Publications of Daniel W. McShea    :chronological  combined listing:

%% Books   
@book{fds139001,
   Author = {D.W. McShea and Robert Brandon},
   Title = {Biology's First Law},
   Publisher = {University of Chicago Press},
   Year = {2010},
   Key = {fds139001}
}

@book{fds229047,
   Author = {McShea, DW and Brandon, RN},
   Title = {Biology's First Law The Tendency for Diversity and
             Complexity to Increase in Evolutionary Systems},
   Pages = {184 pages},
   Publisher = {University of Chicago Press},
   Year = {2010},
   Month = {July},
   ISBN = {9780226562278},
   Abstract = {Intended for evolutionary biologists, paleontologists, and
             other scientists studying complex systems, and written in a
             concise and engaging format that speaks to students and
             interdisciplinary practitioners alike, this book will also
             find ...},
   Key = {fds229047}
}

@book{fds229049,
   Author = {Rosenberg, A and McShea, DW},
   Title = {Philosophy of Biology: A Contemporary Introduction},
   Pages = {1-241},
   Publisher = {Routledge.},
   Year = {2007},
   Month = {January},
   ISBN = {9780415315920},
   url = {http://dx.doi.org/10.4324/9780203926994},
   Abstract = {Is life a purely physical process? What is human nature?
             Which of our traits is essential to us? In this volume,
             Daniel McShea and Alex Rosenberg - a biologist and a
             philosopher, respectively - join forces to create a new
             gateway to the philosophy of biology; making the major
             issues accessible and relevant to biologists and
             philosophers alike. Exploring concepts such as
             supervenience; the controversies about genocentrism and
             genetic determinism; and the debate about major transitions
             central to contemporary thinking about macroevolution; the
             authors lay out the broad terms in which we should assess
             the impact of biology on human capacities, social
             institutions and ethical values.},
   Doi = {10.4324/9780203926994},
   Key = {fds229049}
}

@book{fds353115,
   Author = {Brandon, R and McShea, DW},
   Title = {The Missing Two-Thirds of Evolutionary Theory},
   Pages = {75 pages},
   Publisher = {Cambridge University Press},
   Year = {2020},
   Month = {March},
   ISBN = {9781108716680},
   url = {http://dx.doi.org/10.1017/9781108591508},
   Abstract = {<jats:p>In this Element, we extend our earlier treatment of
             biology's first law. The law says that in any evolutionary
             system in which there is variation and heredity, there is a
             tendency for diversity and complexity to increase. The law
             plays the same role in biology that Newton's first law plays
             in physics, explaining what biological systems are expected
             to do when no forces act, in other words, what happens when
             nothing happens. Here we offer a deeper explanation of
             certain features of the law, develop a quantitative version
             of it, and explore its consequences for our understanding of
             diversity and complexity.</jats:p>},
   Doi = {10.1017/9781108591508},
   Key = {fds353115}
}


%% Papers Published   
@article{fds28542,
   Author = {Marcot, J.D. and D.W. McShea},
   Title = {(Abstract) Phylogenetic tests of directional bias in
             hierarchical evolution},
   Journal = {Abstracts with Programs, Geological Society of America, vol.
             36},
   Pages = {A-18},
   Year = {2004},
   Key = {fds28542}
}

@article{fds28541,
   Author = {D.W. McShea},
   Title = {(Abstract) The evolution of complexity without
             natural selection},
   Journal = {Abstracts with Programs, Geological Society of America, vol.
             36},
   Pages = {A-18},
   Year = {2004},
   Key = {fds28541}
}

@article{fds229080,
   Author = {Marino, L and Uhen, MD and McShea, D},
   Title = {(Abstract) Encephalization trends in cetacean evolution: New
             data and new analyses},
   Journal = {Brain, Behavior, and Evolution},
   Year = {2003},
   Key = {fds229080}
}

@article{fds229079,
   Author = {Novack Gottshall and PM and McShea, DW},
   Title = {(Abstract) Quantifying ecological disparity: comparative
             paleoecology of Ordovician and Recent marine
             assemblages},
   Journal = {Abstracts with Programs, Geological Society of
             America},
   Volume = {35},
   Year = {2003},
   Key = {fds229079}
}

@article{fds229088,
   Author = {McShea, DW},
   Title = {A complexity drain on cells in the evolution of
             multicellularity.},
   Journal = {Evolution; international journal of organic
             evolution},
   Volume = {56},
   Number = {3},
   Pages = {441-452},
   Year = {2002},
   Month = {March},
   ISSN = {0014-3820},
   url = {http://www.ncbi.nlm.nih.gov/pubmed/11989676},
   Abstract = {A hypothesis has been advanced recently predicting that, in
             evolution, as higher-level entities arise from associations
             of lower-level organisms, and as these entities acquire the
             ability to feed, reproduce, defend themselves, and so on,
             the lower-level organisms will tend to lose much of their
             internal complexity (McShea 2001a). In other words, in
             hierarchical transitions, there is a drain on numbers of
             part types at the lower level. One possible rationale is
             that the transfer of functional demands to the higher level
             renders many part types at the lower level useless, and thus
             their loss in evolution is favored by selection for economy.
             Here, a test is conducted at the cell level, comparing
             numbers of part types in free-living eukaryotic cells
             (protists) and the cells of metazoans and land plants.
             Differences are significant and consistent with the
             hypothesis, suggesting that tests at other hierarchical
             levels may be worthwhile.},
   Doi = {10.1111/j.0014-3820.2002.tb01357.x},
   Key = {fds229088}
}

@article{fds3923,
   Author = {D.W. McShea},
   Title = {A hypothesis about hierarchies},
   Booktitle = {Unifying Themes in Complex Systems},
   Editor = {Y. Bar-Yam},
   Year = {2000},
   Key = {fds3923}
}

@article{fds229065,
   Author = {McSHEA, DW},
   Title = {A metric for the study of evolutionary trends in the
             complexity of serial structures},
   Journal = {Biological Journal of the Linnean Society},
   Volume = {45},
   Number = {1},
   Pages = {39-55},
   Publisher = {Oxford University Press (OUP)},
   Year = {1992},
   Month = {January},
   ISSN = {0024-4066},
   url = {http://dx.doi.org/10.1111/j.1095-8312.1992.tb00630.x},
   Abstract = {Little empirical work has been done to see what sort of
             patterns of change in morphological complexity occur in
             evolution, mainly because the complexity of whole organisms
             has been so hard to define and to measure. For serial
             structures within organisms, there are fewer difficulties;
             this paper introduces a set of complexity metrics that are
             designed especially for serial structures, and then explores
             some of the properties of the new metrics. Also, a principle
             proposed in the last century by Herbert Spencer, and offered
             recently in a new form by the thermodynamic school of
             evolutionary thought, predicts that complexity should
             increase in evolution as a consequence of the accumulation
             of perturbations. Here, simulations in which perturbations
             are introduced to ideal and real series of vertebral
             measurements show how the complexity increase predicted by
             Spencer's principle would be captured by the new metrics.
             Copyright © 1992, Wiley Blackwell. All rights
             reserved},
   Doi = {10.1111/j.1095-8312.1992.tb00630.x},
   Key = {fds229065}
}

@article{fds353117,
   Author = {McShea, DW},
   Title = {A post‐modern vision of artificial life},
   Journal = {Complexity},
   Volume = {1},
   Number = {5},
   Pages = {36-38},
   Publisher = {Wiley},
   Year = {1996},
   Month = {May},
   url = {http://dx.doi.org/10.1002/cplx.6130010509},
   Doi = {10.1002/cplx.6130010509},
   Key = {fds353117}
}

@article{fds343318,
   Author = {McShea, DW and Wang, SC and Brandon, RN},
   Title = {A quantitative formulation of biology's first
             law.},
   Journal = {Evolution; international journal of organic
             evolution},
   Volume = {73},
   Number = {6},
   Pages = {1101-1115},
   Year = {2019},
   Month = {June},
   url = {http://dx.doi.org/10.1111/evo.13735},
   Abstract = {The zero-force evolutionary law (ZFEL) states that in
             evolutionary systems, in the absence of forces or
             constraints, diversity and complexity tend to increase. The
             reason is that diversity and complexity are both variance
             measures, and variances tend to increase spontaneously as
             random events accumulate. Here, we use random-walk models to
             quantify the ZFEL expectation, producing equations that give
             the probabilities of diversity or complexity increasing as a
             function of time, and that give the expected magnitude of
             the increase. We produce two sets of equations, one for the
             case in which variation occurs in discrete steps, the other
             for the case in which variation is continuous. The equations
             provide a way to decompose actual trajectories of diversity
             or complexity into two components, the portion due to the
             ZFEL and a remainder due to selection and constraint.
             Application of the equations is demonstrated using real and
             hypothetical data.},
   Doi = {10.1111/evo.13735},
   Key = {fds343318}
}

@article{fds229068,
   Author = {McShea, DW},
   Title = {A revised Darwinism},
   Journal = {Biology and Philosophy},
   Volume = {19},
   Number = {1},
   Pages = {45-53},
   Publisher = {Springer Nature},
   Year = {2004},
   Month = {January},
   url = {http://dx.doi.org/10.1023/B:BIPH.0000013260.40162.dd},
   Doi = {10.1023/B:BIPH.0000013260.40162.dd},
   Key = {fds229068}
}

