {fds277060}
}

@article{fds371217,
   Author = {Welsh-Bohmer, KA and Plassman, BL and Tschanz, JT and Norton, MC and Helms, MJ and Wyse, BW and Breitner, JCS},
   Title = {Cognitive decline in aging: What is the role of
             education?},
   Journal = {CLINICAL NEUROPSYCHOLOGIST},
   Volume = {12},
   Number = {2},
   Pages = {265-265},
   Publisher = {SWETS ZEITLINGER PUBLISHERS},
   Year = {1998},
   Month = {May},
   Key = {fds371217}
}

@article{fds371218,
   Author = {Lowry, CM and Bigler, ED and Plassman, BL and Tshcantz, J and Welsh-Bohmer, KA and Saunders, AM and Breitner,
             JCS},
   Title = {A new method for quantifying cerebral atrophy in Alzheimer's
             disease},
   Journal = {CLINICAL NEUROPSYCHOLOGIST},
   Volume = {12},
   Number = {2},
   Pages = {258-258},
   Publisher = {SWETS ZEITLINGER PUBLISHERS},
   Year = {1998},
   Month = {May},
   Key = {fds371218}
}

@article{fds277138,
   Author = {Ikeuchi, T and Sanpei, K and Takano, H and Sasaki, H and Tashiro, K and Cancel, G and Brice, A and Bird, TD and Schellenberg, GD and Pericak-Vance, MA and Welsh-Bohmer, KA and Clark, LN and Wilhelmsen,
             K and Tsuji, S},
   Title = {A novel long and unstable CAG/CTG trinucleotide repeat on
             chromosome 17q.},
   Journal = {Genomics},
   Volume = {49},
   Number = {2},
   Pages = {321-326},
   Year = {1998},
   Month = {April},
   ISSN = {0888-7543},
   url = {http://www.ncbi.nlm.nih.gov/pubmed/9598323},
   Abstract = {Using the direct identification of repeat expansion and
             cloning technique, we cloned a novel long CAG/CTG
             trinucleotide repeat on chromosome 17. Using radiation
             hybrid panels, the CAG/CTG repeat was mapped to chromosome
             17q. The CAG/CTG repeat is highly polymorphic, with a
             heterozygosity of 85%, and exhibits a bimodal distribution
             (allele S, 10-26 repeat units, and allele L, 50-92 repeat
             units). The CAG/CTG repeat of allele L exhibited
             intergenerational instabilities, which are more prominent in
             maternal transmission than in paternal transmission.
             Analyses of Northern blot and RT-PCR indicate that the
             repeat is transcribed. Although the size of the CAG/CTG
             repeat of allele L is within the range of the expanded CAG
             repeat of disease-causing genes, we did not detect any
             association of allele L with various neurodegenerative
             diseases, including frontotemporal dementia and
             parkinsonism, mapped to 17q21-q23.},
   Doi = {10.1006/geno.1998.5266},
   Key = {fds277138}
}

@article{fds371219,
   Author = {Pericak-Vance, MA and Rimmler, JB and Gaskell, PC and Abou-Donia, S and Welsh-Bohmer, KA and Jackson, CE and Schamel, MA and Yamaoka, LH and Haines, JL},
   Title = {A genomic screen in extended Amish families with Alzheimer's
             disease supports a locus on chromosome 12},
   Journal = {NEUROLOGY},
   Volume = {50},
   Number = {4},
   Pages = {A356-A356},
   Publisher = {LIPPINCOTT WILLIAMS & WILKINS},
   Year = {1998},
   Month = {April},
   Key = {fds371219}
}

@article{fds371220,
   Author = {Meyer, MR and Breitner, JCS and Steffens, DC and Anthony, JC and Norton,
             MA and Tschanz, JT and Wyse, BW and Welsh-Bohmer, KA and Plassman,
             BL},
   Title = {196. Onset of alzheimer's disease in susceptible subjects:
             Influence of APOE},
   Journal = {Biological Psychiatry},
   Volume = {43},
   Number = {8},
   Pages = {S58-S58},
   Publisher = {Elsevier BV},
   Year = {1998},
   Month = {April},
   url = {http://dx.doi.org/10.1016/s0006-3223(98)90644-8},
   Doi = {10.1016/s0006-3223(98)90644-8},
   Key = {fds371220}
}

@article{fds277001,
   Author = {Breitner, JC and Jarvik, GP and Plassman, BL and Saunders, AM and Welsh,
             KA},
   Title = {Risk of Alzheimer disease with the epsilon4 allele for
             apolipoprotein E in a population-based study of men aged
             62-73 years.},
   Journal = {Alzheimer Dis Assoc Disord},
   Volume = {12},
   Number = {1},
   Pages = {40-44},
   Year = {1998},
   Month = {March},
   ISSN = {0893-0341},
   url = {http://www.ncbi.nlm.nih.gov/pubmed/9539409},
   Abstract = {The epsilon4 allele at APOE, the polymorphic locus for
             apolipoprotein E, increases the risk of Alzheimer disease
             (AD), especially among those with the homozygous
             epsilon4/epsilon4 genotype. In family studies, epsilon4
             homozygotes typically develop AD at 55-75 years, an age
             range when AD is otherwise relatively infrequent.
             Population-based studies of the AD risk associated with
             allele epsilon4 (and especially with genotype
             epsilon4/epsilon4) are limited in number, and most such
             studies have included few AD cases between the ages of 55
             and 75 years. In a large population-based twin registry, the
             screening of 12,709 men who were 62-73 years old yielded 38
             prevalent cases of AD whose onset age ranged from 54 to 73.
             Genotype at APOE was determined for 37 of these cases and
             independently, for a similarly aged probability sample of
             344 men from the same registry. The epsilon4 allele
             frequencies among the AD cases and the population samples
             were 0.39 and 0.15, respectively. The odds ratios (ORs) for
             AD were 17.7 for genotype epsilon4/epsilon4 versus
             epsilon3/epsilon3 and 13.8 for epsilon4/epsilon4 versus all
             remaining genotypes. By contrast, the ORs with heterozygous
             epsilon4/epsilon3 were only 2.76 versus epsilon3/epsilon3
             and 2.01 versus all genotypes other than epsilon4/epsilon3
             (p for homozygote vs. heterozygote ORs=0.002). The estimated
             etiologic fraction for AD with homozygous epsilon4 among men
             in their mid-50s to mid 70s is therefore 0.20; for the much
             more common heterozygous genotype epsilon4/epsilon3, the
             fraction is 0.18. In combination with other studies that
             have adjusted statistically for age, these results suggest
             that the effect of the epsilon4 allele dose is neither
             linear nor homogeneous for age. Homozygous epsilon4/epsilon4
             appears to confer an extreme risk of AD at the age when
             onset with this genotype is most likely. These results are
             consistent with the view that individual genotypes modify
             risk by predisposing to substantially different
             distributions of AD onsets.},
   Doi = {10.1097/00002093-199803000-00006},
   Key = {fds277001}
}

@article{fds371221,
   Author = {Heyman, A and Fillenbaum, G and Welsh-Bohmer, K and Gearing, M and Mirra, SS and Mohs, R and Peterson, B and Pieper,
             C},
   Title = {Clinical and neuropsychological findings in patients with
             autopsy evidence of Alzheimer's disease with and without
             cerebral infarction: The CERAD experience},
   Journal = {NEUROLOGY},
   Volume = {50},
   Number = {4},
   Pages = {A408-A408},
   Year = {1998},
   Key = {fds371221}
}

@article{fds277136,
   Author = {Welsh-Bohmer, KA and Gearing, M and Saunders, AM and Roses, AD and Mirra, S},
   Title = {Apolipoprotein E genotypes in a neuropathological series
             from the Consortium to Establish a Registry for Alzheimer's
             Disease.},
   Journal = {Ann Neurol},
   Volume = {42},
   Number = {3},
   Pages = {319-325},
   Year = {1997},
   Month = {September},
   ISSN = {0364-5134},
   url = {http://www.ncbi.nlm.nih.gov/pubmed/9307253},
   Abstract = {We evaluated the sensitivity, specificity, and predictive
             value of the apolipoprotein E (ApoE) epsilon4 allele for the
             neuropathological diagnosis of Alzheimer disease (AD) in a
             clinical series of well-characterized AD patients and
             controls followed longitudinally in the multicenter study of
             the Consortium to Establish a Registry for Alzheimer's
             Disease (CERAD). In the 162 patients for whom autopsy and
             ApoE genotype data were available, the clinical diagnosis of
             AD was verified as the primary cause of dementia in 139
             cases (139 of 162, or 86%). The sensitivity and specificity
             of the epsilon4 allele for AD were both 83%. The positive
             predictive value of the epsilon4 allele was very high at 97%
             (115 of 119); whereas, the negative predictive value was 44%
             (19 of 43), providing no useful information for diagnosis of
             AD when the epsilon4 allele is not present. Of the cases
             where AD was not considered the primary or sole cause of
             dementia (n = 23), 6 cases exhibited concomitant
             neuropathological findings of definite or probable AD and 2
             of these 6 cases had at least one epsilon4 allele. These
             multicenter data extend previous observations reported from
             smaller case series of single laboratories and demonstrate
             that once the clinical diagnosis of AD is established, the
             presence of an epsilon4 allele reliably predicts the
             ultimate CERAD neuropathological diagnosis of AD. The
             findings also suggest that ApoE genotype information is
             useful clinically in bolstering diagnostic confidence when
             an epsilon4 allele is present and in identifying a group of
             patients for whom additional diagnostic evaluations may be
             warranted when the epsilon4 allele is absent.},
   Doi = {10.1002/ana.410420308},
   Key = {fds277136}
}

@article{fds277137,
   Author = {Welsh-Bohmer, KA and Mohs, RC},
   Title = {Neuropsychological assessment of Alzheimer's
             disease.},
   Journal = {Neurology},
   Volume = {49},
   Number = {3 Suppl 3},
   Pages = {S11-S13},
   Year = {1997},
   Month = {September},
   ISSN = {0028-3878},
   url = {http://dx.doi.org/10.1212/wnl.49.3_suppl_3.s11},
   Abstract = {The CERAD neuropsychological battery has been found to be a
             valuable tool for clinical as well as research
             investigations and has contributed considerably to the
             identification and clinical understanding of AD. The CERAD
             data addressing effects of age, education, sex, race, and
             culture on neuropsychological performance are important in
             the clinical interpretation of cognitive performance in the
             elderly. Information on the performance of AD patients at
             different levels of severity of dementia has enhanced the
             reliability of clinical detection and staging of AD dementia
             over time. Studies addressing longitudinal changes on the
             CERAD neuropsychological measures have also helped to define
             the natural history of AD and provide insights into the
             limitations of various neuropsychological measures at
             different stages of the illness. The neuropsychological
             battery, although designed to assess AD, may be valuable in
             differential diagnosis of some dementias.},
   Doi = {10.1212/wnl.49.3_suppl_3.s11},
   Key = {fds277137}
}

@article{fds277018,
   Author = {Hulette, CM and Welsh-Bohmer, KA and Crain, B and Szymanski, MH and Sinclaire, NO and Roses, AD},
   Title = {Rapid brain autopsy. The Joseph and Kathleen Bryan
             Alzheimer's Disease Research Center experience.},
   Journal = {Arch Pathol Lab Med},
   Volume = {121},
   Number = {6},
   Pages = {615-618},
   Year = {1997},
   Month = {June},
   ISSN = {0003-9985},
   url = {http://www.ncbi.nlm.nih.gov/pubmed/9199629},
   Abstract = {OBJECTIVE: To develop a system for retrieving brain tissue
             within 1 hour after death in an effective and useful manner.
             DESIGN: Nurse clinicians were employed as study
             co-ordinators and were available to families 24 hours each
             day. SETTING: Autopsies were performed at Duke University
             Medical Center, Durham, NC, from 1985 through 1995.
             PARTICIPANTS: Neuropathology faculty, fellows, and
             residents, autopsy technicians; and brain bank staff.
             RESULTS: Fifty-one rapid autopsies with a postmortem
             interval of less than 1 hour have been performed. Four of
             these were normal controls, three were disease controls, and
             44 represented Alzheimer's disease patients. Tissue
             retrieved at rapid autopsy has been distributed to 93
             research teams, 30 of these located at Duke University
             Medical Center. Many researchers have received multiple
             shipments of tissue. CONCLUSIONS: The Bryan Alzheimer's
             Disease Research Center Rapid Autopsy Program at Duke
             University Medical Center has been successful in retrieving
             tissue from individuals with dementia and also from controls
             within 1 hour of death. The critical features of the success
             of this program have been the use of nurse clinicians who
             work closely with patients and their families to ensure a
             successful autopsy at the time of death and the maintenance
             of a 24-hour call schedule for nurses and neuropathology
             staff. Similar programs can be implemented for experimental
             work into the pathogenesis of a wide variety of human
             diseases in which the examination of human tissue is
             required.},
   Key = {fds277018}
}

@article{fds276999,
   Author = {Steffens, DC and Plassman, BL and Helms, MJ and Welsh-Bohmer, KA and Saunders, AM and Breitner, JC},
   Title = {A twin study of late-onset depression and apolipoprotein E
             epsilon 4 as risk factors for Alzheimer's
             disease.},
   Journal = {Biol Psychiatry},
   Volume = {41},
   Number = {8},
   Pages = {851-856},
   Year = {1997},
   Month = {April},
   ISSN = {0006-3223},
   url = {http://www.ncbi.nlm.nih.gov/pubmed/9099411},
   Abstract = {A prior history of depression and the epsilon 4 allele of
             apolipoprotein E (APOE) have each been associated with
             development of Alzheimer's disease (AD). In a sample of 142
             elderly twins from a large study of dementia, we examined
             the relation of major depression, APOE genotype and AD using
             time-dependent proportional hazards models. Compared against
             the risk for AD with no history of depression and no epsilon
             4 allele, the risk ratio for AD with two epsilon 4 alleles
             was 2.87 (C.I. = 1.56-5.28), with one epsilon 4 allele, 1.82
             (C.I. = 1.09-3.04) and with late-onset depression and no
             epsilon 4 allele, 2.95 (C.I. = 1.55-5.62). There was no
             suggestion of an interaction between prior depression and
             APOE genotype in their effects on AD risk. Results were
             similar when the sample was stratified by twin pair, so that
             a single genetic marker is unlikely to explain the relation
             among depression, APOE, and dementia. Risk ratios declined
             substantially with increasing intervals between the onset of
             depression and AD. Thus, for many individuals, the
             association of depression and AD may reflect the occurrence
             of prodromal depressive symptoms rather than a true risk
             relationship.},
   Doi = {10.1016/S0006-3223(96)00247-8},
   Key = {fds276999}
}

@article{fds276998,
   Author = {Plassman, BL and Welsh-Bohmer, KA and Bigler, ED and Johnson, SC and Anderson, CV and Helms, MJ and Saunders, AM and Breitner,
             JC},
   Title = {Apolipoprotein E epsilon 4 allele and hippocampal volume in
             twins with normal cognition.},
   Journal = {Neurology},
   Volume = {48},
   Number = {4},
   Pages = {985-989},
   Year = {1997},
   Month = {April},
   ISSN = {0028-3878},
   url = {http://www.ncbi.nlm.nih.gov/pubmed/9109888},
   Abstract = {We examined the relation of APOE-epsilon 4, hippocampal
             volume, and cognitive performance in ten pairs of
             cognitively normal twins who had a mean age of 62.5 years
             (SD = 7.8). There were no significant differences in
             neuropsychological measures of the groups categorized by the
             presence of an epsilon 4 allele. However, the mean
             normalized right and left hippocampal volumes were smaller
             in the epsilon 4 groups compared to the group without
             epsilon 4. Combined with prior reports, these findings
             suggest that epsilon 4 is associated with differences in
             brain morphology that may be evident when no symptoms of
             dementia are present.},
   Doi = {10.1212/wnl.48.4.985},
   Key = {fds276998}
}

@article{fds371222,
   Author = {Rimmler, JB and Gaskell, PC and Aboudonia, S and Welsh-Bohmer, K and Jackson, CE and Schamel, M and Yamaoka, L and Haines, JL and Pericak-Vance, MA},
   Title = {A genomic screen in extended Amish families supports a locus
             on chromosome 12 for Alzheimer disease (AD).},
   Journal = {AMERICAN JOURNAL OF HUMAN GENETICS},
   Volume = {61},
   Number = {4},
   Pages = {A292-A292},
   Year = {1997},
   Key = {fds371222}
}

@article{fds277149,
   Author = {Yamaoka, LH and Welsh-Bohmer, KA and Hulette, CM and Gaskell, PC and Murray, M and Rimmler, JL and Helms, BR and Guerra, M and Roses, AD and Schmechel, DE and Pericak-Vance, MA},
   Title = {Linkage of frontotemporal dementia to chromosome 17:
             clinical and neuropathological characterization of
             phenotype.},
   Journal = {Am J Hum Genet},
   Volume = {59},
   Number = {6},
   Pages = {1306-1312},
   Year = {1996},
   Month = {December},
   ISSN = {0002-9297},
   url = {http://www.ncbi.nlm.nih.gov/pubmed/8940276},
   Abstract = {Frontotemporal dementia is a behavioral disorder of
             insidious onset and variable progression. Clinically, its
             early features reflect frontal lobe dysfunction
             characterized by personality change, deterioration in memory
             and executive functions, and stereotypical and perseverative
             behaviors. Pathologically, there is degeneration of the
             neocortex and subcortical nuclei, without distinctive
             features such as plaques, neurofibrillary tangles, or Pick
             or Lewy bodies. Within-family variation in neuropathology
             and clinical phenotype is observed. In cases where family
             aggregation is observed, it is inherited as an autosomal
             dominant, age-dependent disorder. Family studies recently
             have identified two dementia loci: chromosome 17 for
             disinhibition-dementia-parkinsonism-amyotrophic complex and
             pallido-ponto-nigral degeneration and chromosome 3 for
             familial nonspecific dementia. We describe a family
             (DUK1684) with clinically and neuropathologically confirmed,
             autosomal dominant, non-Alzheimer disease dementia. Linkage
             analysis of this family showed evidence for linkage to
             chromosome 17q21, with a multipoint location score (log10)
             of 5.52. A comparison of the clinical and pathological
             features in DUK1684 with those of the other chromosome
             17-linked families, together with the linkage data, suggests
             that these families are allelic. These studies emphasize
             that genetic linkage analysis remains a useful tool for
             differentiating disease loci in clinically complex
             traits.},
   Key = {fds277149}
}

@article{fds277135,
   Author = {Davidson, M and Harvey, P and Welsh, KA and Powchik, P and Putnam, KM and Mohs, RC},
   Title = {Cognitive functioning in late-life schizophrenia: a
             comparison of elderly schizophrenic patients and patients
             with Alzheimer's disease.},
   Journal = {Am J Psychiatry},
   Volume = {153},
   Number = {10},
   Pages = {1274-1279},
   Year = {1996},
   Month = {October},
   ISSN = {0002-953X},
   url = {http://dx.doi.org/10.1176/ajp.153.10.1274},
   Abstract = {OBJECTIVE: Previous studies have suggested that geriatric
             inpatients with chronic schizophrenia manifest profound
             cognitive impairments. This study investigated how these
             cognitive impairments resemble those seen in degenerative
             dementing conditions. METHOD: The neuropsychological battery
             of the Consortium to Establish a Registry for Alzheimer's
             Disease (CERAD), widely used to characterize the cognitive
             deficits of patients with Alzheimer's disease, was used to
             compare patterns of cognitive impairment in 66 triads of
             subjects consisting of one elderly patient with Alzheimer's
             disease, one elderly, institutionalized patient with chronic
             schizophrenia, and one elderly, cognitively normal
             comparison subject who were matched on age, gender, and
             education. For some analyses, the two groups of patients
             were divided into subgroups according to the degree of their
             cognitive impairment (mild, moderate, or severe) as
             determined by their scores on the Mini-Mental State
             examination. RESULTS: Relative to the comparison subjects,
             both groups of patients showed cognitive deficits on each of
             the neuropsychological measures. The schizophrenic patients
             performed worse than the patients with Alzheimer's disease
             on tests of naming and constructional praxis but were less
             impaired on the test of delayed word recall. These
             differences were consistent across all levels of severity of
             globally measured cognitive impairment. CONCLUSIONS:
             Consistent with earlier findings from postmortem studies,
             these findings suggest that major differences exist in the
             neurobiologic mechanisms responsible for cognitive
             impairment in schizophrenia and Alzheimer's disease. Effects
             directly attributable to social and environmental
             differences between these two groups of patients may also
             play a role.},
   Doi = {10.1176/ajp.153.10.1274},
   Key = {fds277135}
}

@article{fds277075,
   Author = {Saunders, AM and Hulette, O and Welsh-Bohmer, KA and Schmechel, DE and Crain, B and Burke, JR and Alberts, MJ and Strittmatter, WJ and Breitner, JC and Rosenberg, C},
   Title = {Specificity, sensitivity, and predictive value of
             apolipoprotein-E genotyping for sporadic Alzheimer's
             disease.},
   Journal = {Lancet},
   Volume = {348},
   Number = {9020},
   Pages = {90-93},
   Year = {1996},
   Month = {July},
   ISSN = {0140-6736},
   url = {http://www.ncbi.nlm.nih.gov/pubmed/8676723},
   Abstract = {BACKGROUND: We aimed to determine the specificity,
             sensitivity, and predictive value of apolipoprotein E (APOE)
             genotyping in 67 consecutive patients with clinical
             diagnoses of sporadic Alzheimer's disease (AD) who underwent
             necropsy. METHODS: We studied patients who attended the Duke
             Memory Disorders Clinic and were diagnosed as having
             probable AD. These patients were followed up until they
             died. APOE genotyping was done during life in most cases,
             but in some brain tissue obtained at necropsy was used.
             Members of known AD families were excluded. FINDINGS: After
             neuropathological examination 57 (85%) of 67 of our patients
             were confirmed as having AD including all 43 who had at
             least one APOE-epsilon 4 allele. None of the patients found
             not to have AD carried an epsilon 4 allele. In this series,
             the specificity of the epsilon 4 allele was 100%, the
             sensitivity 75%, the positive predictive value 100%, and the
             negative predictive value 42%. In this necropsy-confirmed
             series, the epsilon 4/epsilon 4 genotype predicted AD with
             100% accuracy. The epsilon 3/epsilon 4 and epsilon 2/epsilon
             4 genotypes were also unexpectedly highly specific for AD.
             INTERPRETATION: Data from hundreds of necropsy-confirmed
             non-AD patients in other longitudinal necropsy series will
             allow the predictive value of APOE genotypes to be assessed
             with useful confidence limits.},
   Doi = {10.1016/s0140-6736(96)01251-2},
   Key = {fds277075}
}

@article{fds277054,
   Author = {Hoffman, JM and Hanson, MW and Welsh, KA and Earl, N and Paine, S and Delong, D and Coleman, RE},
   Title = {Interpretation variability of 18FDG-positron emission
             tomography studies in dementia.},
   Journal = {Invest Radiol},
   Volume = {31},
   Number = {6},
   Pages = {316-322},
   Year = {1996},
   Month = {June},
   ISSN = {0020-9996},
   url = {http://www.ncbi.nlm.nih.gov/pubmed/8761863},
   Abstract = {RATIONALE AND OBJECTIVES: Functional imaging studies such as
             18F-fluoro-18-labeled-deoxyglucose-positron emission
             tomography (18FDG-PET) are being used increasingly in the
             evaluation of patients with dementia. The authors evaluate
             inter- and intraobserver interpretation agreement in a
             diverse group of patients with clinically diagnosed dementia
             and subjective memory complaints, as well as two healthy
             control subjects. METHODS: Ninety-six patients with clinical
             diagnoses of probable Alzheimer's disease (n = 18), possible
             Alzheimer's disease (n = 33), dementia (n = 26), and mild
             memory impairment (n = 17), as well as two healthy control
             subjects were studied using 18FDG-PET. Three observers
             graded all studies for regional 18FDG uptake in the
             temporal, parietal, and frontal regions bilaterally. The
             studies also were interpreted for the presence of bilateral
             temporoparietal hypometabolism, which typically is present
             in Alzheimer's disease. The kappa statistic was used to
             determine intra- and interobserver agreement for regional
             18FDG uptake and bilateral temporoparietal hypometabolism.
             RESULTS: There was excellent intraobserver (kappa = .56, P <
             0.0005) and interobserver (kappa = .51, P < 0.0005)
             interpretation agreement for bilateral temporoparietal
             hypometabolism. There also was excellent intraobserver
             (kappa = .61, P < 0.000) and interobserver (kappa = .55, P <
             0.000) interpretation agreement of regional 18FDG uptake.
             Interobserver agreement was extremely high in those patients
             who were considered clinically to have possible (kappa =
             .42, P < 0.001) or probable (kappa = .42, P < 0.01)
             Alzheimer's disease. CONCLUSIONS: Results confirm that
             bilateral temporoparietal hypometabolism is the metabolic
             abnormality associated with the diagnosis of probable
             Alzheimer's disease. Furthermore, intra- and interobserver
             agreement of visual interpretation of 18FDG-PET images
             indicates that 18FDG-PET is acceptable as an imaging
             technique in the clinical evaluation of the dementia
             patient.},
   Doi = {10.1097/00004424-199606000-00002},
   Key = {fds277054}
}

@article{fds277074,
   Author = {Steffens, DC and Welsh, KA and Burke, JR and Helms, MJ and Folstein, MF and Brandt, J and McDonald, WM and Breitner, JCS},
   Title = {Diagnosis of Alzheimer's disease in epidemiologic studies by
             staged review of clinical data},
   Journal = {Neuropsychiatry, Neuropsychology and Behavioral
             Neurology},
   Volume = {9},
   Number = {2},
   Pages = {107-113},
   Year = {1996},
   Month = {April},
   Abstract = {We explored the inter-rater agreement and validity of
             diagnoses of Alzheimer's disease (AD) and other dementias
             made in an epidemiological study. A previously described
             protocol for cognitive screening and clinical assessment was
             applied to a large registry of twins. An expert panel then
             reviewed results from the assessment of 41 subjects whose
             screening results suggested the presence of AD. After review
             of the information at each of four stages of data
             collection, we assessed inter-rater agreement among the
             experts as well as their individual agreement with the final
             consensus diagnosis. We investigated these measures to
             assess the amount and quality, respectively, of new and
             diagnostically useful information that was revealed at each
             stage. A new scheme of weighted differences among the
             available diagnostic categories was developed for these
             analyses. As expected, incremental information from
             successive stages of data collection enabled the panel to
             increase their diagnostic agreement and rates of 'correct'
             diagnoses. Over half of the total information was available,
             however, after review of only the initial telephone
             screening results (stage 1). A brief standardized videotape
             segment of the mental status and neurologic examinations
             provided substantial additional information. We were able to
             compare the final consensus diagnoses with autopsy results
             from seven individuals who had consensus clinical diagnoses
             of Probable or Possible AD (n = 6) or 'demented,
             questionable etiology' (n = 1). All these subjects had
             Definite AD.},
   Key = {fds277074}
}

@article{fds277058,
   Author = {Welsh, KA and Fillenbaum, G and Wilkinson, W and Heyman, A and Mohs, RC and Stern, Y and Harrell, L and Edland, SD and Beekly,
             D},
   Title = {Neuropsychological test performance in African-American and
             white patients with Alzheimer's disease.},
   Journal = {Neurology},
   Volume = {45},
   Number = {12},
   Pages = {2207-2211},
   Year = {1995},
   Month = {December},
   ISSN = {0028-3878},
   url = {http://dx.doi.org/10.1212/wnl.45.12.2207},
   Abstract = {Little information exists on the performance of black versus
             white patients with Alzheimer's disease on
             neuropsychological tests for dementia. In this study, we
             compared performance on the CERAD (Consortium to Establish a
             Registry for Alzheimer's Disease) neuropsychological battery
             between white (n = 830) and black (n = 158) patients with
             Alzheimer's disease enrolled in the CERAD study at 23
             university medical centers in the United States. The black
             patients were older, had fewer years of formal education,
             and were more impaired in their activities of daily living
             than were the white patients. After controlling for these
             characteristics and for duration of the disease and severity
             of dementia, there were differences in the performance of
             black and white patients on several of the cognitive
             measures. Black patients scored lower than whites on tests
             of visual naming and constructional praxis and on the
             Mini-Mental State Examination. There were no statistical
             differences in performance on tests of fluency and word list
             memory. These findings suggest that cultural or experiential
             differences may modify performance on specific
             neuropsychological tests. These factors, in addition to age
             and educational background, should be considered when
             interpreting performance on neuropsychological tests in
             elderly black patients with dementia.},
   Doi = {10.1212/wnl.45.12.2207},
   Key = {fds277058}
}