@article{fds229054,
   Author = {McShea, DW},
   Title = {A universal generative tendency toward increased organismal
             complexity},
   Pages = {435-453},
   Booktitle = {Variation: A Central Concept in Biology},
   Publisher = {Elsevier},
   Editor = {B. Hallgrimsson and B. Hall},
   Year = {2005},
   Month = {December},
   url = {http://dx.doi.org/10.1016/B978-012088777-4/50020-X},
   Abstract = {Characterizing internal variance as complexity needs
             justification, because in colloquial usage, complexity
             connotes so much more. A complex organism is ordinarily
             understood to be not just more internally varied, or more
             differentiated, but more capable as well. The human brain is
             thought to be complex not simply because it has many cell
             types, but because of its impressive functional
             capabilities, because of what it can do. Thus, as
             conventionally understood, complexity depends on both
             structure and function. However, in biology, a narrower view
             has been adopted, herein complexity refers to number of part
             types, or degree of differentiation among parts. Complexity
             has other aspects besides number of part types. For example,
             there is complexity of spatial arrangement of parts, a kind
             of second-order complexity (where number of part types is
             first order), and number of types of connections among
             parts. The chapter introduces three simple models to
             illustrate the internal-variance principle and also to
             reveal its robustness. In each successive model, the
             variations introduced are more finely tuned in such a way as
             to negate or overcome the internal-variance principle. ©
             2005 Elsevier Inc. All rights reserved.},
   Doi = {10.1016/B978-012088777-4/50020-X},
   Key = {fds229054}
}

@article{fds356869,
   Author = {Babcock, G and McShea, DW},
   Title = {An externalist teleology},
   Journal = {Synthese},
   Volume = {199},
   Number = {3-4},
   Pages = {8755-8780},
   Year = {2021},
   Month = {December},
   url = {http://dx.doi.org/10.1007/s11229-021-03181-w},
   Abstract = {Teleology has a complicated history in the biological
             sciences. Some have argued that Darwin’s theory has
             allowed biology to purge itself of teleological
             explanations. Others have been content to retain teleology
             and to treat it as metaphorical, or have sought to replace
             it with less problematic notions like teleonomy. And still
             others have tried to naturalize it in a way that distances
             it from the vitalism of the nineteenth century, focusing on
             the role that function plays in teleological explanation. No
             consensus has seemed possible in this debate. This paper
             takes a different approach. It argues that teleology is a
             perfectly acceptable scientific notion, but that the debate
             took an unfortunate misstep some 2300 years ago, one that
             has confused things ever since. The misstep comes in the
             beginning of Aristotle’s Physics when a distinction is
             made between two types of teleological explanation. One type
             pertains to artifacts while the other pertains to entities
             in nature. For Aristotle, artifacts are guided by something
             external to themselves, human intentions, while natural
             entities are guided by an internal nature. We aim to show
             that there is, in fact, only one type of legitimate
             teleological explanation, what Aristotle would have
             considered a variant of an artifact model, where entities
             are guided by external fields. We begin with an analysis of
             the differences between the two types of explanation. We
             then examine some evidence in Aristotle’s biological works
             suggesting that in his account of the natural-artifactual
             distinction, he encountered difficulties in trying to
             provide teleological accounts of spontaneous generation. And
             we show that it is possible to resolve these difficulties
             with a more robust version of an artifact model of
             teleology, in other words, with an externalist teleology.
             This is McShea’s model, in which goal-directed entities
             are guided by a nested series a of upper-level fields. To
             explain teleological behavior, this account invokes only
             external physical forces rather than mysterious internal
             natures. We then consider how field theory differs from
             other efforts to naturalize teleology in biology. And
             finally, we show how the account enables us to grapple with
             certain difficult cases—genes and intentions—where, even
             in biology today, the temptation to posit internal natures
             remains strong.},
   Doi = {10.1007/s11229-021-03181-w},
   Key = {fds356869}
}

@article{fds363030,
   Author = {De Castro and C and McShea, DW},
   Title = {Applying the Prigogine view of dissipative systems to the
             major transitions in evolution},
   Journal = {Paleobiology},
   Volume = {48},
   Number = {4},
   Pages = {711-728},
   Year = {2022},
   Month = {November},
   url = {http://dx.doi.org/10.1017/pab.2022.7},
   Abstract = {Ilya Prigogine's trinomial concept is, he argued, applicable
             to many complex dissipative systems, from physics to biology
             and even to social systems. For Prigogine, this trinomial -
             functions, structure, fluctuations - was intended to capture
             the feedback-rich relations between upper and lower levels
             in these systems. The main novelty of his vision was his
             view of causation, in which the causal arrow runs downward
             from dissipative structures to their components or
             functions. Following this insight, some physicists and
             biophysicists are beginning to apply terms formerly used
             mainly in biology, such as evolution, adaptation, learning,
             and life-like behavior, to physical and chemical
             nonequilibrium systems. Here, instead, we apply Prigogine's
             view to biology, in particular to evolution, and especially
             the major transitions in evolution (MTE), arguing that at
             least the hierarchical transitions - the transitions in
             individuality - follow a trajectory anticipated by the
             trinomial. In this trajectory, formerly free-living
             organisms are transformed into functions within a larger
             organic structure. The Prigogine view also predicts that,
             consistent with available data, the increase in number of
             hierarchical levels in organisms should accelerate over
             time. Finally, it predicts that, on geological timescales,
             ecosystems and Gaia in particular will tend to de-Darwinize
             or machinify their component organisms.},
   Doi = {10.1017/pab.2022.7},
   Key = {fds363030}
}

@article{fds229056,
   Author = {Mcshea, DW},
   Title = {Arguments, tests, and the Burgess Shale � a commentary on
             the debate},
   Journal = {Paleobiology},
   Volume = {19},
   Number = {4},
   Pages = {399-402},
   Publisher = {Cambridge University Press (CUP)},
   Year = {1993},
   Month = {January},
   ISSN = {0094-8373},
   url = {http://gateway.webofknowledge.com/gateway/Gateway.cgi?GWVersion=2&SrcApp=PARTNER_APP&SrcAuth=LinksAMR&KeyUT=WOS:A1993ML91300001&DestLinkType=FullRecord&DestApp=ALL_WOS&UsrCustomerID=47d3190e77e5a3a53558812f597b0b92},
   Doi = {10.1017/S0094837300014044},
   Key = {fds229056}
}

@article{fds327781,
   Author = {McShea, DW},
   Title = {Bernd Rosslenbroich: On the origin of autonomy: a new look
             at the major transitions in evolution},
   Journal = {Biology & Philosophy},
   Volume = {30},
   Number = {3},
   Pages = {439-446},
   Publisher = {Springer Science and Business Media LLC},
   Year = {2015},
   Month = {May},
   url = {http://dx.doi.org/10.1007/s10539-015-9474-2},
   Doi = {10.1007/s10539-015-9474-2},
   Key = {fds327781}
}

@article{fds366842,
   Author = {Rosenberg, A and McShea, DW},
   Title = {Biological laws and theories},
   Pages = {32-64},
   Booktitle = {PHILOSOPHY OF BIOLOGY: A CONTEMPORARY INTRODUCTION},
   Year = {2008},
   Key = {fds366842}
}

@article{fds3820,
   Author = {R.J. McShea and D.W. McShea},
   Title = {Biology and value theory},
   Booktitle = {Biology and the Foundation of Ethics},
   Publisher = {Cambridge University Press},
   Editor = {J. Maienschein and M. Ruse},
   Year = {1999},
   Key = {fds3820}
}

@article{fds366837,
   Author = {Rosenberg, A and McShea, DW},
   Title = {Biology, human behavior, social science, and moral
             philosophy},
   Pages = {187-225},
   Booktitle = {PHILOSOPHY OF BIOLOGY: A CONTEMPORARY INTRODUCTION},
   Year = {2008},
   Key = {fds366837}
}

@article{fds322304,
   Author = {Smith, FA and Payne, JL and Heim, NA and Balk, MA and Finnegan, S and Kowalewski, M and Lyons, SK and McClain, CR and McShea, DW and Novack-Gottshall, PM and Anich, PS and Wang, SC},
   Title = {Body Size Evolution Across the Geozoic},
   Journal = {Annual Review of Earth and Planetary Sciences},
   Volume = {44},
   Number = {1},
   Pages = {523-553},
   Publisher = {ANNUAL REVIEWS},
   Year = {2016},
   Month = {June},
   url = {http://dx.doi.org/10.1146/annurev-earth-060115-012147},
   Abstract = {The Geozoic encompasses the 3.6 Ga interval in Earth history
             when life has existed. Over this time, life has diversified
             from exclusively tiny, single-celled organisms to include
             large, complex multicellular forms. Just how and why this
             diversification occurred has been a major area of interest
             for paleontologists and evolutionary biologists for
             centuries. Here, we compile data on organism size throughout
             the Geozoic fossil record for the three domains of life. We
             describe canonical trends in the evolution of body size,
             synthesize current understanding of the patterns and causal
             mechanisms at various hierarchical scales, and discuss the
             biological and geological consequences of variation in
             organismal size.},
   Doi = {10.1146/annurev-earth-060115-012147},
   Key = {fds322304}
}

@article{fds229053,
   Author = {McShea, DW},
   Title = {Comments on "evolutionary complexity, " H. Morowitz,
             complexity 3(6): pp 12-14.},
   Journal = {Complex.},
   Volume = {4},
   Number = {2},
   Pages = {11-12},
   Publisher = {WILEY},
   Year = {1998},
   url = {http://dx.doi.org/10.1002/(SICI)1099-0526(199811/12)4:2<11::AID-CPLX2>3.0.CO;2},
   Doi = {10.1002/(SICI)1099-0526(199811/12)4:2<11::AID-CPLX2>3.0.CO;2},
   Key = {fds229053}
}

@article{fds229063,
   Author = {McShea, DW and Raup, DM},
   Title = {Completeness of the geological record.},
   Journal = {The Journal of geology},
   Volume = {94},
   Pages = {569-574},
   Year = {1986},
   Month = {January},
   ISSN = {0022-1376},
   url = {http://www.ncbi.nlm.nih.gov/pubmed/11542057},
   Abstract = {The completeness of a sedimentary section of known timespan
             may be assessed qualitatively by comparing its thickness
             with the average accumulation for that timespan. Average
             accumulations may be estimated from sediment volume and
             continental area data. Quantitative completeness estimation
             methods based on data compiled from the geological
             literature have been proposed, but we argue that the
             literature data are significantly biased and cannot support
             such methods. Interestingly, however, a comparison of the
             literature data and accumulation averages computed from
             sediment volume data suggests that the thickest known
             sections may be extremely complete.},
   Doi = {10.1086/629058},
   Key = {fds229063}
}