@article{fds276897,
   Author = {Deguchi, A and Ohnishi, K and Nakagawa, H and Hamaguchi, H and Shiku, H and Deguchi, K},
   Title = {Lipoprotein(A) and effective renal plasma flow rate in older
             patients with arteriosclerotic diseases.},
   Journal = {J Am Geriatr Soc},
   Volume = {43},
   Number = {9},
   Pages = {1067-1068},
   Year = {1995},
   Month = {September},
   ISSN = {0002-8614},
   url = {http://dx.doi.org/10.1111/j.1532-5415.1995.tb05582.x},
   Doi = {10.1111/j.1532-5415.1995.tb05582.x},
   Key = {fds276897}
}

@article{fds276973,
   Author = {Tupler, LA and Welsh, KA and Asare-Aboagye, Y and Dawson,
             DV},
   Title = {Reliability of the Rey-Osterrieth Complex Figure in use with
             memory-impaired patients.},
   Journal = {J Clin Exp Neuropsychol},
   Volume = {17},
   Number = {4},
   Pages = {566-579},
   Year = {1995},
   Month = {August},
   ISSN = {1380-3395},
   url = {http://www.ncbi.nlm.nih.gov/pubmed/7593476},
   Abstract = {Rater reliability was evaluated for the system most widely
             used to assess copy and recall of the Rey Complex Figure:
             the Osterrieth (1944) 18-item scoring system. The study
             sample consisted of 95 subjects (49 males, 46 females), most
             of whom were elderly individuals (M = 59.83, SD = 15.21
             years) suffering from memory impairment. Four raters rated
             copy and delayed-recall protocols, and three raters re-rated
             the protocols after an interval of 3 months. Results
             revealed excellent inter- and intra-rater reliability
             coefficients (.85-.97) for total scores. However,
             reliabilities for the 18 individual items ranged from poor
             (.14) to excellent (.96). Differences in both reliability
             and level of subject performance were observed as a function
             of item and conditions of copy versus recall. It is
             concluded that the Osterrieth scoring system supports
             excellent reliability in use with memory-impaired patients
             using total scores. Nevertheless, individual-item
             reliability would benefit from enhancement, for example, via
             amplified delineation of relevant decision
             criteria.},
   Doi = {10.1080/01688639508405146},
   Key = {fds276973}
}

@article{fds276995,
   Author = {Breitner, JC and Welsh, KA and Gau, BA and McDonald, WM and Steffens,
             DC and Saunders, AM and Magruder, KM and Helms, MJ and Plassman, BL and Folstein, MF},
   Title = {Alzheimer's disease in the National Academy of
             Sciences-National Research Council Registry of Aging Twin
             Veterans. III. Detection of cases, longitudinal results, and
             observations on twin concordance.},
   Journal = {Arch Neurol},
   Volume = {52},
   Number = {8},
   Pages = {763-771},
   Year = {1995},
   Month = {August},
   ISSN = {0003-9942},
   url = {http://www.ncbi.nlm.nih.gov/pubmed/7639628},
   Abstract = {OBJECTIVES: To detect cases of Alzheimer's disease (AD) in a
             large population of twins living throughout the United
             States and to examine concordance for AD in twins as a
             function of age and genotype for apolipoprotein E (APOE).
             SETTING: Nationwide survey. DESIGN: Multistage screening and
             field evaluation beginning with two telephone interviews and
             culminating with laboratory tests, longitudinal
             neuropsychological measures, physician examination, and
             diagnostic consensus among experts. PARTICIPANTS: Membership
             in 1990-1991 of intact pairs in the National Academy of
             Sciences--National Research Council Registry of veteran
             twins, then aged 62 to 73 years. MAIN OUTCOME MEASURES:
             Completeness of case detection was examined in collateral
             studies. Zygosity and APOE genotypes were determined by
             restriction mapping. Concordance was calculated by the
             proband method. RESULTS: Ninety subjects who screened
             positively for AD were studied in person, and 60 whose
             differential diagnoses included AD were followed up, as were
             their co-twins. Sensitivity of screening was estimated at
             greater than 99%, but 24% of subjects refused participation
             after initial screening. Seven of 38 diagnoses of AD have
             been confirmed at autopsy, and 31 other subjects eventually
             met criteria for probable or possible AD (prevalence
             estimate, 0.42%, 95% confidence interval, 0.29% to 0.56%),
             with good interrater reliability (intraclass r = .86).
             Excluding one discordant pair with unknown zygosity,
             concordance rates were 21.1% (4/19) for monozygotic and
             11.1% (2/18) for dizygotic probands. Concordance was 50% for
             twins sharing the epsilon 4/epsilon 4 genotype at APOE, but
             there were no affected co-twins of 15 probands with onset
             before age 70 years, no epsilon 4 allele, and no family
             history of AD. The mean (SD) period of discordance in the
             latter pairs was 11.3 (3.3) years. CONCLUSIONS: The
             multistage case-detection approach achieved reliable and
             valid diagnoses of AD with high apparent sensitivity but
             substantial attrition after initial screening. Genetic
             influences in AD at this age are limited, except among
             homozygotes for allele epsilon 4 at APOE. Subjects with
             early-onset AD who lack the epsilon 4 allele are not rare,
             and their condition appears to have little genetic
             influence. They should be ideal for studies on environmental
             cause of AD.},
   Doi = {10.1001/archneur.1995.00540320035011},
   Key = {fds276995}
}

@article{fds276997,
   Author = {Plassman, BL and Welsh, KA and Helms, M and Brandt, J and Page, WF and Breitner, JC},
   Title = {Intelligence and education as predictors of cognitive state
             in late life: a 50-year follow-up.},
   Journal = {Neurology},
   Volume = {45},
   Number = {8},
   Pages = {1446-1450},
   Year = {1995},
   Month = {August},
   ISSN = {0028-3878},
   url = {http://www.ncbi.nlm.nih.gov/pubmed/7644038},
   Abstract = {We evaluated the relation of education and intelligence in
             early adult life to cognitive function in a group of elderly
             male twins. The Army General Classification Test (AGCT) was
             administered to US armed forces inductees in the early
             1940s. Fifty years later, as part of a study of dementia in
             twins, we tested the cognitive status of 930 of these men
             using the modified Telephone Interview for Cognitive Status
             (TICS-m). TICS-m scores obtained in later life were
             correlated with AGCT scores (r = 0.457) and with years of
             education (r = 0.408). Thus, in univariate analyses, the
             AGCT score accounted for 20.6% and education accounted for
             16.7% of variance in cognitive status. However, these two
             effects were not fully independent. A multivariable model
             using AGCT score, education, and the interaction of the two
             variables as predictors of the TICS-m score explained 24.8%
             of the variance, a slightly but significantly greater
             proportion than was explained by either factor alone. In a
             separate analysis based on 604 pairs of twins who took the
             AGCT, heritability of intelligence (estimated by AGCT score)
             was 0.503. Although this study does not address the issue of
             education and premorbid IQ as risk factors for dementia, the
             findings suggest that basic cognitive abilities in late life
             are related to cognitive performance measures from early
             adult life (ie, education and IQ).},
   Doi = {10.1212/wnl.45.8.1446},
   Key = {fds276997}
}

@article{fds276996,
   Author = {Smith, CJ and Lippiello, PM and Ashford, JW},
   Title = {Smoking, Alzheimer's disease, and confounding with
             genes.},
   Journal = {Lancet},
   Volume = {345},
   Number = {8956},
   Pages = {1054},
   Year = {1995},
   Month = {April},
   url = {http://dx.doi.org/10.1016/s0140-6736(95)90796-3},
   Doi = {10.1016/s0140-6736(95)90796-3},
   Key = {fds276996}
}

@article{fds276993,
   Author = {Plassman, BL and Helms, MJ and Welsh, KA and Saunders, AM and Breitner,
             JC},
   Title = {Smoking, Alzheimer's disease, and confounding with
             genes.},
   Journal = {Lancet},
   Volume = {345},
   Number = {8946},
   Pages = {387},
   Year = {1995},
   Month = {February},
   ISSN = {0140-6736},
   url = {http://www.ncbi.nlm.nih.gov/pubmed/7772122},
   Doi = {10.1016/s0140-6736(95)90374-7},
   Key = {fds276993}
}

@article{fds277134,
   Author = {Breitner, JC and Welsh, KA},
   Title = {Diagnosis and management of memory loss and cognitive
             disorders among elderly persons.},
   Journal = {Psychiatr Serv},
   Volume = {46},
   Number = {1},
   Pages = {29-35},
   Year = {1995},
   Month = {January},
   url = {http://dx.doi.org/10.1176/ps.46.1.29},
   Abstract = {Memory loss and other cognitive dysfunctions, although
             common in elderly persons, are not universal features of old
             age. Instead they herald the presence of various
             neuropsychiatric diseases, which are first recognized as
             syndromes. The two most common neuropsychiatric syndromes,
             dementia and delirium, produce global changes in cognition
             and other capacities. They are differentiated by the
             patient's level of consciousness, which is impaired in
             delirium but intact in dementia. Delirium is generally
             reversible and often indicates serious physical illness.
             Although dementia is occasionally reversible, the mainstays
             of its management and treatment are palliative. Comorbid
             psychiatric symptoms are common in patients with both
             delirium or dementia, and their recognition and treatment
             constitute an important task for the geropsychiatrist. The
             differential diagnosis of primary dementing illness and
             other psychiatric illnesses such as depression is complex,
             because symptoms of the two kinds of disorders often coexist
             and common pathogenetic mechanisms may underlie both
             syndromes.},
   Doi = {10.1176/ps.46.1.29},
   Key = {fds277134}
}

@article{fds276941,
   Author = {Breitner, JC and Welsh, KA},
   Title = {Genes and recent developments in the epidemiology of
             Alzheimer's disease and related dementia.},
   Journal = {Epidemiol Rev},
   Volume = {17},
   Number = {1},
   Pages = {39-47},
   Year = {1995},
   url = {http://dx.doi.org/10.1093/oxfordjournals.epirev.a036184},
   Abstract = {In the search for cause or prevention of Alzheimer's
             disease, the traditional aims of analytic epidemiology have
             been hindered by several technical difficulties. The
             heterogeneous genetic influences on Alzheimer's disease have
             probably contributed substantially to difficulties in the
             detection of host or environmental factors associated with
             modified disease risk. We have discussed five areas of
             improved technical or theoretical approach, each of which
             has materially improved the prospects for future success in
             this endeavor. Improved methods of diagnosis have yielded
             purer samples of "cases" and have made it more practical to
             undertake population-based studies. Genetic determinants of
             Alzheimer's disease risk are being understood with
             increasing sophistication. A growing recognition of the
             time-dependent nature of the Alzheimer process has led to
             new and heuristically valuable ways of thinking about the
             disease, its causes (genetic and otherwise), and its
             prevention. While there is a growing consensus that
             Alzheimer's disease is probably not one disease but several,
             the limited success of efforts to identify distinguishable
             phenotypes has largely given way to the identification of
             those with various measures of disease risk and (probably)
             mechanisms on the basis of identifiable genotypic variation.
             Thus, several forms of this disease may soon be segregated
             for separate analysis. Two of the predisposing genes have
             apparent implications for disease pathogenesis; others
             remain identified only as anonymous "loci" implicated by
             linkage analyses. The greater understanding of genetic
             mechanisms serves to enable not only studies of gene effects
             in pathogenesis, but also the influence of various
             environmental factors that may modify the effects.(ABSTRACT
             TRUNCATED AT 250 WORDS)},
   Doi = {10.1093/oxfordjournals.epirev.a036184},
   Key = {fds276941}
}

@article{fds276942,
   Author = {Breitner, JC and Welsh, KA and Helms, MJ and Gaskell, PC and Gau, BA and Roses, AD and Pericak-Vance, MA and Saunders, AM},
   Title = {Delayed onset of Alzheimer's disease with nonsteroidal
             anti-inflammatory and histamine H2 blocking
             drugs.},
   Journal = {Neurobiol Aging},
   Volume = {16},
   Number = {4},
   Pages = {523-530},
   Year = {1995},
   ISSN = {0197-4580},
   url = {http://www.ncbi.nlm.nih.gov/pubmed/8544901},
   Abstract = {Factors that modify onset of Alzheimer's disease (AD) may be
             revealed by comparing environmental exposures in affected
             and unaffected members of discordant twin pairs or sibships.
             Among siblings at high risk of AD, sustained use of
             nonsteroidal anti-inflammatory drugs (NSAIDs) was associated
             with delayed onset and reduced risk of AD. After adjustment
             for use of NSAIDs, there was minimal effect on onset with
             reported history of any of three common illnesses
             (arthritis, diabetes, or acid-peptic disease). However,
             independent of exposure to NSAIDs, onset was unexpectedly
             delayed in those reporting extended use of histamine H2
             blocking drugs. Randomized clinical trials will be needed to
             affirm the utility of these drugs for prevention, but the
             present findings may have implications for pathogenesis:
             because NSAIDs block the calcium-dependent postsynaptic
             cascade that induces excitotoxic cell death in NMDA-reactive
             neurons, and because histamine potentiates such events,
             excitotoxicity may deserve additional investigation in
             AD.},
   Doi = {10.1016/0197-4580(95)00049-k},
   Key = {fds276942}
}

@article{fds276937,
   Author = {Reed, T and Carmelli, D and Swan, GE and Breitner, JC and Welsh, KA and Jarvik, GP and Deeb, S and Auwerx, J},
   Title = {Lower cognitive performance in normal older adult male twins
             carrying the apolipoprotein E epsilon 4 allele.},
   Journal = {Arch Neurol},
   Volume = {51},
   Number = {12},
   Pages = {1189-1192},
   Year = {1994},
   Month = {December},
   ISSN = {0003-9942},
   url = {http://dx.doi.org/10.1001/archneur.1994.00540240033012},
   Abstract = {OBJECTIVE: Given the strong association of the
             apolipoprotein E (apoE) allele epsilon 4 with late-onset
             Alzheimer dementia or multi-infarct dementia, we tested
             whether normal older adult men with at least one epsilon 4
             allele demonstrate subclinical changes in cognition and
             perform more poorly on tests of cognitive function compared
             with subjects without the epsilon 4 allele. DESIGN:
             Matched-pair design of normal adult male (average age, 63
             years) fraternal twins. SETTING: Subjects voluntarily
             participated on an outpatient basis at a research or medical
             center facility. PARTICIPANTS: Members of the National
             Heart, Lung, and Blood Institute twin panel third
             examination previously genotyped for apoE. MAIN OUTCOME
             MEASURE: Education-adjusted scores on several
             neuropsychological tests were compared in twins discordant
             for the apoE epsilon 4 allele. Subjects with documented
             cerebrovascular disease were excluded. RESULTS: Among 20
             fraternal twin pairs discordant for the presence of epsilon
             4, twins with the epsilon 4 allele demonstrated poorer mean
             performance than their co-twins without the epsilon 4
             allele. This relationship was also noted cross-sectionally
             where age- and education-adjusted scores of 50 individual
             twin subjects with at least one epsilon 4 allele
             demonstrated poorer performance compared with 138 individual
             twins without an epsilon 4 allele. CONCLUSIONS: The apoE
             epsilon 4 allele may be associated with decreased cognitive
             function in discordant twin pairs. Our results suggest that
             epsilon 4 may represent a potential marker for accelerated
             cognitive aging and such individuals may be at greater risk
             for development of late-onset Alzheimer dementia or
             multi-infarct dementia.},
   Doi = {10.1001/archneur.1994.00540240033012},
   Key = {fds276937}
}

@article{fds276938,
   Author = {Norton, MC and Breitner, JC and Welsh, KA and Wyse,
             BW},
   Title = {Characteristics of nonresponders in a community survey of
             the elderly.},
   Journal = {J Am Geriatr Soc},
   Volume = {42},
   Number = {12},
   Pages = {1252-1256},
   Year = {1994},
   Month = {December},
   url = {http://dx.doi.org/10.1111/j.1532-5415.1994.tb06506.x},
   Abstract = {OBJECTIVE: To identify predictors of nonresponse in a
             community survey of cognitive status in the elderly. DESIGN:
             Cross-sectional community survey with two stages of
             recruitment: an initial, less-intensive method, followed by
             a more aggressive approach that included face-to-face
             contact. Characteristics of initial nonresponders and
             responders were compared. SETTING: A close-knit rural
             community with higher than usual proportions of elderly,
             especially the very old. Subjects were interviewed in their
             homes. Collateral informants were subsequently interviewed
             by telephone. PARTICIPANTS: Utah heads of household aged 75
             and older who resided in a noninstitutionalized setting.
             MEASUREMENTS: Mini-Mental State Examination (MMSE), Dementia
             Questionnaire, and an autobiographical risk factor and
             family history questionnaire provided measures for all
             independent variables. The dependent variable was status as
             initial responders or initial nonresponders. RESULTS: An
             initial participation rate of 63% was achieved, but a final
             rate of 93% was achieved when initial nonresponders were
             contacted later face-to-face. MMSE score was significantly
             related to responder status when analyzed alone (beta =
             -.19, P = 0.02) and remained a significant predictor after
             adjusting for education and whether born in Cache County
             (beta = -.16, P = 0.041) or current drinking, diabetes, or
             "other" health problems (beta = -.18, P = 0.028). After
             controlling for the informant report of subject's problems
             with activities of daily living, MMSE score fell just below
             statistical significance (beta = -.16, P = 0.079).
             CONCLUSIONS: Nonresponders in community surveys of the
             elderly appear to be disproportionately cognitively
             impaired. The increase in participation rates achieved after
             more persistent recruitment suggests that many initial
             nonresponders can still be recruited if intensive methods
             are used.},
   Doi = {10.1111/j.1532-5415.1994.tb06506.x},
   Key = {fds276938}
}

@article{fds277053,
   Author = {Welsh, KA and Hoffman, JM and Earl, NL and Hanson,
             MW},
   Title = {Neural correlates of dementia: regional brain metabolism
             (FDG-PET) and the CERAD neuropsychological
             battery.},
   Journal = {Arch Clin Neuropsychol},
   Volume = {9},
   Number = {5},
   Pages = {395-409},
   Year = {1994},
   Month = {October},
   ISSN = {0887-6177},
   url = {http://www.ncbi.nlm.nih.gov/pubmed/14589655},
   Abstract = {The present Investigation examined the biological correlates
             of the cognitive deficits of Alzheimer's disease and related
             dementias using the neuropsychological assessment battery of
             the Consortium to Establish a Registry of Alzheimer's
             Disease (CERAD) and positron emission tomography (PET).
             Resting state cerebral glucose metabolism was measured using
             the labelled radiotracer, [18F] Fluoro-2-deoxyglucose (FDG),
             in a sample of patients with mild to moderate dementia (n =
             66). Specific and predictable relationships were seen
             between regional brain metabolism (left and right, frontal,
             temporal, and parietal lobes) and the neuropsychological
             measures of verbal fluency, constructional praxis, and
             verbal list learning. On tests of naming and delayed verbal
             recall only diminished FDG uptake in the left frontal lobe
             and the left temporal lobe, respectively, approached
             significance. This study demonstrates the expected
             relationships between neuropsychological performance and
             regional cerebral metabolism, thereby providing support for
             the CERAD battery as a valid measure in the clinical
             evaluation of dementia and for the use of FDG-PET in
             brain-behavior studies of dementia.},
   Doi = {10.1093/arclin/9.5.395},
   Key = {fds277053}
}

@article{fds276983,
   Author = {Siegler, IC and Dawson, DV and Welsh, KA},
   Title = {Caregiver ratings of personality change in Alzheimer's
             disease patients: a replication.},
   Journal = {Psychol Aging},
   Volume = {9},
   Number = {3},
   Pages = {464-466},
   Year = {1994},
   Month = {September},
   ISSN = {0882-7974},
   url = {http://www.ncbi.nlm.nih.gov/pubmed/7999331},
   Abstract = {Caregivers of 26 patients with Alzheimer's disease (AD)
             rated current and premorbid personality patterns with the
             NEO Personality Inventory. Results replicated previous
             findings on the degree of change reported in a previous
             group of patients with mixed memory disorder diagnoses.
             After a diagnosis of AD, the patients were rated as
             significantly more neurotic, less extraverted, less open,
             and less conscientious. There were no rated differences of
             changes in the personality domain of Agreeableness. These
             results strengthen the usefulness of caregiver ratings of
             personality change of patients with memory problems who
             cannot be useful informants on their own
             behalf.},
   Doi = {10.1037//0882-7974.9.3.464},
   Key = {fds276983}
}

@article{fds277056,
   Author = {Fillenbaum, GG and Wilkinson, WE and Welsh, KA and Mohs,
             RC},
   Title = {Discrimination between stages of Alzheimer's disease with
             subsets of Mini-Mental State Examination items. An analysis
             of Consortium to Establish a Registry for Alzheimer's
             Disease data.},
   Journal = {Arch Neurol},
   Volume = {51},
   Number = {9},
   Pages = {916-921},
   Year = {1994},
   Month = {September},
   ISSN = {0003-9942},
   url = {http://www.ncbi.nlm.nih.gov/pubmed/8080392},
   Abstract = {OBJECTIVE: To identify minimal sets of Mini-Mental State
             Examination (MMSE) items that can distinguish normal control
             subjects from patients with mild Alzheimer's disease (AD),
             patients with mild from those with moderate AD, and those
             with moderate from those with severe AD. DESIGN: Two
             randomly selected equivalent half samples. Results of
             logistic regression analysis from data from the first half
             of the sample were confirmed by receiver operating
             characteristic curves on the second half. SETTING: Memory
             disorders clinics at major medical centers in the United
             States affiliated with the Consortium to establish a
             Registry for Alzheimer's Disease (CERAD). PARTICIPANTS:
             White, normal control subjects (n = 412) and patients with
             AD (n = 621) who met CERAD criteria; nonwhite subjects (n =
             165) and persons with missing data (n = 27) were excluded.
             MAIN OUTCOME MEASURES: Three four-item sets of MMSE items
             that discriminate, respectively, (1) normal controls from
             patients with mild AD, (2) patients with mild from those
             with moderate AD, and (3) patients with moderate from those
             with severe AD. RESULTS: The MMSE items discriminating
             normal controls from patients with mild AD were day, date,
             recall of apple, and recall of penny; those discriminating
             patients with mild from those with moderate AD were month,
             city, spelling world backward, and county, and those
             discriminating patients with moderate from those with severe
             AD were floor of building, repeating the word table, naming
             watch, and folding paper in half. Performance on the first
             two four-item sets was comparable with that of the full
             MMSE; the third set distinguished patients with moderate
             from those with severe AD better than chance. CONCLUSIONS: A
             minimum set of MMSE items can effectively discriminate
             normal controls from patients with mild AD and between
             successive levels of severity of AD. Data apply only to
             white patients with AD. Performance in minorities, more
             heterogeneous groups, or normal subjects with questionable
             cognitive status has not been assessed.},
   Doi = {10.1001/archneur.1994.00540210088017},
   Key = {fds277056}
}

@article{fds277057,
   Author = {Welsh, KA and Butters, N and Mohs, RC and Beekly, D and Edland, S and Fillenbaum, G and Heyman, A},
   Title = {The Consortium to Establish a Registry for Alzheimer's
             Disease (CERAD). Part V. A normative study of the
             neuropsychological battery.},
   Journal = {Neurology},
   Volume = {44},
   Number = {4},
   Pages = {609-614},
   Year = {1994},
   Month = {April},
   ISSN = {0028-3878},
   url = {http://dx.doi.org/10.1212/wnl.44.4.609},
   Abstract = {The neuropsychological tests developed for the Consortium to
             Establish a Registry for Alzheimer's Disease (CERAD) are
             currently used to measure cognitive impairments of
             Alzheimer's disease (AD) in clinical investigations of this
             disorder. This report presents the normative information for
             the CERAD battery, obtained in a large sample (n = 413) of
             control subjects (ages 50 to 89) who were enrolled in 23
             university medical centers in the United States
             participating in the CERAD study from 1987 to 1992. We
             compared separately the performance of subjects with high (>
             or = 12) and low (< 12) years of formal education. For many
             of the individual cognitive measures in the highly educated
             group, we observed significant age and gender effects. Only
             the praxis measure showed a significant age effect in the
             low-education group. Delayed recall, when adjusted for
             amount of material acquired (savings), was relatively
             unaffected by age, gender, and level of education. Our
             findings suggest that the savings scores, in particular, may
             be useful in distinguishing between AD and normal
             aging.},
   Doi = {10.1212/wnl.44.4.609},
   Key = {fds277057}
}

@article{fds276994,
   Author = {Breitner, JC and Gau, BA and Welsh, KA and Plassman, BL and McDonald,
             WM and Helms, MJ and Anthony, JC},
   Title = {Inverse association of anti-inflammatory treatments and
             Alzheimer's disease: initial results of a co-twin control
             study.},
   Journal = {Neurology},
   Volume = {44},
   Number = {2},
   Pages = {227-232},
   Year = {1994},
   Month = {February},
   ISSN = {0028-3878},
   url = {http://www.ncbi.nlm.nih.gov/pubmed/8309563},
   Abstract = {We conducted a co-twin control study among 50 elderly twin
             pairs with onsets of Alzheimer's disease (AD) separated by 3
             or more years. Twenty-three male pairs (46%) were screened
             from the (U.S.) National Academy of Sciences-National
             Research Council Registry (NAS-NRC Registry) of World War II
             veteran twins; others (mostly women) had responded to
             advertisements or were referred from AD clinics. Twenty-six
             pairs (52%) were monozygous. The onset of AD was inversely
             associated with prior use of corticosteroids or ACTH (odds
             ratio [OR], 0.25; 95% confidence interval [CI], 0.06 to
             0.95; p = 0.04). Similar but weaker trends were present
             among pairs discordant for history of arthritis or for prior
             daily use of nonsteroidal anti-inflammatory drugs (NSAIDs)
             or aspirin. The association was strongest when we combined
             use of steroids/ACTH or NSAIDs post hoc into a single
             variable of anti-inflammatory drugs (AIs) (OR, 0.24; CI,
             0.07 to 0.74; p = 0.01). The inverse relation was strong in
             female (volunteer) twin pairs but was not present in the
             younger men from the NAS-NRC Registry. AIs had typically
             been taken for arthritis or related conditions, but a
             similar result was apparent after controlling statistically
             for the arthritis variable (OR, 0.08; CI, 0.01 to 0.69; p =
             0.02). AIs have been proposed as a means of retarding the
             progression of AD symptoms, and these data suggest that AIs
             may also prevent or delay the initial onset of AD symptoms.
             Because of limitations in the case-control method, our
             results require corroboration with hypothesis-driven
             research designed to control bias and confounding.},
   Doi = {10.1212/wnl.44.2.227},
   Key = {fds276994}
}

@article{fds276992,
   Author = {Plassman, BL and Newman, TT and Welsh, KA and Helms, M and Breitner,
             JCS},
   Title = {Properties of the telephone interview for cognitive status:
             Application in epidemiological and longitudinal
             studies},
   Journal = {Neuropsychiatry, Neuropsychology and Behavioral
             Neurology},
   Volume = {7},
   Number = {3},
   Pages = {235-241},
   Year = {1994},
   Month = {January},
   ISSN = {0894-878X},
   Abstract = {We evaluated the utility of telephone screening for dementia
             in epidemiologic research by comparing performance on the
             modified Telephone Interview for Cognitive Status (TICS-m)
             with results from in-person neuropsychological measures in
             67 elderly males. Longitudinal performance on the TICS-m was
             also evaluated over an average of 15 months in the same
             subjects. After comprehensive clinical evaluation, subjects
             were assigned to one of three diagnostic groups: normal,
             demented, or "mild-ambiguous" cognitive syndrome. As
             expected, the normal group scored highest on the TICS-m,
             followed in turn by the mild-ambiguous and demented groups.
             Among various neuropsychological measures, the Mini-Mental
             State Examination correlated most strongly with the TICS-m.
             The scores on the first and second administration of the
             TICS-m were significantly correlated for both the normal and
             demented groups. The normal and mild-ambiguous groups showed
             slight improvement on the second administration of the
             TICS-m, but the demented group showed a slight decline in
             their scores. Thus, the TICS-m is able to detect dementia
             and decline in cognitive function over time, and therefore
             appears useful for population studies as an economical
             alternative to standard in-person screening. © 1994 Raven
             Press, Ltd., New York.},
   Key = {fds276992}
}