@article{fds229064,
   Author = {McShea, DW},
   Title = {Complexity and evolution: What everybody
             knows},
   Journal = {Biology and Philosophy},
   Volume = {6},
   Number = {3},
   Pages = {303-324},
   Publisher = {Springer Nature},
   Year = {1991},
   Month = {July},
   ISSN = {0169-3867},
   url = {http://dx.doi.org/10.1007/BF00132234},
   Abstract = {The consensus among evolutionists seems to be (and has been
             for at least a century) that the morphological complexity of
             organisms increases in evolution, although almost no
             empirical evidence for such a trend exists. Most studies of
             complexity have been theoretical, and the few empirical
             studies have not, with the exception of certain recent ones,
             been especially rigorous; reviews are presented of both the
             theoretical and empirical literature. The paucity of
             evidence raises the question of what sustains the consensus,
             and a number of suggestions are offered, including the
             possibility that certain cultural and/or perceptual biases
             are at work. In addition, a shift in emphasis from
             theoretical to empirical inquiry is recommended for the
             study of complexity, and guidelines for future empirical
             studies are proposed. © 1991 Kluwer Academic
             Publishers.},
   Doi = {10.1007/BF00132234},
   Key = {fds229064}
}

@article{fds355923,
   Author = {McSHEA, DW},
   Title = {COMPLEXITY AND HOMOPLASY},
   Pages = {207-225},
   Booktitle = {Homoplasy},
   Publisher = {Elsevier},
   Year = {1996},
   ISBN = {9780126180305},
   url = {http://dx.doi.org/10.1016/b978-012618030-5/50010-1},
   Doi = {10.1016/b978-012618030-5/50010-1},
   Key = {fds355923}
}

@article{fds353116,
   Author = {McShea, DW and Hordijk, W},
   Title = {Complexity by Subtraction},
   Journal = {Evolutionary Biology},
   Pages = {1-17},
   Year = {2013},
   Key = {fds353116}
}

@article{fds366838,
   Author = {Rosenberg, A and McShea, DW},
   Title = {Complexity, directionality, and progress in
             evolution},
   Pages = {127-156},
   Booktitle = {PHILOSOPHY OF BIOLOGY: A CONTEMPORARY INTRODUCTION},
   Year = {2008},
   Key = {fds366838}
}

@article{fds366843,
   Author = {Rosenberg, A and McShea, DW},
   Title = {Darwin makes a science},
   Pages = {12-31},
   Booktitle = {PHILOSOPHY OF BIOLOGY: A CONTEMPORARY INTRODUCTION},
   Year = {2008},
   Key = {fds366843}
}

@article{fds229089,
   Author = {Ciampaglio, CN and Kemp, M and McShea, DW},
   Title = {Detecting changes in morphospace occupation patterns in the
             fossil record: Characterization and analysis of measures of
             disparity},
   Journal = {Paleobiology},
   Volume = {27},
   Number = {4},
   Pages = {695-715},
   Publisher = {Cambridge University Press (CUP)},
   Year = {2001},
   Month = {Fall},
   url = {http://dx.doi.org/10.1666/0094-8373(2001)027<0695:DCIMOP>2.0.CO;2},
   Abstract = {Recently, there has been much interest in detecting and
             measuring patterns of change in disparity. Although most
             studies have used one or two measures of disparity to
             quantify and characterize the occupation of morphospace,
             multiple measures may be necessary to fully detect changes
             in patterns of morphospace occupation. Also, the ability to
             detect morphological trends and occupation patterns within
             morphospace depends on using the appropriate measure(s) of
             disparity. In this study, seven measures were used to
             determine and characterize sensitivity to sample size of the
             data, number of morphological characters, percentage of
             missing data, and changes in morphospace occupation pattern.
             These consist of five distance measures - sum of univariate
             variances, total range, mean distance, principal coordinate
             analysis volume, average pairwise dissimilarity - and two
             non-distance measures - participation ratio and number of
             unique pairwise character combinations. Evaluation of each
             measure with respect to sensitivity to sample size, number
             of morphological characters, and percentage of missing data
             was accomplished by using both simulated and Ordovician
             crinoid data. For simulated data, each measure of disparity
             was evaluated for its response to changes of morphospace
             occupation pattern, and with respect to simulated random and
             nonrandom extinction events. Changes in disparity were also
             measured within the Crinoidea across the Permian extinction
             event. Although all measures vary in sensitivity with
             respect to species sample size, number of morphological
             characters, and percentage of missing data, the non-distance
             measures overall produce the lowest estimates of variance
             (in bootstrap analyses). The non-distance measures appear to
             be relatively insensitive to changes in morphospace
             occupation pattern. All measures, except average pairwise
             dissimilarity, detect changes in occupation pattern in
             simulated nonrandom extinction events, but all measures,
             except number of unique pairwise character combinations and
             principal coordinate analysis volume, are relatively
             insensitive to changes in pattern in simulated random
             extinction events. The distance measures report similar
             changes in disparity over the Permian extinction event,
             whereas the non-distance measures differ. This study
             suggests that each measures of disparity is designed for
             different purposes, and that by using a combination of
             techniques a clearer picture of disparity should
             emerge.},
   Doi = {10.1666/0094-8373(2001)027<0695:DCIMOP>2.0.CO;2},
   Key = {fds229089}
}

@article{fds229072,
   Author = {Fleming, L and McShea, DW},
   Title = {Drosophila mutants suggest a strong drive toward complexity
             in evolution},
   Journal = {Evolution and Development},
   Volume = {15},
   Number = {1},
   Pages = {53-62},
   Year = {2012},
   url = {http://www.ncbi.nlm.nih.gov/pubmed/23331917},
   Abstract = {The view that complexity increases in evolution is
             uncontroversial, yet little is known about the possible
             causes of such a trend. One hypothesis, the Zero Force
             Evolutionary Law (ZFEL), predicts a strong drive toward
             complexity, although such a tendency can be overwhelmed by
             selection and constraints. In the absence of strong
             opposition, heritable variation accumulates and complexity
             increases. In order to investigate this claim, we evaluate
             the gross morphological complexity of laboratory mutants in
             Drosophila melanogaster, which represent organisms that
             arise in a context where selective forces are greatly
             reduced. Complexity was measured with respect to part types,
             shape, and color over two independent focal levels. Compared
             to the wild type, we find that D. melanogaster mutants are
             significantly more complex. When the parts of mutants are
             categorized by degree of constraint, we find that weakly
             constrained parts are significantly more complex than more
             constrained parts. These results support the ZFEL
             hypothesis. They also represent a first step in establishing
             the domain of application of the ZFEL and show one way in
             which a larger empirical investigation of the principle
             might proceed.},
   Doi = {10.1111/ede.12014},
   Key = {fds229072}
}

@article{fds3926,
   Author = {D.W. McShea},
   Title = {Dynamics of large-scale trends},
   Booktitle = {Biodiversity Dynamics},
   Editor = {M.L. McKinney and J.A. Drake},
   Year = {1998},
   Key = {fds3926}
}

@article{fds355920,
   Author = {McShea, DW},
   Title = {Evolution of Complexity},
   Pages = {169-179},
   Booktitle = {Evolutionary Developmental Biology},
   Publisher = {Springer International Publishing},
   Year = {2021},
   ISBN = {9783319329772},
   url = {http://dx.doi.org/10.1007/978-3-319-32979-6_123},
   Doi = {10.1007/978-3-319-32979-6_123},
   Key = {fds355920}
}

@article{fds355921,
   Author = {McShea, DW},
   Title = {Evolution of Complexity},
   Pages = {1-11},
   Booktitle = {Evolutionary Developmental Biology},
   Publisher = {Springer International Publishing},
   Year = {2017},
   ISBN = {9783319330389},
   url = {http://dx.doi.org/10.1007/978-3-319-33038-9_123-1},
   Doi = {10.1007/978-3-319-33038-9_123-1},
   Key = {fds355921}
}

@article{fds229060,
   Author = {MCSHEA, DW},
   Title = {EVOLUTIONARY CHANGE IN THE MORPHOLOGICAL COMPLEXITY OF THE
             MAMMALIAN VERTEBRAL COLUMN},
   Journal = {EVOLUTION},
   Volume = {47},
   Number = {3},
   Pages = {730-740},
   Publisher = {JSTOR},
   Year = {1993},
   ISSN = {0014-3820},
   url = {http://gateway.webofknowledge.com/gateway/Gateway.cgi?GWVersion=2&SrcApp=PARTNER_APP&SrcAuth=LinksAMR&KeyUT=WOS:A1993MA41600002&DestLinkType=FullRecord&DestApp=ALL_WOS&UsrCustomerID=47d3190e77e5a3a53558812f597b0b92},
   Abstract = {The notion that morphological complexity increases in
             evolution is widely accepted in biology and paleontology.
             Several possible explanations have been offered for this
             trend, among them the suggestion that it has an active
             forcing mechanism, such as natural selection or the second
             law of thermodynamics. No such mechanism has yet been
             empirically demonstrated, but testing is possible: if a
             forcing mechanism has operated, the expectation is that
             complexity would have increased in evolutionary lineages
             more frequently than it decreased. However, a quantitative
             analysis of changes in the complexity of the vertebral
             column in a random sample of mammalian lineages reveals a
             nearly equal number of increases and decreases. This finding
             raises the possibility that no forcing mechanism exists, or
             at least that it may not be as powerful or pervasive as has
             been assumed. The finding also highlights the need for more
             empirical tests.},
   Doi = {10.2307/2410179},
   Key = {fds229060}
}