@article{fds276939,
   Author = {Breitner, JC and Welsh, KA and Robinette, CD and Gau, BA and Folstein,
             MF and Brandt, J},
   Title = {Alzheimer's disease in the NAS-NRC registry of aging twin
             veterans. II. Longitudinal findings in a pilot series.
             National Academy of Sciences. National Research Council
             Registry.},
   Journal = {Dementia},
   Volume = {5},
   Number = {2},
   Pages = {99-105},
   Year = {1994},
   url = {http://dx.doi.org/10.1159/000106703},
   Abstract = {Over 3 years we followed 8 pairs of male twins one or both
             of whom had suspected Alzheimer's disease (AD) including
             'mild/ambiguous' changes suggestive of incident AD. These
             pairs were screened in 1988 and 1989 from 339 pairs in the
             (US) National Academy of Sciences-National Research Council
             Registry (NASR) of aging veteran twins, then 61-72 years of
             age. Most of the suspected cases (10 of 12) had
             mild/ambiguous changes. Including these subjects, we had
             estimated the prevalence of AD in the NASR as about 2%. We
             now describe briefly the longitudinal evaluation of these 8
             pairs. Only 1 of the 10 individuals with mild/ambiguous
             changes has progressed to show well-defined clinical
             symptoms of AD. Two others remain in their original research
             category, while 7 clearly do not have AD. Thus, we now
             estimate the 1988-1989 prevalence of AD in the NASR as 0.5%.
             These results contrast with other follow-up studies of mild
             cases from a university-based Alzheimer's clinic. We suggest
             that the contrasting findings reflect the nature of the
             samples studied, and we show that the present results are
             predicted by Bayesian reasoning.},
   Doi = {10.1159/000106703},
   Key = {fds276939}
}

@article{fds276940,
   Author = {Welsh, KA and Ballard, E and Nash, F and Raiford, K and Harrell,
             L},
   Title = {Issues affecting minority participation in research studies
             of Alzheimer disease.},
   Journal = {Alzheimer Dis Assoc Disord},
   Volume = {8},
   Number = {Suppl. 4},
   Pages = {38-48},
   Year = {1994},
   Abstract = {Despite the need for minority subjects in research studies
             of Alzheimer disease (AD), the successful involvement of
             minority patients in such studies has been difficult. This
             report discusses the many societal, economic, logistical,
             and attitudinal barriers that have inhibited the
             participation of minority patients and their families in
             medical research programs of AD. Special consideration is
             given to the unique cultural issues that arise when
             conducting studies involving African-American elderly
             subjects. Methods are considered for overcoming the barriers
             to participation gleaned from the national study CERAD
             (Consortium to Establish a Registry of Alzheimer Disease)
             and other investigations of AD. Recommendations are made for
             future research programs targeted on the specific health
             care needs and concerns of the minority segments of our
             population.},
   Key = {fds276940}
}

@article{fds277131,
   Author = {Brandt, J and Welsh, KA and Breitner, JC and Folstein, MF and Helms, M and Christian, JC},
   Title = {Hereditary influences on cognitive functioning in older men.
             A study of 4000 twin pairs.},
   Journal = {Arch Neurol},
   Volume = {50},
   Number = {6},
   Pages = {599-603},
   Year = {1993},
   Month = {June},
   url = {http://dx.doi.org/10.1001/archneur.1993.00540060039014},
   Abstract = {OBJECTIVE: To determine the contribution of genetic factors
             to cognitive functioning in older men. DESIGN: Cognitive
             testing by telephone interview in an epidemiologically
             defined population. PARTICIPANTS: 2077 monozygotic and 2225
             dizygotic male twin pairs, all between the ages of 62 and 73
             years, recruited from the National Academy of Sciences twin
             registry. MAIN OUTCOME MEASURES: The Telephone Interview for
             Cognitive Status--Modified total score and factor scores
             were analyzed. The Falconer heritability statistic and
             maximum likelihood estimates of genetic and environmental
             components were computed. RESULTS: Heritability of the total
             Telephone Interview for Cognitive Status--Modified score was
             estimated to be 30%. Shared environmental effects accounted
             for an additional 18% of the variance; most of this was
             related to years of education. Of the four cognitive factors
             derived, the language/attention factor had the highest
             heritability estimate. CONCLUSIONS: Genetic factors and
             educational achievement together account for almost half of
             the variance in the cognitive functioning of older men.
             Studies of the genetics of dementing illnesses need to
             consider the degree to which cognitive capacities are
             themselves under genetic control.},
   Doi = {10.1001/archneur.1993.00540060039014},
   Key = {fds277131}
}

@article{fds277132,
   Author = {Breitner, JC and Gatz, M and Bergem, AL and Christian, JC and Mortimer,
             JA and McClearn, GE and Heston, LL and Welsh, KA and Anthony, JC and Folstein, MF},
   Title = {Use of twin cohorts for research in Alzheimer's
             disease.},
   Journal = {Neurology},
   Volume = {43},
   Number = {2},
   Pages = {261-267},
   Year = {1993},
   Month = {February},
   url = {http://dx.doi.org/10.1212/wnl.43.2.261},
   Abstract = {The causes of Alzheimer's disease (AD) remain a mystery
             despite the recent identification of several putative
             environmental risk factors and the discovery of several
             linked genetic loci and point mutations associated with the
             disease. Particularly uncertain is the generalizability of
             the genetic findings to the common forms of disease
             encountered in clinical practice or population research.
             Twin studies of AD can illuminate causal mechanisms, both
             genetic and environmental. This consensus document explores
             the rationale for such twin studies, as well as a number of
             methodologic problems that render them difficult to
             implement or interpret. We review existing twin studies of
             AD and note several ambitious new studies. Finally, we
             delineate several practical strategies for the near future
             of twin research in AD.},
   Doi = {10.1212/wnl.43.2.261},
   Key = {fds277132}
}

@article{fds276936,
   Author = {Welsh, KA and Hoffman, JM and McDonald, WM and Earl, NL and Breitner,
             JCS},
   Title = {Concordant but Different: Cognitive Function, Cerebral
             Anatomy, and Metabolism in Monozygotic Twins With
             Alzheimer's Disease},
   Journal = {Neuropsychology},
   Volume = {7},
   Number = {2},
   Pages = {158-171},
   Publisher = {American Psychological Association (APA)},
   Year = {1993},
   Month = {January},
   ISSN = {0894-4105},
   url = {http://dx.doi.org/10.1037/0894-4105.7.2.158},
   Abstract = {To illustrate the utility of the twin method in Alzheimer's
             disease (AD) research, we studied in detail a pair of
             monozygotic twins concordant for the disease but markedly
             different in their clinical presentations.
             Neuropsychological evaluation, magnetic resonance brain
             imaging, and cerebral glucose metabolic studies revealed a
             typical behavioral presentation for AD in Twin A. In
             contrast, Twin B showed prominent visuospatial impairments.
             Although there was no identified cause for the disparate
             presentations, a close correspondence was observed between
             the neuropsychological findings and the regional brain
             measures. The results suggest that the trajectory of AD may
             vary widely even in genetically identical individuals.
             Factors accounting for the variability include potential
             intrauterine, early developmental, and environmental
             differences.},
   Doi = {10.1037/0894-4105.7.2.158},
   Key = {fds276936}
}

@article{fds277133,
   Author = {Welsh, KA and Breitner, JCS and Magruder-Habib,
             KM},
   Title = {Detection of dementia in the elderly using telephone
             screening of cognitive status},
   Journal = {Neuropsychiatry, Neuropsychology and Behavioral
             Neurology},
   Volume = {6},
   Number = {2},
   Pages = {103-110},
   Year = {1993},
   Month = {January},
   Abstract = {Detection of dementia in large, geographically dispersed
             populations is difficult. Conventional in-person
             neuropsychological assessment techniques, no matter how
             brief, are too costly to be practical for this purpose.
             Telephone interviewing is an obvious alternative for
             cognitive screening, but its practical utility is relatively
             unexplored. We therefore investigated the performance
             characteristics of a telephone screen for dementia in
             elderly residents of congregate housing facilities. We
             interviewed 209 subjects using the Telephone Interview for
             Cognitive Status (TICS) and a modified version (TICS-m) that
             includes items sensitive to early dementia (delayed recall)
             and eliminates other items difficult to verify in survey
             work. After the subjects received a brief in-person
             neuropsychological assessment, TICS and TICS-m scores were
             compared as predictors of the resulting clinical assignment
             (normal, mildly impaired, or demented). Although the TICS-m
             yielded slightly better results, both versions of the
             instrument were sensitive and specific indicators of
             dementia in this community sample. In a separate exercise,
             both instruments also correctly identified 17 clinic
             patients with carefully diagnosed Probable AD. Telephone
             interviewing of cognitive function may therefore provide an
             economical approach to mental status screening in research
             studies where in-person assessment is impractical. © 1993
             Raven Press, Ltd., New York.},
   Key = {fds277133}
}

@article{fds277129,
   Author = {Welsh, KA and Butters, N and Hughes, JP and Mohs, RC and Heyman,
             A},
   Title = {Detection and staging of dementia in Alzheimer's disease.
             Use of the neuropsychological measures developed for the
             Consortium to Establish a Registry for Alzheimer's
             Disease.},
   Journal = {Arch Neurol},
   Volume = {49},
   Number = {5},
   Pages = {448-452},
   Year = {1992},
   Month = {May},
   url = {http://dx.doi.org/10.1001/archneur.1992.00530290030008},
   Abstract = {Our earlier studies using the Consortium to Establish a
             Registry of Alzheimer's Disease neuropsychological battery
             showed that delayed recall was a highly sensitive indicator
             of early Alzheimer's disease. None of the learning and
             memory measures in the battery were found to be useful in
             staging the severity of this form of dementia. This study
             explores the nonmemory functions (fluency, naming, and
             praxis) of the Consortium to Establish a Registry of
             Alzheimer's Disease battery and asks whether performance on
             any of these measures adds to the detection of early
             Alzheimer's disease or is sensitive to the later progression
             of the illness. We stratified patients with this disease
             according to severity (mild, moderate, severe), and compared
             them with age-, education-, and gender-matched control
             subjects (group N = 49 each). Multivariate procedures and
             cutting scores were used to determine the efficacy of the
             various measures in distinguishing between the cases and
             control subjects. Impairment of delayed recall was again
             found to be the best discriminator for detecting mild cases
             of Alzheimer's disease. Confrontation naming was the only
             nonmemory factor that assisted in this discrimination. For
             staging the illness, a combination of measures including
             fluency, praxis, and recognition memory best differentiated
             cases with mild dementia from those with either moderate or
             severe stages of disease. Measures of delayed recall quickly
             "bottomed out" in the patients with Alzheimer's disease and
             proved of little value in staging the disorder.(ABSTRACT
             TRUNCATED AT 250 WORDS)},
   Doi = {10.1001/archneur.1992.00530290030008},
   Key = {fds277129}
}

@article{fds277126,
   Author = {Pericak-Vance, MA and Bebout, JL and Gaskell, PC and Yamaoka, LH and Hung, WY and Alberts, MJ and Walker, AP and Bartlett, RJ and Haynes, CA and Welsh, KA},
   Title = {Linkage studies in familial Alzheimer disease: evidence for
             chromosome 19 linkage.},
   Journal = {Am J Hum Genet},
   Volume = {48},
   Number = {6},
   Pages = {1034-1050},
   Year = {1991},
   Month = {June},
   ISSN = {0002-9297},
   url = {http://www.ncbi.nlm.nih.gov/pubmed/2035524},
   Abstract = {A genetic component in the etiology of Alzheimer disease
             (AD) has been supported by indirect evidence for several
             years, with autosomal dominant inheritance with
             age-dependent penetrance being suggested to explain the
             familial aggregation of affecteds. St. George Hyslop et al.
             reported linkage of familial AD (FAD) in four early-onset
             families (mean age at onset [M] less than 50 years).
             Subsequent studies have been inconsistent in their results;
             Goate et al. also reported positive lod scores. However,
             both Pericak-Vance et al.'s study of a series of mainly
             late-onset FAD families (M greater than 60 years) and
             Schellenberg et al.'s study failed to confirm linkage to
             chromosome 21 (CH21). These various studies suggest the
             possibility of genetic heterogeneity, with some families
             linked to CH21 and others unlocalized. Recently, St. George
             Hyslop et al. extended their analysis to include additional
             families. The extended analyses supported their earlier
             finding of linkage to CH21, while showing strong evidence of
             heterogeneity between early-onset (M less than 65 years) and
             late-onset (M greater than 60 years) FAD families. Because
             our families did not show linkage to CH21, we undertook a
             genomic search for an additional locus for FAD. Because of
             both the confounding factor of late age at onset of FAD and
             the lack of clear evidence of Mendelian transmission in some
             of our families, we employed the affected-pedigree-member
             (APM) method of linkage analysis as an initial screen for
             possible linkage. Using this method, we identified two
             regions suggesting linkage: the proximal long arm of
             chromosome 19 (CH19) and the CH21 region of FAD linkage
             reported by St. George Hyslop et al. Application of standard
             likelihood (LOD score) analysis to these data support the
             possibility of an FAD gene locate on CH19, particularly in
             the late-onset FAD families. These data further suggest
             genetic heterogeneity and delineate this region of CH19 as
             an area needing additional investigation in
             FAD.},
   Key = {fds277126}
}

@article{fds277127,
   Author = {Welsh, K and Butters, N and Hughes, J and Mohs, R and Heyman,
             A},
   Title = {Detection of abnormal memory decline in mild cases of
             Alzheimer's disease using CERAD neuropsychological
             measures.},
   Journal = {Arch Neurol},
   Volume = {48},
   Number = {3},
   Pages = {278-281},
   Year = {1991},
   Month = {March},
   url = {http://dx.doi.org/10.1001/archneur.1991.00530150046016},
   Abstract = {The present study was designed to determine which of the
             memory tasks included in the CERAD (Consortium to Establish
             a Registry for Alzheimer's Disease) neuropsychological
             battery best differentiate patients with early Alzheimer's
             disease from cognitively normal elderly control subjects and
             also best distinguish between the various levels of severity
             of the dementia process. A sample of CERAD patients with
             Alzheimer's disease was stratified by disease severity into
             those with mild, moderate, or severe dementia and matched
             with control subjects for sex, age, and education. Using
             multivariate procedures and cutting scores, the efficacy of
             each memory measure in distinguishing between these groups
             and control subjects was determined. The test for delayed
             recall was found to be the best overall discriminatory
             measure. The other tests of memory, ie, immediate recall,
             intrusion errors, and recognition memory, had poor overall
             discriminability. None of the CERAD memory measures were
             found to be particularly powerful in staging the severity of
             dementia. These findings suggest that tests for delayed
             recall may be particularly useful in the early detection of
             Alzheimer's disease and should be considered in screening
             batteries for dementia in community surveys.},
   Doi = {10.1001/archneur.1991.00530150046016},
   Key = {fds277127}
}

@article{fds277128,
   Author = {Schellenberg, GD and Pericak-Vance, MA and Wijsman, EM and Moore, DK and Gaskell, PC and Yamaoka, LA and Bebout, JL and Anderson, L and Welsh,
             KA and Clark, CM},
   Title = {Linkage analysis of familial Alzheimer disease, using
             chromosome 21 markers.},
   Journal = {Am J Hum Genet},
   Volume = {48},
   Number = {3},
   Pages = {563-583},
   Year = {1991},
   Month = {March},
   ISSN = {0002-9297},
   url = {http://www.ncbi.nlm.nih.gov/pubmed/1998342},
   Abstract = {Chromosome 21 markers were tested for linkage to familial
             Alzheimer disease (FAD) in 48 kindreds. These families had
             multiple cases of Alzheimer disease (AD) in 2 or more
             generations with family age-at-onset means (M) ranging from
             41 to 83 years. Included in this group are seven Volga
             German families which are thought to be genetically
             homogeneous with respect to FAD. Autopsy documentation of AD
             was available for 32 families. Linkage to the 21 q11-q21
             region was tested using D21S16, D21S13, D21S110, D21S1/S11,
             and the APP gene as genetic markers. When linkage results
             for all the families were summed, the LOD scores for these
             markers were consistently negative and the entire region was
             formally excluded. Linkage results were also summed for the
             following family groups; late-onset (M greater than 60),
             early-onset (M less than or equal to 60), Volga Germans (M =
             56), and early-onset non-Volga Germans (M less than or equal
             to 60). For the first three groups, LOD scores were negative
             for this region. For the early-onset non-Volga German group
             (six families), small positive LOD scores of Zmax = 0.78
             (recombination fraction theta = .15), Zmax = 0.27 (theta =
             .15), and Zmax = 0.64 (theta = .0), were observed for
             D21S13, D21S16, and D21S110, respectively. The remainder of
             the long arm of chromosome 21 was tested for linkage to FAD
             using seven markers spanning the q22 region. Results for
             these markers were also predominantly negative. Thus it is
             highly unlikely that a chromosome 21 gene is responsible for
             late-onset FAD and at least some forms of early-onset FAD
             represented by the Volga German kindreds.},
   Key = {fds277128}
}

@article{fds277130,
   Author = {Welsh, KA},
   Title = {Detection of early dementia in the elderly.},
   Journal = {Experimental Aging Research},
   Volume = {17},
   Number = {2},
   Pages = {101-},
   Year = {1991},
   Key = {fds277130}
}

@article{fds277150,
   Author = {Siegler, IC and Welsh, KA and Dawson, DV and Fillenbaum, GG and Earl,
             NL and Kaplan, EB and Clark, CM},
   Title = {Ratings of personality change in patients being evaluated
             for memory disorders.},
   Journal = {Alzheimer Dis Assoc Disord},
   Volume = {5},
   Number = {4},
   Pages = {240-250},
   Year = {1991},
   ISSN = {0893-0341},
   url = {http://www.ncbi.nlm.nih.gov/pubmed/1781966},
   Abstract = {Caregivers of 35 mildly to moderately memory-impaired
             patients rated current and premorbid personalities with the
             NEO Personality Inventory. We then examined changes in the
             five domains of personality tapped by the NEO. There were
             significant changes in four of the five domains of normal
             personality functioning toward less conscientiousness, lower
             extraversion, higher neuroticism, and lower openness. The
             difference toward lower agreeableness was not significant
             when controlling for multiple comparisons. Spearman rank
             correlation coefficients indicated that changes in
             conscientiousness and vulnerability were not related to
             rated premorbid personality patterns and thus appear to
             describe shifts for all patients evaluated for memory
             disorders. These data suggest that personality inventories
             may be helpful in characterizing caregivers' observations of
             memory-impaired patients and thus represent a critical
             source of information for the clinician in charge of
             care.},
   Doi = {10.1097/00002093-199100540-00003},
   Key = {fds277150}
}

@article{fds276935,
   Author = {Breitner, JCS and Welsh, KA and Magruder-Habib, KM and Churchill, CM and Robinette, CD and Folslein, MF and Murphy, EA and Priolo, CC and Brandt,
             J},
   Title = {Alzheimer’s disease in the national academy of sciences
             registry of aging twin veterans: I. Pilot
             investigations},
   Journal = {Dementia and Geriatric Cognitive Disorders},
   Volume = {1},
   Number = {6},
   Pages = {297-303},
   Publisher = {S. Karger AG},
   Year = {1990},
   Month = {January},
   url = {http://dx.doi.org/10.1159/000107157},
   Abstract = {The (US) National Academy of Sciences Registry of aging twin
             veterans contains 15.924 pairs of white male twins born
             between 1917 and 1927. About 8.000 pairs are living today.
             In preparation for a study of Alzheimer’s disease (AD) in
             this Registry, we investigated 829 members of pairs living
             in North and South Carolina. Virginia. in the District of
             Columbia, and Maryland. A brief telephone interview for
             cognitive symptoms was administered to 678 (91.4%) of 742
             subjects located. Cognitive dysfunction was identified
             initially in 124 individuals (18.3%). whose clinical
             histories were then obtained over the telephone. Results
             suggested that 108 subjects did not have AD. Ten (62.5%) of
             the remaining 16 subjects underwent diagnostic evaluation
             for dementia. One had cerebrovascular disease with concident
             depression. Two others with probable AD were a monozygotic
             (MZ) pair. The remaining 7 subjects had possible AD or a
             mild but progressive cognitive disorder suggestive of early
             AD. Upon subsequent examination, 3 of 4 MZ co-twins showed
             significant symptoms that had not been detected during
             screening procedures. None of 3 dizygotic co-twin showed any
             significant abnormality. These results suggest: (1) that a
             combination of mailed information, telephone interviews, and
             clinical examination provides a feasible means of detecting
             AD in the Registry. (2) that about 150 pairs with
             presumptive AD in 1 or both subjects will be identified in a
             full study of the Registry, (3) that concordance for AD in
             MZ pairs may exceed prior estimates, but (4) that such rates
             of concordance are apparent only upon detailed evaluation of
             apparently normal co-twins as well as their impaired
             brothers. Longitudinal observation of pairs with apparently
             affected individuals will be required for definitive
             conclusions. © 1990 S. Karger AG, Basel.},
   Doi = {10.1159/000107157},
   Key = {fds276935}
}

@article{fds277125,
   Author = {Pericak-Vance, MA and Bebout, JL and Yamaoka, LA and Gaskell, PC and Hung, WY and Alberts, MJ and Clark, CM and Haynes, CS and Welsh, KA and Earl, NL and Heyman, A and Roses, AD},
   Title = {Linkage studies in familial Alzheimer's disease, application
             of the affected pedigree member (APM) method of linkage
             analysis},
   Journal = {Molecular biology and genetics of Alzheimer's disease:
             proceedings of the International Symposium on Dementia:
             Molecular Biology and Genetics of Alzheimer's Disease.
             ICS884},
   Pages = {215-228},
   Year = {1990},
   Month = {January},
   Key = {fds277125}
}

@article{fds276934,
   Author = {Gold, PE and Welsh, KA},
   Title = {Regional brain catecholamines and memory: effects of
             footshock, amygdala implantation, and stimulation.},
   Journal = {Behav Neural Biol},
   Volume = {47},
   Number = {2},
   Pages = {116-129},
   Year = {1987},
   Month = {March},
   ISSN = {0163-1047},
   url = {http://dx.doi.org/10.1016/s0163-1047(87)90215-9},
   Abstract = {Previous findings have revealed a correlation between
             post-training release of whole brain norepinephrine (NE) and
             later retention performance. The present experiment examined
             changes after a training footshock in NE levels, as well as
             the levels of the major central NE metabolite,
             3-methoxy-4-hydroxyphenylglycol (MHPG), dopamine (DA), and
             epinephrine (EPI) in eight brain regions. Brain levels of
             these amines and the metabolite were assessed 10 min after
             training in a one-trial inhibitory (passive) avoidance task.
             The results indicate that NE levels decreased significantly
             in neocortex, neostriatum, hypothalamus, frontal pole,
             septum, and brainstem, but not in hippocampus or thalamus.
             The decreases in NE levels were generally accompanied by
             increases in MHPG; the MHPG/NE ratio increased significantly
             in all areas in which decreases in NE were observed. DA
             levels decreased in neostriatum and increased in neocortex
             and brainstem. Epinephrine levels decreased only in the
             brainstem sample. Thus, the effects of training on NE are
             widespread, probably reflecting the release of the amine in
             most brain regions. Such findings are consistent with the
             view that posttraining release of brain NE may modulate the
             storage of new information in many brain regions. One
             especially potent treatment for modulating memory storage is
             electrical stimulation of the amygdala. Therefore, we also
             examined the effects of amygdala implantation and
             stimulation on brain catecholamine levels to determine
             whether such changes might be correlated with the effects of
             amygdala stimulation on memory. The results indicate that
             electrode implantation into the amygdala results in
             pervasive changes in NE levels in most brain regions tested.
             Against this modified baseline, the results of training and
             electrical stimulation were region specific and very
             difficult to interpret. The major conclusion which can be
             derived from this portion of the experiment is that the
             amygdala damage produced by electrode implantation produces
             a brain which is substantially different from that of intact
             animals.},
   Doi = {10.1016/s0163-1047(87)90215-9},
   Key = {fds276934}
}

@article{fds276932,
   Author = {Welsh, KA and Gold, PE},
   Title = {Epinephrine proactive retardation of amygdala-kindled
             epileptogenesis.},
   Journal = {Behav Neurosci},
   Volume = {100},
   Number = {2},
   Pages = {236-245},
   Year = {1986},
   Month = {April},
   ISSN = {0735-7044},
   url = {http://dx.doi.org/10.1037//0735-7044.100.2.236},
   Abstract = {Recent findings indicate that a single injection (ip) of
             epinephrine can proactively retard the development of
             amygdala-kindled seizures. In these experiments, these
             findings were extended by examining the dose and temporal
             properties of this phenomenon. Rats were prepared with
             stimulating electrodes placed in the amygdala. At 30 min or
             24, 48, or 72 hr before the first kindling trial, the
             animals received an epinephrine injection (0.01-1.0 mg/kg).
             The results indicate that the high epinephrine dose delayed
             the onset of kindling when injected within 24 hr before
             kindling, the intermediate dose delayed kindling when
             injected at 30 min before the first kindling trial, and the
             low dose was ineffective at all intervals. Injections
             administered after kindling had been established had no
             effect on later seizures, although injections during early
             kindling stages had a small effect on the further
             development of seizures. Examination of the time courses of
             increases in plasma catecholamine levels, blood pressure,
             and heart rate demonstrated that long-term changes in these
             measures cannot account for the retardation of kindling. The
             findings of these experiments indicate that although
             epinephrine does not directly affect seizure production with
             these procedures, the hormone does have long-lasting
             proactive effects on kindled epileptogenesis.},
   Doi = {10.1037//0735-7044.100.2.236},
   Key = {fds276932}
}

@article{fds276933,
   Author = {Welsh, KA and Gold, PE},
   Title = {Brain catecholamines and memory modulation: effects of
             footshock, amygdala implantation, and stimulation.},
   Journal = {Behav Neural Biol},
   Volume = {43},
   Number = {2},
   Pages = {119-131},
   Year = {1985},
   Month = {March},
   ISSN = {0163-1047},
   url = {http://dx.doi.org/10.1016/s0163-1047(85)91317-2},
   Abstract = {The results of previous studies indicate that the extent of
             a transient decline in brain norepinephrine (NE) levels
             shortly after training and administration of any of several
             memory modulating treatments is correlated with later
             retention performance. The present experiment assessed such
             changes after one-trial inhibitory (passive) avoidance
             training and, in addition, measured concentration changes in
             3-methoxy-4-hydroxyphenylglycol (MHPG), the major metabolite
             of brain NE, as well as dopamine (DA) and epinephrine (EPI)
             levels. The results indicate that the decreases in brain NE
             after footshock are accompanied by an increase in MHPG, thus
             providing additional evidence that brain NE is released
             after training. DA levels were unchanged after training;
             brainstem EPI levels increased after the training footshock,
             but forebrain EPI levels were unchanged. A second experiment
             examined brain catecholamine levels in animals which
             received post-training electrical stimulation of the
             amygdala. The findings of this experiment indicate that the
             amygdala damage which accompanies electrode implantation
             apparently results in a chronic change in whole brain NE
             levels and metabolism. After amygdala, NE concentrations in
             both brainstem and forebrain samples were reduced by 20% and
             MHPG was increased by 22-34%. Furthermore, NE levels were
             not responsive to training in implanted animals. Thus, brain
             NE levels after training were not predictive of retention
             performance in amygdala-implanted or -stimulated animals.
             However, the significance of such findings for understanding
             the possible role of central NE in memory storage is
             complicated by the severe modification of the dynamics of
             brain aminergic systems in animals bearing amygdala
             electrodes.},
   Doi = {10.1016/s0163-1047(85)91317-2},
   Key = {fds276933}
}

@article{fds276931,
   Author = {Welsh, KA and Gold, PE},
   Title = {Attenuation of epileptogenesis: proactive effect of a single
             epinephrine injection of amygdaloid kindling.},
   Journal = {Behav Neural Biol},
   Volume = {40},
   Number = {2},
   Pages = {179-185},
   Year = {1984},
   Month = {March},
   ISSN = {0163-1047},
   url = {http://dx.doi.org/10.1016/s0163-1047(84)90279-6},
   Abstract = {Repeated daily electrical stimulation of the amygdala can
             lead to a progressive increase in brain and behavioral
             seizures. This phenomenon, termed kindling, has been viewed
             as a model for epileptogenesis. The results reported here
             demonstrate that a single systemic epinephrine injection can
             significantly retard such epileptogenesis for a period of at
             least several days. These findings suggest that peripheral
             catecholamines, responding either to stress near the time of
             seizure initiation or to treatments administered at that
             time, may be important in regulating the development of
             epileptic states. In addition, the results indicate that an
             acute episode of high plasma epinephrine levels may result
             in a durable modification of brain function.},
   Doi = {10.1016/s0163-1047(84)90279-6},
   Key = {fds276931}
}