@article{fds229045,
   Author = {McShea, DW},
   Title = {Evolutionary progress},
   Pages = {550-557},
   Booktitle = {Evolution: The First Four Billion Years},
   Publisher = {Harvard University Press},
   Editor = {Ruse, M and Travis, J},
   Year = {2011},
   Key = {fds229045}
}

@article{fds370659,
   Author = {McShea, DW},
   Title = {Evolutionary Success: Standards of Value},
   Pages = {17-39},
   Booktitle = {Human Success: Evolutionary Origins and Ethical
             Implications},
   Year = {2023},
   Month = {January},
   ISBN = {9780190096168},
   url = {http://dx.doi.org/10.1093/oso/9780190096168.003.0002},
   Doi = {10.1093/oso/9780190096168.003.0002},
   Key = {fds370659}
}

@article{fds322305,
   Author = {McShea, DW},
   Title = {Evolutionary Trends},
   Pages = {206-211},
   Booktitle = {Palaeobiology II},
   Year = {2007},
   Month = {December},
   ISBN = {9780632051496},
   url = {http://dx.doi.org/10.1002/9780470999295ch.44},
   Doi = {10.1002/9780470999295ch.44},
   Key = {fds322305}
}

@article{fds372477,
   Author = {Mcshea, DW},
   Title = {Evolutionary Trends},
   Series = {pp. 206-210},
   Pages = {206-211},
   Booktitle = {Palaeobiology II},
   Publisher = {Wiley},
   Editor = {DEG Briggs and PR Crowther},
   Year = {2001},
   Month = {January},
   ISBN = {9780632051496},
   url = {http://dx.doi.org/10.1002/9780470999295.ch44},
   Doi = {10.1002/9780470999295.ch44},
   Key = {fds372477}
}

@article{fds370848,
   Author = {McShea, DW},
   Title = {Evolutionary trends and goal directedness.},
   Journal = {Synthese},
   Volume = {201},
   Number = {5},
   Pages = {178},
   Year = {2023},
   Month = {January},
   url = {http://dx.doi.org/10.1007/s11229-023-04164-9},
   Abstract = {The conventional wisdom declares that evolution is not goal
             directed, that teleological considerations play no part in
             our understanding of evolutionary trends. Here I argue that,
             to the contrary, under a current view of teleology, field
             theory, most evolutionary trends would have to be considered
             goal directed to some degree. Further, this view is
             consistent with a modern scientific outlook, and more
             particularly with evolutionary theory today. Field theory
             argues that goal directedness is produced by higher-level
             fields that direct entities contained within them to behave
             persistently and plastically, that is, returning them to a
             goal-directed trajectory following perturbations
             (persistence) and directing them to a goal-directed
             trajectory from a large range of alternative starting points
             (plasticity). The behavior of a bacterium climbing a
             chemical food gradient is persistent and plastic, with
             guidance provided by the external "food field," the chemical
             gradient. Likewise, an evolutionary trend that is produced
             by natural selection is a lineage behaving persistently and
             plastically under the direction of its local ecology, an
             "ecological field." Trends directed by selection-generated
             boundaries, thermodynamic gradients, and certain internal
             constraints, would also count as goal directed. In other
             words, most of the causes of evolutionary trends that have
             been proposed imply goal directedness. However, under field
             theory, not all trends are goal directed. Examples are
             discussed. Importantly, nothing in this view suggests that
             evolution is guided by intentionality, at least none at the
             level of animal intentionality. Finally, possible
             implications for our thinking about evolutionary
             directionality in the history of life are
             discussed.},
   Doi = {10.1007/s11229-023-04164-9},
   Key = {fds370848}
}

@article{fds322306,
   Author = {McShea, DW},
   Title = {Evolutionary Trends and the Salience Bias (with Apologies to
             Oil Tankers, Karl Marx, and Others)},
   Journal = {Technical Communication Quarterly},
   Volume = {3},
   Number = {1},
   Pages = {21-38},
   Publisher = {Informa UK Limited},
   Year = {1994},
   Month = {January},
   url = {http://dx.doi.org/10.1080/10572259409364556},
   Abstract = {Salient examples may bias human judgments about the
             probability or frequency of events, an effect known as the
             “availability heuristic” or the “salience bias.”
             Scientific work has not been immune to this bias; in
             particular, the existence of certain large-scale trends in
             evolution, such as those in size, complexity, and fitness,
             is widely accepted among professionals within evolutionary
             biology and paleontology, as well as among people outside
             these fields, even though these trends are poorly
             documented. Often, what documentation exists consists mainly
             of long lists of cases exemplifying the trend, or detailed
             descriptions of a small number of salient cases. Here, it is
             argued that although these lists and salient cases are not
             good evidence that a trend is pervasive, they may convince
             both the trend researcher and his or her audience. The
             possibility is raised that the bias may be pervasive in
             science and everyday thought, and a strategy for avoiding
             it, namely the use of random samples, is offered. © 1994,
             Taylor & Francis Group, LLC. All rights reserved.},
   Doi = {10.1080/10572259409364556},
   Key = {fds322306}
}

@article{fds3803,
   Author = {D.W. McShea},
   Title = {Feelings as the proximate cause of behavior},
   Booktitle = {Where Psychology Meets Biology: Philosophical
             Essays},
   Publisher = {Cambridge University Press},
   Editor = {V.G. Hardcastle},
   Year = {1999},
   Key = {fds3803}
}

@article{fds369050,
   Author = {McShea, DW},
   Title = {Four reasons for scepticism about a human major transition
             in social individuality.},
   Journal = {Philosophical transactions of the Royal Society of London.
             Series B, Biological sciences},
   Volume = {378},
   Number = {1872},
   Pages = {20210403},
   Year = {2023},
   Month = {March},
   url = {http://dx.doi.org/10.1098/rstb.2021.0403},
   Abstract = {The 'major transitions in evolution' are mainly about the
             rise of hierarchy, new individuals arising at ever higher
             levels of nestedness, in particular the eukaryotic cell
             arising from prokaryotes, multicellular individuals from
             solitary protists and individuated societies from
             multicellular individuals. Some lists include human
             societies as a major transition, but based on a comparison
             with the non-human transitions, there are reasons for
             scepticism. (i) The foundation of the major transitions is
             hierarchy, but the cross-cutting interactions in human
             societies undermine hierarchical structure. (ii) Natural
             selection operates in three modes-stability, growth and
             reproductive success-and only the third produces the complex
             adaptations seen in fully individuated higher levels. But
             human societies probably evolve mainly in the stability and
             growth modes. (iii) Highly individuated entities are marked
             by division of labour and commitment to morphological
             differentiation, but in humans differentiation is mostly
             behavioural and mostly reversible. (iv) As higher-level
             individuals arise, selection drains complexity, drains
             parts, from lower-level individuals. But there is little
             evidence of a drain in humans. In sum, a comparison with the
             other transitions gives reasons to doubt that human social
             individuation has proceeded very far, or if it has, to doubt
             that it is a transition of the same sort. This article is
             part of the theme issue 'Human socio-cultural evolution in
             light of evolutionary transitions'.},
   Doi = {10.1098/rstb.2021.0403},
   Key = {fds369050}
}

@article{fds229073,
   Author = {Brandon, RN and McShea, DW},
   Title = {Four solutions for four puzzles},
   Journal = {Biology and Philosophy},
   Volume = {27},
   Number = {5},
   Pages = {737-744},
   Publisher = {Springer Nature},
   Year = {2012},
   Month = {September},
   ISSN = {0169-3867},
   url = {http://dx.doi.org/10.1007/s10539-012-9330-6},
   Abstract = {Barrett et al. (Biol Philos, 2012) present four puzzles for
             the ZFEL-view of evolution that we present in our 2010 book,
             Biology's First Law: The Tendency for Diversity and
             Complexity to Increase in Evolutionary Systems. Our intent
             in writing this book was to present a radically different
             way to think about evolution. To the extent that it really
             is radical, it will be easy to misunderstand. We think
             Barrett et al. have misunderstood several crucial points and
             so we welcome the opportunity to clarify. © 2012 Springer
             Science+Business Media B.V.},
   Doi = {10.1007/s10539-012-9330-6},
   Key = {fds229073}
}

@article{fds322303,
   Author = {McShea, DW},
   Title = {Freedom and purpose in biology.},
   Journal = {Studies in history and philosophy of biological and
             biomedical sciences},
   Volume = {58},
   Pages = {64-72},
   Year = {2016},
   Month = {August},
   url = {http://dx.doi.org/10.1016/j.shpsc.2015.12.002},
   Abstract = {All seemingly teleological systems share a common
             hierarchical structure. They consist of a small entity
             moving or changing within a larger field that directs it
             from above (what I call "upper direction"). This is true for
             organisms seeking some external resource, for the organized
             behavior of cells and other parts in organismal development,
             and for lineages evolving by natural selection. In all
             cases, the lower-level entity is partly "free," tending to
             wander under the influence of purely local forces, and
             partly directed by a larger enveloping field. The persistent
             and plastic behavior that characterizes goal-directedness
             arises, I argue, at intermediate levels of freedom and upper
             direction, when the two are in a delicate balance. I
             tentatively extend the argument to human teleology (wants,
             purposes).},
   Doi = {10.1016/j.shpsc.2015.12.002},
   Key = {fds322303}
}

@article{fds229083,
   Author = {McShea, DW},
   Title = {Functional complexity in organisms: Parts as
             proxies},
   Journal = {Biology and Philosophy},
   Volume = {15},
   Number = {5},
   Pages = {641-668},
   Publisher = {Springer Nature},
   Year = {2000},
   Month = {December},
   ISSN = {0169-3867},
   url = {http://dx.doi.org/10.1023/A:1006695908715},
   Abstract = {The functional complexity, or the number of functions, of
             organisms has figured prominently in certain theoretical and
             empirical work in evolutionary biology. Large-scale trends
             in functional complexity and correlations between functional
             complexity and other variables, such as size, have been
             proposed. However, the notion of number of functions has
             also been operationally intractable, in that no method has
             been developed for counting functions in an organism in a
             systematic and reliable way. Thus, studies have had to rely
             on the largely unsupported assumption that number of
             functions can be measured indirectly, by using number of
             morphological, physiological, and behavioral partsas a
             proxy. Here, a model is developed that supports this
             assumption. Specifically, the model predicts that few parts
             will have many functions overlapping in them, and therefore
             the variance in number of functions per part will be low. If
             so, then number of parts is expected to be well correlated
             with number of functions, and we can use part counts as
             proxies for function counts in comparative studies of
             organisms, even when part counts are low. Also discussed
             briefly is a strategy for identifying certain kinds of parts
             in organisms in a systematic way.},
   Doi = {10.1023/A:1006695908715},
   Key = {fds229083}
}

@article{fds366841,
   Author = {Rosenberg, A and McShea, DW},
   Title = {Further problems of Darwinism Constraint, drift,
             function},
   Pages = {65-95},
   Booktitle = {PHILOSOPHY OF BIOLOGY: A CONTEMPORARY INTRODUCTION},
   Year = {2008},
   Key = {fds366841}
}

@article{fds366839,
   Author = {Rosenberg, A and McShea, DW},
   Title = {Genes, groups, teleosemantics, and the major
             transitions},
   Pages = {157-186},
   Booktitle = {PHILOSOPHY OF BIOLOGY: A CONTEMPORARY INTRODUCTION},
   Year = {2008},
   Key = {fds366839}
}

@article{fds355924,
   Author = {McShea, DW},
   Title = {Gene‐talk talk about sociobiology},
   Journal = {Social Epistemology},
   Volume = {6},
   Number = {2},
   Pages = {183-192},
   Publisher = {Informa UK Limited},
   Year = {1992},
   Month = {April},
   url = {http://dx.doi.org/10.1080/02691729208578653},
   Doi = {10.1080/02691729208578653},
   Key = {fds355924}
}

@article{fds374935,
   Author = {Babcock, G and McShea, DW},
   Title = {Goal Directedness and the Field Concept},
   Journal = {Philosophy of Science},
   Pages = {1-10},
   Publisher = {Cambridge University Press (CUP)},
   Year = {2023},
   Month = {October},
   url = {http://dx.doi.org/10.1017/psa.2023.121},
   Abstract = {<jats:title>Abstract</jats:title> <jats:p>A long-standing
             problem in understanding goal-directed systems has been the
             insufficiency of mechanistic explanations to make sense of
             them. This article offers a solution to this problem. It
             begins by observing the limitations of mechanistic
             decompositions when it comes to understanding physical
             fields. We argue that introducing the field concept, as it
             has been developed in <jats:italic>field
             theory</jats:italic>, alongside mechanisms is able to
             provide an account of goal directedness in the
             sciences.</jats:p>},
   Doi = {10.1017/psa.2023.121},
   Key = {fds374935}
}

@article{fds327277,
   Author = {Heim, NA and Payne, JL and Finnegan, S and Knope, ML and Kowalewski, M and Lyons, SK and McShea, DW and Novack-Gottshall, PM and Smith, FA and Wang, SC},
   Title = {Hierarchical complexity and the size limits of
             life.},
   Journal = {Proceedings. Biological sciences},
   Volume = {284},
   Number = {1857},
   Pages = {20171039},
   Year = {2017},
   Month = {June},
   url = {http://dx.doi.org/10.1098/rspb.2017.1039},
   Abstract = {Over the past 3.8 billion years, the maximum size of life
             has increased by approximately 18 orders of magnitude. Much
             of this increase is associated with two major evolutionary
             innovations: the evolution of eukaryotes from prokaryotic
             cells approximately 1.9 billion years ago (Ga), and
             multicellular life diversifying from unicellular ancestors
             approximately 0.6 Ga. However, the quantitative relationship
             between organismal size and structural complexity remains
             poorly documented. We assessed this relationship using a
             comprehensive dataset that includes organismal size and
             level of biological complexity for 11 172 extant genera. We
             find that the distributions of sizes within complexity
             levels are unimodal, whereas the aggregate distribution is
             multimodal. Moreover, both the mean size and the range of
             size occupied increases with each additional level of
             complexity. Increases in size range are non-symmetric: the
             maximum organismal size increases more than the minimum. The
             majority of the observed increase in organismal size over
             the history of life on the Earth is accounted for by two
             discrete jumps in complexity rather than evolutionary trends
             within levels of complexity. Our results provide
             quantitative support for an evolutionary expansion away from
             a minimal size constraint and suggest a fundamental
             rescaling of the constraints on minimal and maximal size as
             biological complexity increases.},
   Doi = {10.1098/rspb.2017.1039},
   Key = {fds327277}
}

@article{fds229092,
   Author = {McShea, DW and Venit, EP and Simon, VB},
   Title = {Hierarchical complexity of organisms: dynamics of a
             well-known trend},
   Journal = {Abstracts with Programs, Geological Society of
             America},
   Volume = {31},
   Pages = {A-171},
   Year = {1999},
   Key = {fds229092}
}

@article{fds229062,
   Author = {Golob, RS and McShea, DW},
   Title = {IMPLICATIONS OF THE IXTOC 1 BLOW-OUT AND OIL
             SPILL.},
   Pages = {743-759},
   Year = {1981},
   Month = {December},
   Key = {fds229062}
}

@article{fds229077,
   Author = {Marcot, JD and McShea, DW},
   Title = {Increasing hierarchical complexity throughout the history of
             life: Phylogenetic tests of trend mechanisms},
   Journal = {Paleobiology},
   Volume = {33},
   Number = {2},
   Pages = {182-200},
   Publisher = {Cambridge University Press (CUP)},
   Year = {2007},
   Month = {March},
   ISSN = {0094-8373},
   url = {http://dx.doi.org/10.1666/06028.1},
   Abstract = {The history of life is punctuated by a number of major
             transitions in hierarchy, defined here as the degree of
             nestedness of lower-level individuals within higher-level
             ones: the combination of single-celled prokaryotic cells to
             form the first eukaryotic cell, the aggregation of single
             eukaryotic cells to form complex multicellular organisms,
             and finally, the association of multicellular organisms to
             form complex colonial individuals. These transitions
             together constitute one of the most salient and certain
             trends in the history of life, in particular, a trend in
             maximum hierarchical structure, which can be understood as a
             trend in complexity. This trend could be produced by a
             biased mechanism, in which increases in hierarchy are more
             likely than decreases, or by an unbiased one, in which
             increases and decreases are about equally likely. At stake
             is whether or not natural selection or some other force acts
             powerfully over the history of life to drive complexity
             upward. Too few major transitions are known to permit
             rigorous statistical discrimination of trend mechanisms
             based on these transitions alone. However, the mechanism can
             be investigated by using "minor transitions" in hierarchy,
             or, in other words, changes in the degree of individuation
             of the upper level. This study tests the null hypothesis
             that the probability (or rate) of increase and decrease in
             individuation are equal in a phylogenetic context. We found
             published phylogenetic trees for clades spanning minor
             transitions across the tree of life and identified changes
             in character states associated with those minor transitions.
             We then used both parsimony- and maximum-likelihood-based
             methods to test for asymmetrical rates of character
             evolution. Most analyses failed to reject equal rates of
             hierarchical increase and decrease. In fact, a bias toward
             decreasing complexity was observed for several clades. These
             results suggest that no strong tendency exists for
             hierarchical complexity to increase. © 2007 The
             Paleontological Society. All rights reserved.},
   Doi = {10.1666/06028.1},
   Key = {fds229077}
}

@article{fds229084,
   Author = {Anderson, C and McShea, DW},
   Title = {Individual versus social complexity, with particular
             reference to ant colonies.},
   Journal = {Biological reviews of the Cambridge Philosophical
             Society},
   Volume = {76},
   Number = {2},
   Pages = {211-237},
   Year = {2001},
   Month = {May},
   ISSN = {1464-7931},
   url = {http://www.ncbi.nlm.nih.gov/pubmed/11396847},
   Abstract = {Insect societies colonies of ants, bees, wasps and
             termites--vary enormously in their social complexity. Social
             complexity is a broadly used term that encompasses many
             individual and colony-level traits and characteristics such
             as colony size, polymorphism and foraging strategy. A number
             of earlier studies have considered the relationships among
             various correlates of social complexity in insect societies;
             in this review, we build upon those studies by proposing
             additional correlates and show how all correlates can be
             integrated in a common explanatory framework. The various
             correlates are divided among four broad categories
             (sections). Under 'polyphenism' we consider the differences
             among individuals, in particular focusing upon 'caste' and
             specialization of individuals. This is followed by a section
             on 'totipotency' in which we consider the autonomy and
             subjugation of individuals. Under this heading we consider
             various aspects such as intracolony conflict, worker
             reproductive potential and physiological or morphological
             restrictions which limit individuals' capacities to perform
             a range of tasks or functions. A section entitled
             'organization of work' considers a variety of aspects, e.g.
             the ability to tackle group, team or partitioned tasks,
             foraging strategies and colony reliability and efficiency. A
             final section, 'communication and functional integration',
             considers how individual activity is coordinated to produce
             an integrated and adaptive colony. Within each section we
             use illustrative examples drawn from the social insect
             literature (mostly from ants, for which there is the best
             data) to illustrate concepts or trends and make a number of
             predictions concerning how a particular trait is expected to
             correlate with other aspects of social complexity. Within
             each section we also expand the scope of the arguments to
             consider these relationships in a much broader sense
             of'sociality' by drawing parallels with other 'social'
             entities such as multicellular individuals, which can be
             understood as 'societies' of cells. The aim is to draw out
             any parallels and common causal relationships among the
             correlates. Two themes run through the study. The first is
             the role of colony size as an important factor affecting
             social complexity. The second is the complexity of
             individual workers in relation to the complexity of the
             colony. Consequently, this is an ideal opportunity to test a
             previously proposed hypothesis that 'individuals of highly
             social ant species are less complex than individuals from
             simple ant species' in light of numerous social correlates.
             Our findings support this hypothesis. In summary, we
             conclude that, in general, complex societies are
             characterized by large colony size, worker polymorphism,
             strong behavioural specialization and loss of totipotency in
             its workers, low individual complexity, decentralized colony
             control and high system redundancy, low individual
             competence, a high degree of worker cooperation wher
             tackling tasks, group foraging strategies, high tempo,
             multi-chambered tailor-made nests, high functional
             integration, relatively greater use of cues and modulatory
             signals to coordinate individuals and heterogeneous patterns
             of worker-worker interaction.},
   Doi = {10.1017/s1464793101005656},
   Key = {fds229084}
}