@article{fds277124,
   Author = {Welsh, KA and Gold, PE},
   Title = {Age-related changes in brain catecholamine responses to a
             single footshock.},
   Journal = {Neurobiol Aging},
   Volume = {5},
   Number = {1},
   Pages = {55-59},
   Year = {1984},
   ISSN = {0197-4580},
   url = {http://dx.doi.org/10.1016/0197-4580(84)90086-1},
   Abstract = {The responses of forebrain and brainstem catecholamine
             levels to a single footshock were studied in 70-day,
             one-year, and two-year-old Fischer-344 rats. Brain
             catecholamine concentrations were assessed 10 minutes after
             a single 2 second footshock (0, 0.3, or 2.0 mA). In samples
             taken from non-footshocked rats, only forebrain dopamine
             concentrations showed a significant age-related decline.
             However, because the net weights of both the forebrain and
             the brainstem samples increased significantly with age, the
             content of forebrain dopamine did not exhibit a significant
             decline. Both norepinephrine and dopamine levels showed
             age-related changes in responsiveness to footshock.
             Norepinephrine concentrations were reduced in both the
             forebrain and brainstem samples obtained 10 minutes after
             the high footshock in both the 70-day and one-year-old
             animals. In two-year-old rats, however, neither forebrain
             nor brainstem norepinephrine concentrations were altered in
             response to footshock. Seventy-day-old rats demonstrated
             significant footshock-induced increases in brainstem
             dopamine levels, one-year olds showed no appreciable change,
             and two-year olds demonstrated a non-significant
             footshock-induced decrease. Thus, both noradrenergic and
             dopaminergic systems demonstrated age-related changes in
             their responsiveness to a single brief footshock. These
             alterations may contribute to the declining ability of the
             senescent animals to adapt to stressful situations.},
   Doi = {10.1016/0197-4580(84)90086-1},
   Key = {fds277124}
}


%% Papers Published   
@article{fds135091,
   Title = {Silverman, D.H.S., Small, G.W., Chang, C.Y., Lu, C.S., Kung
             de Aburto, M.A., Chen, W., Czernin, J., Rapoport, S.I.,
             Pietrini, P., Alexander, G.E., Schapiro, M.B., Jagust, W.J.,
             Hoffman, J.M., Welsh-Bohmer, K.A., ....& Phelps, M.E. (in
             press). Neuroimaging in evaluation of dementia: Regional
             brain metabolism and long-term outcome. Journal of the
             American Medical Association.},
   Year = {2001},
   Key = {fds135091}
}

@article{fds135092,
   Title = {Welsh-Bohmer, K.A. & Madden, D.J. (accepted). Benign
             Senescent Forgetfulness, Age-Associated Memory Impairment,
             and Age-related cognitive decline. In: J.Copeland, M.
             Abou-Saleh, & D. Blazer (Eds). Principles and Practice of
             Geriatric Psychiatry- 2nd Edition. Sussex, England: John
             Wiley & Sons, Ltd.},
   Year = {2001},
   Key = {fds135092}
}

@article{fds135093,
   Title = {Publications in print},
   Year = {2001},
   Key = {fds135093}
}

@article{fds135094,
   Title = {Sugarman, J., Cain, C., Wallace, R., & Welsh-Bohmer, K.A.
             (2001). How proxies make decisions about research for
             patients with Alzheimer's disease. Journal of the American
             Geriatrics Society, 49, 1110-1119.},
   Year = {2001},
   Key = {fds135094}
}

@article{fds135122,
   Title = {Yi-Ju, L., Scott, W.K., Hedges, D.J., Zhang, F., Gaskell,
             P.C., Nance, M.A., Watts, R.L., Hubble, J.P., Koller, W.C.,
             Pahwa, R., Stern, M.B., Hiner, B.C., Jankovic, J., Allen
             Jr., F.A., Goetz, C.G., Mastaglia, F., Stajich, J.M.,
             Gibson, R.A., Middleton, L.T., Saunders, A.M., Scott, B.L.,
             Small, G.W., Reed, A.D., Schmechel, D.E., Welsh-Bohmer,
             K.A., Conneally, P.M., Roses, A.D., Gilbert, J.R., Vance,
             J.M., Haines, J.L., Pericak-Vance, M.A.. (accepted). Onset
             in neurodegenerative diseases is genetically controlled.
             American Journal of Human Genetics.},
   Year = {2001},
   Key = {fds135122}
}

@article{fds135123,
   Title = {Schiffman, S.S., Graham, B.G., Sattely-Miller, E.A.,
             Zervakis, J., Welsh-Bohmer, K.A. (in press). Taste, smell,
             and neuropsychological performance of individuals at risk
             for Alzheimer's disease. Neurobiology of
             Aging.},
   Year = {2001},
   Key = {fds135123}
}

@article{fds135124,
   Title = {Steffens, D.C., Payne, M.E., Greenberg, D.L., Byrum, C.E.,
             Welsh-Bohmer, K.A., Wagner, H.R., & MacFall, J.R.
             (accepted). Hippocampal volume and incident dementia in
             geriatric depression. American Journal of Geriatric
             Psychiatry.},
   Year = {2001},
   Key = {fds135124}
}

@article{fds135125,
   Title = {Welsh-Bohmer, K.A., Koltai, D.C., & Mason D.J. (accepted).
             The clinical utility of neuropsychological evaluation of
             patients with known or suspected dementia. In: G. Prigatano
             & N. Pliskin (eds). Demonstrating Utility and Cost
             Effectiveness in Clinical Neuropsychology. Philadelphia:
             Psychology Press- Taylor & Francis Group.},
   Year = {2001},
   Key = {fds135125}
}

@article{fds135126,
   Title = {Carlson, M., Tschanz, J., Norton, M., Welsh-Bohmer, K.A.,
             Martin, B., Breitner, J.C.S. (accepted). H2 Histamine
             receptor blockage in the treatment of Alzheimer's disease: A
             randomized, double-blind, placebo-controlled trial of
             nizatidine. Alzheimer's Disease and Associated
             Disorders.},
   Year = {2001},
   Key = {fds135126}
}

@article{fds135127,
   Title = {Koltai, D.C., Welsh-Bohmer, K.A., & Mason, D.J. (accepted).
             Neuropsychological consultation and training of family
             members with dementia. In: G. Prigatano & N. Pliskin (eds).
             Demonstrating Utility and Cost Effectiveness in Clinical
             Neuropsychology. Philadelphia: Psychology Press- Taylor &
             Francis Group.},
   Year = {2001},
   Key = {fds135127}
}

@article{fds135128,
   Title = {Fillenbaum, G.G., Unverzagt, F.W., Ganguli, M.,
             Welsh-Bohmer, K.A. (accepted). The CERAD neuropsychological
             battery: Performance of representative community and
             tertiary care samples of African-American and European
             American elderly. In: F.R. Ferraro (ed). Minority and
             Cross-Cultural Aspects of Neuropsychological Assessment.
             London: Swets & Zeitlineger.},
   Year = {2001},
   Key = {fds135128}
}

@article{fds135129,
   Title = {Dawson, D.V., Welsh-Bohmer, K.A., Siegler, I.C. (accepted).
             Premorbid personality predicts level of rated personality
             change in AD patients. Alzheimer's Disease and Associated
             Disorders.},
   Year = {2001},
   Key = {fds135129}
}

@article{fds135130,
   Title = {Meich, R.A, Breitner, J.C.S., Zandi, P.P., Khachaturian,
             A.S., Anthony, J.C., Mayer, L., for the Cache County Study
             Group. (2002). Incidence of AD may decline in the early 90's
             for men, later for women: The Cache County Study. Neurology,
             58:209-218.},
   Year = {2001},
   Key = {fds135130}
}

@article{fds135131,
   Title = {Carlson, M.C., Zandi, P.A., Plassman, B.L., Tschanz, J.T.,
             Welsh-Bohmer, K.A. Steffens, D.C., Bastian, L, Mehta, K.M.,
             & Breitner, J.C.S. for the Cache County Study Group (2002).
             Hormone Replacement Therapy Predicts Improved Cognitive
             Trajectory In Older Women. The Cache County Study.
             Neurology, 57: 2210-2216},
   Year = {2001},
   Key = {fds135131}
}

@article{fds135132,
   Title = {Koltai, D.C., Welsh-Bohmer, K.A., & Schmechel, D.E. (2001).
             Influence of anosognosia on treatment outcome among dementia
             patients. Neuropsychological Rehabilitation, 11:
             455-475},
   Year = {2001},
   Key = {fds135132}
}

@article{fds135133,
   Title = {Dawson, D.V., Welsh-Bohmer, K.A., & Siegler, I.C. (2001).
             Informant rated personality change in Alzheimer's disease
             patients: Replication, influence of premorbid profile, and
             covariate relationships. Research and Practice in
             Alzheimer's Disease. Volume 5. In: B. Vellas & J. Fitten
             (Eds). Auzeville-Tolosane France: Serdi Press, pp
             27-32.},
   Year = {2001},
   Key = {fds135133}
}

@article{fds135134,
   Title = {Grundman, M., Kim, H.T., Salmon, D., Storandt, M., Smith,
             G., Ferris, S., Mohs, R., Kawas, C., Doody, R.,
             Welsh-Bohmer, K.A.,...Kukull, W., Thal, L.J. for
             participants in the Alzheimer's Disease Centers'
             Neuropsychological Database Initiative (2001). The
             Alzheimer's Disease Centers' Neuropsychological Database
             Initiative: A Resource for Alzheimer's Disease Prevention
             Trials. In: K. Iqbal, S.S. Sisodia, and B. Winblad (eds)
             Alzheimer's disease: Advances in etiology, pathogenesis and
             therapeutics. The Proceedings of the 7th International
             conference on Alzheimer's Disease and Related Disorders."
             London: John Wiley & Son, pp129-140.},
   Year = {2001},
   Key = {fds135134}
}

@article{fds135135,
   Title = {Welsh-Bohmer, K.A., Hulette, C., Schmechel, D., Burke, J. &
             Saunders, A. (2001). Neuropsychological detection of
             preclinical Alzheimer's disease: Results of a
             neuropathological series of "normal" controls. In: K. Iqbal,
             S.S. Sisodia, and B. Winblad (eds) Alzheimer's disease:
             Advances in etiology, pathogenesis and therapeutics. The
             Proceedings of the 7th International conference on
             Alzheimer's Disease and Related Disorders." London: John
             Wiley & Sons, pp 111-122.},
   Year = {2001},
   Key = {fds135135}
}

@article{fds135090,
   Title = {Welsh-Bohmer, K.A. & Ogrocki, P. K. (1998). Clinical
             differentiation of memory disorders in neurodegenerative
             diseases. In: A.I. Troster (Ed), Memory in Neurodegenerative
             Disease: Biological, Cognitive, and Clinical Perspectives.
             Cambridge University Press, London, pp 290313.},
   Year = {1998},
   Key = {fds135090}
}

@article{fds135120,
   Title = {Ogrocki, P.K. & Welsh-Bohmer, K.A. (accepted). Assessment of
             Cognitive and Functional Impairment in the Elderly. In:
             C.Clark (Ed) Neurodegenerative Dementia: Clinical Features
             and Pathological Mechanisms. New York: McGraw
             Hill.},
   Year = {1998},
   Key = {fds135120}
}

@article{fds135121,
   Title = {Koltai, D.C. & Welsh-Bohmer, K.A. (accepted). Geriatric
             Neuropsychological Assessment. In: R.D. Vanderploeg (Ed)
             Clinician's Guide to Neuropsychological Assessment, 2nd
             Edition. New Jersey: Lawrence Erlbaum Associates.},
   Year = {1998},
   Key = {fds135121}
}

@article{fds135096,
   Title = {Welsh-Bohmer, K.A., Gearing, M., Saunders, A.M., Roses,
             A.D., & Mirra, S.M. (1997). Apolipoprotein E genotypes in a
             neuropathological series from the Consortium to Establish a
             Registry for Alzheimer's Disease (CERAD). Annals of
             Neurology, 42: 319-325.},
   Year = {1997},
   Key = {fds135096}
}

@article{fds135087,
   Title = {Fillenbaum, G.G., Peterson, B., Welsh-Bohmer, K.A., Kukull,
             W., Heyman, A. (accepted) Progression of Alzheimer's disease
             in African-American and White patients: The CERAD
             experience, Part XVI. Neurology.},
   Year = {1996},
   Key = {fds135087}
}

@article{fds135088,
   Title = {Welsh-Bohmer, K.A. & Hoffman, J.M. (1996). Positron Emission
             Tomography Neuroimaging in Dementia. In: E.Bigler (Ed),
             Handbook of Human Brain Function. Plenum Press: New York,
             pp. 185-222.},
   Year = {1996},
   Key = {fds135088}
}

@article{fds135089,
   Title = {Siegler, I., Poon, L., Madden, D., & Welsh, K.A. (1996).
             Psychological Aspects of Normal Aging. In: E.W. Busse & D.G.
             Blazer (Eds). Geriatric Psychiatry, 2nd Edition. Washington
             D.C.: American Psychatric Press, pp 105-127.},
   Year = {1996},
   Key = {fds135089}
}

@article{fds135095,
   Title = {Yamaoka, L.H., Welsh-Bohmer, K.A., Hulette, C.M. Gaskell
             Jr., P.C., Murray, M., Rimmler, J.L., Helms, B.R., Guerra,
             M., Roses, A.D., Schmechel, D.E., Pericak-Vance, M.A.
             (1996). Linkage of Frontotemporal dementia to Chromosome 17:
             Clinical and neuropathological characterization of
             phenotype. American Journal of Human Genetics, 59,
             1306-1312.},
   Year = {1996},
   Key = {fds135095}
}

@article{fds135113,
   Title = {Hulette, C., Welsh-Bohmer, K.A., Crain, B., Szymanski, M.H.,
             Sinclaire, N.O.,& Roses, A.D. (accepted). Rapid brain
             autopsy. The Bryan Alzheimer's Disease Research Center
             Experience. Journal of Neuropathology. Plassman, B.L.,
             Welsh-Bohmer, K.A., Bigler, E.D., Johnson, S.C., Anderson,
             C.V., Helms, M.J., Breitner, J.C.S. (in press).
             Apolipoprotein E e4 and hippocampal volume in twins with
             normal cognition. Neurology},
   Year = {1996},
   Key = {fds135113}
}