@article{fds229094,
   Author = {Anderson, C and McShea, DW},
   Title = {Intermediate-level parts in insect societies: Adaptive
             structures that ants build away from the
             nest},
   Journal = {Insectes Sociaux},
   Volume = {48},
   Number = {4},
   Pages = {291-301},
   Publisher = {Springer Nature},
   Year = {2001},
   Month = {January},
   ISSN = {0020-1812},
   url = {http://dx.doi.org/10.1007/PL00001781},
   Abstract = {Insect societies function at various organisational levels.
             Most research has focused on one or other organisational
             extreme. At one extreme, it is the adaptive behaviours at
             the individual level, the behaviour of workers, which is of
             interest. At the other extreme, colony-level adaptive
             behaviour and swarm intelligence is the focus. However,
             between these two extremes, numerous functional adaptive
             units, or "parts," exist. These intermediate-level parts
             include the behavioural properties of "groups" or "teams" in
             which the functionality only emerges at the group-level and
             not within the individuals themselves, and also the
             structural properties of "self-assemblages" in which
             individuals link themselves together to form an adaptive
             configuration, such as a living bridge. We review another
             type of intermediate-level part in insect societies: these
             are the physical structures that ants build away from the
             nest. The structures, that are larger than an individual
             worker but smaller than the colony (hence intermediate),
             include cleared trails, walled trenches, arcades, tunnels,
             outstations, shelters, protective pens, shelters over
             nectaries, food coverings on foraging trails, elevated
             corridors, and bridges. They are found in a diverse range of
             species, and are constructed using a variety of materials.
             We detail the structures built by ants focussing chiefly on
             the adaptive benefits these structures may confer to the
             colony.},
   Doi = {10.1007/PL00001781},
   Key = {fds229094}
}

@article{fds326590,
   Author = {McShea, DW},
   Title = {Logic, passion and the problem of convergence.},
   Journal = {Interface focus},
   Volume = {7},
   Number = {3},
   Pages = {20160122},
   Year = {2017},
   Month = {June},
   url = {http://dx.doi.org/10.1098/rsfs.2016.0122},
   Abstract = {Our estimate of the likelihood of convergence on human-style
             intelligence depends on how we understand our various mental
             capacities. Here I revive David Hume's theory of motivation
             and action to argue that the most common understanding of
             the two conventionally recognized components of
             intelligence-reason and emotion-is confused. We say things
             like, 'Reason can overcome emotion', but to make this
             statement meaningful, we are forced to treat reason as a
             compound notion, as a forced and unhappy mixture of concepts
             that are incommensurate. An alternative is to parse
             intelligence in a different way, into two sets of
             capacities: (i) non-affective capacities, including logic,
             calculation and problem-solving; (ii) affective capacities,
             including wants, preferences and cares, along with the
             emotions. Thus, the question of convergence becomes two
             questions, one having to do with affective and one with
             non-affective capacities. What is the likelihood of
             convergence of these in non-human lineages, in other
             ecologies, on other worlds? Given certain assumptions,
             convergence of the non-affective capacities in thinking
             species seems likely, I argue, while convergence of the
             affective capacities seems much less likely.},
   Doi = {10.1098/rsfs.2016.0122},
   Key = {fds326590}
}

@article{fds229055,
   Author = {McShea, DW},
   Title = {Machine wanting.},
   Journal = {Studies in history and philosophy of biological and
             biomedical sciences},
   Volume = {44},
   Number = {4 Pt B},
   Pages = {679-687},
   Year = {2013},
   Month = {December},
   ISSN = {1369-8486},
   url = {http://www.ncbi.nlm.nih.gov/pubmed/23792091},
   Abstract = {Wants, preferences, and cares are physical things or events,
             not ideas or propositions, and therefore no chain of pure
             logic can conclude with a want, preference, or care. It
             follows that no pure-logic machine will ever want, prefer,
             or care. And its behavior will never be driven in the way
             that deliberate human behavior is driven, in other words, it
             will not be motivated or goal directed. Therefore, if we
             want to simulate human-style interactions with the world, we
             will need to first understand the physical structure of
             goal-directed systems. I argue that all such systems share a
             common nested structure, consisting of a smaller entity that
             moves within and is driven by a larger field that contains
             it. In such systems, the smaller contained entity is
             directed by the field, but also moves to some degree
             independently of it, allowing the entity to deviate and
             return, to show the plasticity and persistence that is
             characteristic of goal direction. If all this is right, then
             human want-driven behavior probably involves a
             behavior-generating mechanism that is contained within a
             neural field of some kind. In principle, for goal
             directedness generally, the containment can be virtual,
             raising the possibility that want-driven behavior could be
             simulated in standard computational systems. But there are
             also reasons to believe that goal-direction works better
             when containment is also physical, suggesting that a new
             kind of hardware may be necessary.},
   Doi = {10.1016/j.shpsc.2013.05.015},
   Key = {fds229055}
}

@article{fds229066,
   Author = {McShea, DW},
   Title = {MECHANISMS OF LARGE-SCALE EVOLUTIONARY TRENDS},
   Journal = {Evolution},
   Volume = {48},
   Number = {6},
   Pages = {1747-1763},
   Publisher = {Wiley},
   Year = {1994},
   Month = {December},
   url = {http://dx.doi.org/10.1111/j.1558-5646.1994.tb02211.x},
   Abstract = {Large-scale evolutionary trends may result from driving
             forces or from passive diffusion in bounded spaces. Such
             trends are persistent directional changes in higher taxa
             spanning significant periods of geological time; examples
             include the frequently cited long-term trends in size,
             complexity, and fitness in life as a whole, as well as
             trends in lesser supraspecific taxa and trends in space. In
             a driven trend, the distribution mean increases on account
             of a force (which may manifest itself as a bias in the
             direction of change) that acts on lineages throughout the
             space in which diversification occurs. In a passive system,
             no pervasive force or bias exists, but the mean increases
             because change in one direction is blocked by a boundary, or
             other inhomogeneity, in some limited region of the space.
             Two tests have been used to distinguish these trend
             mechanisms: (1) the test based on the behavior of the
             minimum; and (2) the ancestor-descendant test, based on
             comparisons in a random sample of ancestor-descendant pairs
             that lie far from any possible lower bound. For skewed
             distributions, a third test is introduced here: (3) the
             subclade test, based on the mean skewness of a sample of
             subclades drawn from the tail of a terminal distribution.
             With certain restrictions, a system is driven if the minimum
             increases, if increases significantly outnumber decreases
             among ancestor-descendant pairs, and if the mean skew of
             subclades is significantly positive. A passive mechanism is
             more difficult to demonstrate but is the more likely
             mechanism if decreases outnumber increases and if the mean
             skew of subclades is negative. Unlike the other tests, the
             subclade test requires no detailed phylogeny or
             paleontological time series, but only terminal (e.g.,
             modern) distributions. Monte Carlo simulations of the
             diversification of a clade are used to show how the subclade
             test works. In the empirical cases examined, the three tests
             gave concordant results, suggesting first, that they work,
             and second, that the passive and driven mechanisms may
             correspond to natural categories of causes of large-scale
             trends.},
   Doi = {10.1111/j.1558-5646.1994.tb02211.x},
   Key = {fds229066}
}

@article{fds365439,
   Author = {Lee, JG and McShea, D},
   Title = {Operationalizing goal directedness: An empirical route to
             advancing a philosophical discussion},
   Journal = {Philosophy, Theory, and Practice in Biology},
   Volume = {12},
   Number = {5},
   Publisher = {University of Michigan Library},
   Year = {2020},
   Month = {June},
   url = {http://dx.doi.org/10.3998/ptpbio.16039257.0012.005},
   Doi = {10.3998/ptpbio.16039257.0012.005},
   Key = {fds365439}
}

@article{fds229081,
   Author = {Marino, L and McShea, DW and Uhen, MD},
   Title = {Origin and evolution of large brains in toothed
             whales.},
   Journal = {The anatomical record. Part A, Discoveries in molecular,
             cellular, and evolutionary biology},
   Volume = {281},
   Number = {2},
   Pages = {1247-1255},
   Year = {2004},
   Month = {December},
   ISSN = {1552-4884},
   url = {http://www.ncbi.nlm.nih.gov/pubmed/15497142},
   Abstract = {Toothed whales (order Cetacea: suborder Odontoceti) are
             highly encephalized, possessing brains that are
             significantly larger than expected for their body sizes. In
             particular, the odontocete superfamily Delphinoidea
             (dolphins, porpoises, belugas, and narwhals) comprises
             numerous species with encephalization levels second only to
             modern humans and greater than all other mammals.
             Odontocetes have also demonstrated behavioral faculties
             previously only ascribed to humans and, to some extent,
             other great apes. How did the large brains of odontocetes
             evolve? To begin to investigate this question, we quantified
             and averaged estimates of brain and body size for 36 fossil
             cetacean species using computed tomography and analyzed
             these data along with those for modern odontocetes. We
             provide the first description and statistical tests of the
             pattern of change in brain size relative to body size in
             cetaceans over 47 million years. We show that brain size
             increased significantly in two critical phases in the
             evolution of odontocetes. The first increase occurred with
             the origin of odontocetes from the ancestral group
             Archaeoceti near the Eocene-Oligocene boundary and was
             accompanied by a decrease in body size. The second occurred
             in the origin of Delphinoidea only by 15 million years
             ago.},
   Doi = {10.1002/ar.a.20128},
   Key = {fds229081}
}