@article{fds135114,
   Title = {Yamaoka, L.H., Welsh-Bohmer, K.A., Hulette, C.M. Gaskell
             Jr., P.C., Murray, M., Rimmler, J.L., Helms, B.R., Guerra,
             M., Roses, A.D., Schmechel, D.E., Pericak-Vance, M.A.
             Linkage of Frontotemporal dementia to Chromosome 17:
             Clinical and neuropathological characterization of
             phenotype. American Journal of Human Genetics.
             (accepted).},
   Year = {1996},
   Key = {fds135114}
}

@article{fds135115,
   Title = {Steffens, D.C., Plassman B.L., Helms M.J., Welsh, K.A.,
             Saunders A.M., & Breitner J.C.S. A twin study of late-onset
             depression and apolipoprotein E e4 as risk factors for
             Alzheimer's disease. Biological Psychiatry
             (accepted).},
   Year = {1996},
   Key = {fds135115}
}

@article{fds135116,
   Title = {Davidson M., Harvey P.D., Welsh K.A., Powchik P., Putnam
             K.M., Mohs R.C. (1996). Defining the dementia of old age
             schizophrenia: Psychometric comparisons of schizophrenic
             patients to patients with Alzheimer's disease. American
             Journal of Psychiatry, 153, 1274-1279.},
   Year = {1996},
   Key = {fds135116}
}

@article{fds135117,
   Title = {Hoffman, J.M., Hanson, M.W., Welsh, K.A., Earl, N.L., Paine,
             S., Delong, D. & Coleman, R.E. (1996). Interpretation
             variability of 18F-fluoro-2-deoxyglucose (FDG) Positron
             Emission Tomography (PET) studies in dementia. Investigative
             Radiology, 31, 316-322.},
   Year = {1996},
   Key = {fds135117}
}

@article{fds135118,
   Title = {Saunders, A.M., Hulette, C., Welsh-Bohmer, K.A., Schmechel,
             D.E., Crain, B., Burke, J.R., Alberts, M.A., Strittmatter,
             W.J., Breitner, J.C.S., Earl, N., Clark, C., Heyman, A.,
             Gaskell, P.C., Pericak-Vance, M.A., & Roses, A.D. (1996).
             Specificity and sensitivity of apolipoprotein E genotyping
             in a prospectively ascertained series of probable Alzheimer
             disease patients with autopsy-confirmed diagnoses. Lancet,
             348, 90-93.},
   Year = {1996},
   Key = {fds135118}
}

@article{fds135119,
   Title = {Steffens, D.C., Welsh, K.A., Burke, J.R., Helms, M.J.,
             Folstein, M.F., Brandt, J., McDonald, W.M., & Breitner,
             J.C.S. (1996). Diagnosis of Alzheimer's disease in
             epidemiological studies by staged review of clinical data.
             Neuropsychiatry, Neuropsychology, and Behavioral Neurology,
             9, 107-113.},
   Year = {1996},
   Key = {fds135119}
}

@article{fds135084,
   Title = {1995},
   Year = {1995},
   Key = {fds135084}
}

@article{fds135085,
   Title = {Tupler, L.A., Welsh, K.A., Asare-Aboagye, Y., Dawson, D.V.
             (1995). Reliability of the Rey-Osterrieth Complex Figure in
             use with memory impaired patients. Journal of Clinical and
             Experimental Neuropsychology, 17, 566-579.},
   Year = {1995},
   Key = {fds135085}
}

@article{fds135086,
   Title = {Plassman, B.L., Saunders, A.M., Helms, M.J., Breitner JCS, &
             Welsh, K.A.,. (1995). Smoking, Alzheimer's disease, and
             confounding with genes. Author's reply. Lancet, 345, 1054
             .},
   Year = {1995},
   Key = {fds135086}
}

@article{fds135107,
   Title = {Welsh, K.A., Fillenbaum, G., Wilkinson, W., Heyman, A.,
             Mohs, R.C., Stern, Y., & Harrell, L. (1995).
             Neuropsychological performance of black and white patients
             with Alzheimer's disease. Neurology, 45:
             2207-2211.},
   Year = {1995},
   Key = {fds135107}
}

@article{fds135108,
   Title = {Breitner, J.C.S. & Welsh, K.A. (1995). Genes and recent
             developments in the epidemiology of Alzheimer's disease and
             related dementia. Epidemiology Reviews, 17,
             39-47.},
   Year = {1995},
   Key = {fds135108}
}

@article{fds135109,
   Title = {Plassman, B.L., Welsh, K.A., Helms, M., Brandt, J., Page,
             W.F., & Breitner, J.C.S. (1995) Intelligence and education
             as predictor of cognitive state in late life: A 50 year
             follow-up. Neurology, 45: 1446-1450.},
   Year = {1995},
   Key = {fds135109}
}

@article{fds135110,
   Title = {Breitner, J.C.S., Welsh, K.A., Gau, B.A., McDonald WM,
             Steffens DC, Saunders AM, Magruder KM, Helms MJ, Blassman
             BL, Folstein MF, Brandt J, Robinette CD, Page WF. (1995).
             Alzheimer's disease in the National Academy of Sciences-
             National Research Council Registry of Aging Twin Veterans.
             III. Detection of Cases, longitudinal results, and
             observations on twin concordance. Archives of Neurology, 52,
             763-771.},
   Year = {1995},
   Key = {fds135110}
}

@article{fds135111,
   Title = {Plassman, B.L., Breitner, J.C.S., Helms, M.J., Welsh, K.A.,
             & Saunders, A.M. (1995) Confounding with genes may explain
             the inverse association of smoking and Alzheimer's disease.
             Lancet, 345, 387.},
   Year = {1995},
   Key = {fds135111}
}

@article{fds135112,
   Title = {Breitner, J.C.S. & Welsh, K.A. (1995). An approach to
             diagnosis and management of memory loss and other cognitive
             syndromes of aging. Psychiatric Services: A Journal of the
             American Psychiatric Association , 46, 29-35.},
   Year = {1995},
   Key = {fds135112}
}

@article{fds135080,
   Title = {1994},
   Year = {1994},
   Key = {fds135080}
}

@article{fds135081,
   Title = {Welsh K.A., Ballard E., Nash F., Raiford K., & Harrell L.
             (1994). Issues affecting minority participation in studies
             of Alzheimer's disease. Alzheimer's Disease and Associated
             Disorders, 8 (Suppl 4), 38-48.},
   Year = {1994},
   Key = {fds135081}
}

@article{fds135082,
   Title = {Reed T., Swan G., Carmeli D., Christian, J., Breitner
             J.C.S., Welsh, K.A. (1994). Lower cognitive performance in
             normal older adult male twins carrying the Apolipoprotein E
             e4 allele. Archives of Neurology, 51, 1189-1192.},
   Year = {1994},
   Key = {fds135082}
}

@article{fds135083,
   Title = {Welsh KA, Butters N, Mohs RC, Beekly D, Edland S, Fillenbaum
             G, Heyman A: The consortius to establish a registry of
             Alzheimer's disease (CERAD) Part V: A normative study of the
             neuropsychological battery. Neurology 44: 609-614,
             1994.},
   Year = {1994},
   Key = {fds135083}
}

@article{fds135100,
   Title = {Norton M.C., Breitner, J.C.S., Welsh, K.A., & Wyse, B.W.
             (1994). Characteristics of nonresponders in a community
             survey of the elderly. Journal of the American Geriatrics
             Society, 42: 1252-1256.},
   Year = {1994},
   Key = {fds135100}
}

@article{fds135101,
   Title = {Fillenbaum, G. G., Burchett, B.M., & Welsh, K.A. The 20-item
             word list as a measure of cognitive functioning in the
             Health and Retirement Survey: Norms and validity for White,
             Black, and Hispanic respondents. Health and Retirement Study
             Working Paper Series, Paper #94-005. Institute for Social
             Research at the University of Michigan & the National
             Institute of Aging.},
   Year = {1994},
   Key = {fds135101}
}

@article{fds135102,
   Title = {Welsh K.A., Hoffman J.M., Earl N.L., & Hansen, M. W. (1994)
             Neural correlates of dementia: Regional brain metabolism
             (FDG-PET) and the CERAD neuropsychological battery. Archives
             of Clinical Neuropsychology, 9, 395-409.},
   Year = {1994},
   Key = {fds135102}
}

@article{fds135103,
   Title = {Welsh K.A., Butters N., Mohs R.C., Beekly D., Edland S.,
             Fillenbaum, G, & Heyman A.},
   Year = {1994},
   Key = {fds135103}
}

@article{fds135104,
   Title = {Welsh KA, Hoffman JM, Earl NL, Hansen MW: Neural correlates
             of dementia: Regional brain metabolism (FDG-PET) and the
             CERAD neuropsychological battery. Archives of Clinical
             Neuropsychology 9: 395-409, 1994.},
   Year = {1994},
   Key = {fds135104}
}

@article{fds135105,
   Title = {Breitner JCS, Welsh KA, Robinette CD, Gau BA, Folstein MF,
             Brandt J: Alzheimer's disease in the National Academy of
             Sciences Registry of Aging Twin Veterans II. Longitudinal
             findings in a pilot series. Dementia 5: 99-105,
             1994.},
   Year = {1994},
   Key = {fds135105}
}

@article{fds135106,
   Title = {Breitner JCS, Gau B, Welsh KA, Plassman B, Helms M, Anthony
             J: Inverse association betrween anti-inflammatory agents and
             Alzheimer's disease: Initial results of a co-twin control
             study. Neurology 44: 227-232, 19994.},
   Year = {1994},
   Key = {fds135106}
}

@article{fds135136,
   Title = {Welsh K.A., Hoffman J.M., Earl N.L., & Hansen, M. W. (1994)
             Neural correlates of dementia: Regional brain metabolism
             (FDG-PET) and the CERAD neuropsychological battery. Archives
             of Clinical Neuropsychology, 9, 395-409.},
   Year = {1994},
   Key = {fds135136}
}

@article{fds135077,
   Title = {1993},
   Year = {1993},
   Key = {fds135077}
}

@article{fds135078,
   Title = {Welsh K.A., Breitner J.C.S., & Magruder-Habib, K.M. (1993).
             Detection of dementia in community volunteers using
             telephone screening of cognitive status. Neuropsychiatry,
             Neuropsychology, and Behavioral Neurology, 6,
             103-110.},
   Year = {1993},
   Key = {fds135078}
}

@article{fds135079,
   Title = {Welsh K.A., Hoffman J.M., McDonald J.M., Earl N.L., &
             Breitner J.C.S. (1993). Concordant but different: Cognitive
             function, cerebral anatomy, and brain metabolism in
             monozygotic twins with Alzheimer's disease. Neuropsychology,
             7, 158-171.},
   Year = {1993},
   Key = {fds135079}
}

@article{fds135098,
   Title = {Brandt J., Welsh K.A., Breitner J.C.S., Folstein, M.F.,
             Helms, M, & Christian, J.C. (1993). Hereditary influences on
             cognitive functioning in older men: A study of 4,000
             twin-pairs. Archives of Neurology, 50, 599-603.},
   Year = {1993},
   Key = {fds135098}
}

@article{fds135099,
   Title = {Breitner, J.C.S., Gatz, M., Bergem A.L.M., Christian J.C.,
             Mortimer, J.A., McClearn, G.E., Heston L.L., Welsh, K.A.,
             Anthony, J.C., Folstein, M.F., & Radebaugh, T.S. (1993). The
             use of Twin Cohorts for Research in Alzheimer's Disease: A
             consensus document sponsored by the NIA and Duke University
             Collaborative Twin Studies. Neurology, 43,
             261-267.},
   Year = {1993},
   Key = {fds135099}
}

@article{fds135076,
   Title = {Welsh, K.A., Butters, N., Hughes, J.P., Mohs, R.C., and
             Heyman, A. (1991) Detection of abnormal memory decline in
             mild Alzheimer\'s disease using CERAD neuropsychological
             measures. Archives of Neurology, 48, 278-281.},
   Year = {1991},
   Key = {fds135076}
}

@article{fds135097,
   Title = {Hulette, C.M., Welsh-Bohmer, K.A., Murray, M.G., Mash, D.&
             McIntyre, L.M. Neuropathological and neuropsychological
             changes in "normal" aging: Evidence for preclinical
             Alzheimer's disease. Journal of Neuropathology and
             Experimental Neurology, 57, 1168-1174.},
   Key = {fds135097}
}


%% Chapters in Books   
@misc{fds364315,
   Author = {Welsh-Bohmer, KA and Johnson, S},
   Title = {Neuropsychological Assessment of Dementia},
   Pages = {59-87},
   Booktitle = {Handbook of Dementing Illnesses, Second Edition},
   Year = {2006},
   Month = {January},
   ISBN = {9780824758387},
   url = {http://dx.doi.org/10.3109/9780849354847-8},
   Abstract = {Department of Psychiatry and Joseph and Kathleen Bryan
             Alzheimer’s Disease Research Center-Division of Neurology,
             Department of Medicine, Duke University Medical Center,
             Durham, North Carolina, U.S.A. Neuropsychological assessment
             plays an important part in the differential diagnosis of
             dementing disorders, particularly in clinically ambiguous
             situations (e.g., suspected early dementia) and in the
             context of confounding factors, such as the presence of a
             superimposed depression or advanced age (1, 2). Within the
             medical evaluation of dementia, the neuropsychological
             evaluation provides unique information in the form of a
             “behavioral sample” that can be used to determine in an
             objective, quantifiable manner the presence of cognitive
             symptoms and their functional significance. To accomplish
             these basic aims, the neuropsychological examination employs
             rigorous, standardized psychometric tests of memory and
             cognitive function and relies heavily on the application of
             normative standards and on a fundamental understanding of
             brain function. The examination results in metric values for
             discrete cognitive and behavioral capacities, which can then
             be used by the examining clinician to arrive at diagnostic
             inferences, to make judgments of functional abilities, and
             to establish a benchmark for monitoring future change and
             responsiveness to medical treatments and
             therapies.},
   Doi = {10.3109/9780849354847-8},
   Key = {fds364315}
}


Duke University * Arts & Sciences * Faculty * Staff * Grad * Postdocs * Reload * Login