@article{fds3809,
   Author = {D.W. McShea},
   Title = {Parts and integration: consequences of hierarchy},
   Series = {pp. 27-60},
   Booktitle = {Evolutionary Patterns: Growth, Form, and Tempo in the Fossil
             Record},
   Publisher = {Chicago: Univ. Chicago Press},
   Editor = {JBC Jackson and S Lidgard and FK McKinney},
   Year = {2001},
   Key = {fds3809}
}

@article{fds229067,
   Author = {McShea, DW},
   Title = {Perspective: Metazoan Complexity and Evolution: Is There a
             Trend?},
   Journal = {Evolution},
   Volume = {50},
   Number = {2},
   Pages = {477-477},
   Publisher = {Oxford University Press (OUP)},
   Year = {1996},
   Month = {April},
   url = {http://dx.doi.org/10.2307/2410824},
   Abstract = {The notion that complexity increases in evolution is widely
             accepted, but the best-known evidence is highly
             impressionistic. Here I propose a scheme for understanding
             complexity that provides a conceptual basis for objective
             measurement. The scheme also shows complexity to be a broad
             term covering four independent types. For each type, I
             describe some of the measures that have been devised and
             review the evidence for trends in the maximum and mean. In
             metazoans as a whole, there is good evidence only for an
             early-Phanerozoic trend, and only in one type of complexity.
             For each of the other types, some trends have been
             documented, but only in a small number of metazoan
             subgroups.},
   Doi = {10.2307/2410824},
   Key = {fds229067}
}

@article{fds229070,
   Author = {Liow, LH and Simpson, C and Bouchard, F and Damuth, J and Hallgrimsson,
             B and Hunt, G and McShea, DW and Powell, JR and Stenseth, NC and Stoller,
             MK and Wagner, G},
   Title = {Pioneering paradigms and magnificent manifestos--Leigh Van
             Valen's priceless contributions to evolutionary
             biology.},
   Journal = {Evolution; international journal of organic
             evolution},
   Volume = {65},
   Number = {4},
   Pages = {917-922},
   Year = {2011},
   Month = {April},
   url = {http://www.ncbi.nlm.nih.gov/pubmed/21463292},
   Doi = {10.1111/j.1558-5646.2011.01242.x},
   Key = {fds229070}
}

@article{fds229082,
   Author = {McShea, DW},
   Title = {Possible largest-scale trends in organismal evolution: Eight
             'live hypotheses'},
   Journal = {Annual Review of Ecology and Systematics},
   Volume = {29},
   Number = {1},
   Pages = {293-318},
   Publisher = {ANNUAL REVIEWS},
   Year = {1998},
   Month = {December},
   url = {http://dx.doi.org/10.1146/annurev.ecolsys.29.1.293},
   Abstract = {Historically, a great many features of organisms have been
             said to show a trend over the history of life, and many
             rationales for such trends have been proposed. Here I review
             eight candidates, eight 'live hypotheses' that are inspiring
             research on largest-scale trends today: entropy, energy
             intensiveness, evolutionary versatility, developmental
             depth, structural depth, adaptedness, size, and complexity.
             For each, the review covers the principal arguments that
             have been advanced for why a trend is expected, as well as
             some of the empirical approaches that have been adopted.
             Also discussed are three conceptual matters arising in
             connection with trend studies: 1. Alternative bases for
             classifying trends: pattern versus dynamics; 2. alternative
             modes in which largest-scale trends have been studied:
             'exploratory' versus 'skeptical'; and 3. evolutionary
             progress.},
   Doi = {10.1146/annurev.ecolsys.29.1.293},
   Key = {fds229082}
}

@article{fds366840,
   Author = {Rosenberg, A and McShea, DW},
   Title = {Reductionism about biology},
   Pages = {96-126},
   Booktitle = {PHILOSOPHY OF BIOLOGY: A CONTEMPORARY INTRODUCTION},
   Year = {2008},
   Key = {fds366840}
}

@article{fds366835,
   Author = {Babcock, G and McShea, DW},
   Title = {Resolving teleology's false dilemma},
   Journal = {Biological Journal of the Linnean Society},
   Volume = {139},
   Number = {4},
   Pages = {415-432},
   Publisher = {Oxford University Press (OUP)},
   Year = {2023},
   Month = {August},
   url = {http://dx.doi.org/10.1093/biolinnean/blac058},
   Abstract = {This paper argues that the account of teleology previously
             proposed by the authors is consistent with the physical
             determinism that is implicit across many of the sciences. We
             suggest that much of the current aversion to teleological
             thinking found in the sciences is rooted in debates that can
             be traced back to ancient natural science, which pitted
             mechanistic and deterministic theories against teleological
             ones. These debates saw a deterministic world as one where
             freedom and agency is impossible. And, because teleological
             entities seem to be free to either reach their ends or not,
             it was assumed that they could not be deterministic. Mayr's
             modern account of teleonomy adheres to this basic
             assumption. Yet, the seeming tension between teleology and
             determinism is illusory because freedom and agency do not,
             in fact, conflict with a deterministic world. To show this,
             we present a taxonomy of different types of freedom that we
             see as inherent in teleological systems. Then we show that
             our taxonomy of freedom, which is crucial to understanding
             teleology, shares many of the features of a philosophical
             position regarding free will that is known in the
             contemporary literature as 'compatibilism'. This position
             maintains that an agent is free when the sources of its
             actions are internal, when the agent itself is the
             deterministic cause of those actions. Our view shows that
             freedom is not only indispensable to teleology, but also
             that, contrary to common intuitions, there is no conflict
             between teleology and causal determinism.},
   Doi = {10.1093/biolinnean/blac058},
   Key = {fds366835}
}

@article{fds229061,
   Author = {McShea, DW},
   Title = {Sense and Depth},
   Journal = {Biology & Philosophy},
   Volume = {15},
   Number = {5},
   Pages = {751-758},
   Publisher = {Springer Science and Business Media LLC},
   Year = {2000},
   Month = {November},
   ISSN = {0169-3867},
   url = {http://gateway.webofknowledge.com/gateway/Gateway.cgi?GWVersion=2&SrcApp=PARTNER_APP&SrcAuth=LinksAMR&KeyUT=WOS:000165088500008&DestLinkType=FullRecord&DestApp=ALL_WOS&UsrCustomerID=47d3190e77e5a3a53558812f597b0b92},
   Doi = {10.1023/a:1006754311040},
   Key = {fds229061}
}

@article{fds368096,
   Author = {Keenan, JP and McShea, DW},
   Title = {Synergies Among Behaviors Drive the Discovery of Productive
             Interactions},
   Journal = {Biological Theory},
   Volume = {18},
   Number = {1},
   Pages = {43-62},
   Year = {2023},
   Month = {March},
   url = {http://dx.doi.org/10.1007/s13752-022-00420-2},
   Abstract = {When behaviors assemble into combinations, then synergies
             have a central role in the discovery of productive patterns
             of behavior. In our view—what we call the Synergy
             Emergence Principle (SEP)—synergies are dynamic
             attractors, drawing interactions toward greater returns as
             they happen, in the moment. This Principle offers an
             alternative to the two conventionally acknowledged routes to
             discovery: directed problem solving, involving forethought
             and planning; and the complete randomness of trial and
             error. Natural selection has a role in the process, in
             humans favoring the maintenance and improvement of certain
             key underlying capabilities, such as prosocial helping and
             episodic foresight, but selection is not required for
             discovery by synergy (which occurs too rapidly for selection
             anyway). Here we discuss the consequences of the SEP for
             the evolution in humans of key synergies such as tool
             usage and interactions that reward cooperation, show how
             discovery by synergy and the selection of synergy-supporting
             abilities formed a positive feedback loop, and show how
             synergies can combine, forming clusters and packages that
             are the core of institutions and cultures. Finally, clusters
             and packages represent an intermediate level of organization
             above the individual and below whole society, with
             consequences for our understanding of the major transitions
             in evolution.},
   Doi = {10.1007/s13752-022-00420-2},
   Key = {fds368096}
}

@article{fds229087,
   Author = {McShea, DW and Venit, EP},
   Title = {Testing for bias in the evolution of coloniality: A
             demonstration in cyclostome bryozoans},
   Journal = {Paleobiology},
   Volume = {28},
   Number = {3},
   Pages = {308-327},
   Publisher = {Cambridge University Press (CUP)},
   Year = {2002},
   Month = {Summer},
   url = {http://dx.doi.org/10.1666/0094-8373(2002)028<0308:TFBITE>2.0.CO;2},
   Abstract = {Colonial organisms vary in the degree to which they are
             individuated at the colony level, i.e., in the degree to
             which the colony constitutes a unified whole, as opposed to
             a group of independent lower-level entities. Various
             arguments have been offered suggesting that evolutionary
             change along this continuum may be biased, that increases
             may be more probable than decreases. However,
             counterarguments can be devised, and the existing evidence
             is meager and inconclusive. In this paper, we demonstrate
             how the question can be addressed empirically by conducting
             a test for bias in a group of stenolaemate bryozoans, the
             cyclostomes. More specifically, we suggest three criteria
             for colony individuation: degree of connectedness among
             lower-level entities (in this case, zooids), degree of
             differentiation among lower-level entities, and number of
             intermediate-level parts. And we show these criteria can be
             used together with a phylogeny and ancestral-state
             reconstruction methods to test for bias. In this case,
             results do not unambiguously support any single
             interpretation but are somewhat supportive of a null
             hypothesis of no bias in favor of increase. As part of the
             demonstration, we also show how results can be transformed
             into a quantitative estimate of an upper limit on bias.
             Finally, we place the question of bias in a larger context,
             arguing that the same criteria and methods we employ here
             can be used to test for bias in other colonial taxa, and
             also at other hierarchical levels, for example, in the
             transitions from free-living eukaryotic cells to
             multicellular organisms. © 2002 Paleontological
             Society.},
   Doi = {10.1666/0094-8373(2002)028<0308:TFBITE>2.0.CO;2},
   Key = {fds229087}
}

@article{fds229085,
   Author = {Anderson, C and Franks, NR and McShea, DW},
   Title = {The complexity and hierarchical structure of tasks in insect
             societies},
   Journal = {Animal Behaviour},
   Volume = {62},
   Number = {4},
   Pages = {643-651},
   Publisher = {Elsevier BV},
   Year = {2001},
   Month = {January},
   ISSN = {0003-3472},
   url = {http://dx.doi.org/10.1006/anbe.2001.1795},
   Abstract = {To understand the functioning and organizational complexity
             of insect societies, a combination of different approaches
             is needed. One such approach, which we adopt in this study,
             is to consider tasks in insect societies not based upon
             their function, as is traditional, but upon their structure.
             Four types of task in insect societies have been proposed:
             individual, group, team and partitioned tasks. We examine
             the relationships among these four task types and consider
             'task complexity' to mean the degree of cooperation and
             coordination required to complete a particular task
             successfully. In this respect, individual tasks are
             considered the simplest (low complexity), group tasks are
             more complex (medium), and team and partitioned tasks the
             most complex (high). We decompose tasks into their component
             subtasks to understand how the demands of a task influence
             how workers must work together to complete it successfully.
             We describe a simple method to measure the complexity of
             tasks using task deconstruction. Points are assigned to each
             subtask within the task and summed to give a total score.
             This measure, the task's score, allows objective comparison
             of tasks (different tasks may be ranked in order of their
             complexity) within and between species, or even higher taxa,
             and we hope it will be of practical use to researchers. We
             propose that both team and partitioned tasks may contain
             individual, group, team and partitioned subtasks. We examine
             each of the possible task-subtask relationships and provide
             examples from known social insect behaviour. © 2001 The
             Association for the Study of Animal Behaviour.},
   Doi = {10.1006/anbe.2001.1795},
   Key = {fds229085}
}

@article{fds229078,
   Author = {McShea, DW},
   Title = {The evolution of complexity without natural selection, a
             possible large-scale trend of the fourth
             kind},
   Journal = {Paleobiology},
   Volume = {31},
   Number = {2 SUPPL.},
   Pages = {146-156},
   Publisher = {Cambridge University Press (CUP)},
   Year = {2005},
   Month = {July},
   url = {http://dx.doi.org/10.1666/0094-8373(2005)031[0146:TEOCWN]2.0.CO;2},
   Abstract = {A simple principle predicts a tendency, or vector, toward
             increasing organismal complexity in the history of life: As
             the parts of an organism accumulate variations in evolution,
             they should tend to become more different from each other.
             In other words, the variance among the parts, or what I call
             the "internal variance" of the organism, will tend to
             increase spontaneously. Internal variance is complexity, I
             argue, albeit complexity in a purely structural sense,
             divorced from any notion of function. If the principle is
             correct, this tendency should exist in all lineages, and the
             resulting trend (if there is one) will be driven, or more
             precisely, driven by constraint (as opposed to selection).
             The existence of a trend is uncertain, because the
             internal-variance principle predicts only that the range of
             options offered up to selection will be increasingly
             complex, on average. And it is unclear whether selection
             will enhance this vector, act neutrally, or oppose it,
             perhaps negating it. The vector might also be negated if
             variations producing certain kinds of developmental
             truncations are especially common in evolution.
             Constraint-driven trends - or what I ca ll large-scale
             trends of the fourth kind - have been in bad odor in
             evolutionary studies since the Modern Synthesis. Indeed, one
             such trend, orthogenesis, is famous for having been
             discredited. In Stephen Jay Gould's last book, The Structure
             of Evolutionary Thought, he tried to rehabilitate this
             category (although not orthogenesis), showing how
             constraint-driven trends could be produced by processes well
             within the mainstream of contemporary evolutionary theory.
             The internal-variance principle contributes to Gould's
             project by adding another candidate trend to this category.
             © 2005 The Paleontological Society. All rights
             reserved.},
   Doi = {10.1666/0094-8373(2005)031[0146:TEOCWN]2.0.CO;2},
   Key = {fds229078}
}

@article{fds229075,
   Author = {Payne, JL and McClain, CR and Boyer, AG and Brown, JH and Finnegan, S and Kowalewski, M and Krause, RA and Lyons, SK and McShea, DW and Novack-Gottshall, PM and Smith, FA and Spaeth, P and Stempien, JA and Wang, SC},
   Title = {The evolutionary consequences of oxygenic photosynthesis: a
             body size perspective.},
   Journal = {Photosynthesis research},
   Volume = {107},
   Number = {1},
   Pages = {37-57},
   Year = {2011},
   Month = {January},
   url = {http://www.ncbi.nlm.nih.gov/pubmed/20821265},
   Abstract = {The high concentration of molecular oxygen in Earth's
             atmosphere is arguably the most conspicuous and geologically
             important signature of life. Earth's early atmosphere lacked
             oxygen; accumulation began after the evolution of oxygenic
             photosynthesis in cyanobacteria around 3.0-2.5 billion
             years ago (Gya). Concentrations of oxygen have since varied,
             first reaching near-modern values ~600 million years ago
             (Mya). These fluctuations have been hypothesized to
             constrain many biological patterns, among them the evolution
             of body size. Here, we review the state of knowledge
             relating oxygen availability to body size. Laboratory
             studies increasingly illuminate the mechanisms by which
             organisms can adapt physiologically to the variation in
             oxygen availability, but the extent to which these findings
             can be extrapolated to evolutionary timescales remains
             poorly understood. Experiments confirm that animal size is
             limited by experimental hypoxia, but show that plant
             vegetative growth is enhanced due to reduced
             photorespiration at lower O(2):CO(2). Field studies of size
             distributions across extant higher taxa and individual
             species in the modern provide qualitative support for a
             correlation between animal and protist size and oxygen
             availability, but few allow prediction of maximum or mean
             size from oxygen concentrations in unstudied regions. There
             is qualitative support for a link between oxygen
             availability and body size from the fossil record of
             protists and animals, but there have been few quantitative
             analyses confirming or refuting this impression. As oxygen
             transport limits the thickness or volume-to-surface area
             ratio-rather than mass or volume-predictions of maximum
             possible size cannot be constructed simply from metabolic
             rate and oxygen availability. Thus, it remains difficult to
             confirm that the largest representatives of fossil or living
             taxa are limited by oxygen transport rather than other
             factors. Despite the challenges of integrating findings from
             experiments on model organisms, comparative observations
             across living species, and fossil specimens spanning
             millions to billions of years, numerous tractable avenues of
             research could greatly improve quantitative constraints on
             the role of oxygen in the macroevolutionary history of
             organismal size.},
   Doi = {10.1007/s11120-010-9593-1},
   Key = {fds229075}
}

@article{fds229071,
   Author = {Kowalewski, M and Payne, JL and Smith, FA and Wang, SC and McShea, DW and Xiao, S and Novack-Gottshall, PM and McClain, CR and Krause, RA and Boyer, AG and Finnegan, S and Lyons, SK and Stempien, JA and Alroy, J and Spaeth, PA},
   Title = {The geozoic supereon},
   Journal = {Palaios},
   Volume = {26},
   Number = {5},
   Pages = {251-255},
   Publisher = {Society for Sedimentary Geology},
   Year = {2011},
   Month = {May},
   ISSN = {0883-1351},
   url = {http://dx.doi.org/10.2110/palo.2011.S03},
   Doi = {10.2110/palo.2011.S03},
   Key = {fds229071}
}

@article{fds229090,
   Author = {McShea, DW},
   Title = {The hierarchical structure of organisms: A scale and
             documentation of a trend in the maximum},
   Journal = {Paleobiology},
   Volume = {27},
   Number = {2},
   Pages = {405-423},
   Publisher = {Cambridge University Press (CUP)},
   Year = {2001},
   Month = {January},
   url = {http://dx.doi.org/10.1666/0094-8373(2001)027<0405:THSOOA>2.0.CO;2},
   Abstract = {The degree of hierarchical structure of organisms-the number
             of levels of nesting of lower-level entities within
             higher-level individuals-has apparently increased a number
             of times in the history of life, notably in the origin of
             the eukaryotic cell from an association of prokaryotic
             cells, of multicellular organisms from clones of eukaryotic
             cells, and of intergrated colonies from aggregates of
             multicellular individuals. Arranged in order of first
             occurrence, these three transitions suggest a trend, in
             particular a trend in the maximum, or an increase in the
             degree of hierarchical structure present in the
             hierarchically deepest organism on Earth. However, no
             rigorous documentation of such a trend-based on operational
             and consistent criteria for hierarchical levels-has been
             attempted. Also, the trajectory of increase has not been
             examined in any detail. One limitation is that no hierarchy
             scale has been developed with sufficient resolution to
             document more than these three major increases. Here, a
             higher-resolution scale is proposed in which hierarchical
             structure is decomposed into levels and sublevels, with
             levels reflecting number of layers of nestedness, and
             sublevels reflecting degree of individuation at the highest
             level. The scale is then used, together with the body-fossil
             record, to plot the trajectory of the maximum. Two
             alternative interpretations of the record are considered,
             and both reveal a long-term trend extending from the Archean
             through the early Phanerozoic. In one, the pattern of
             increase was incremental, with almost all sublevels arising
             precisely in order. The data also raise the possibility that
             waiting times for transitions between sublevels may have
             decreased with increasing hierarchical level (and with
             time). These last two findings-incremental increase in level
             and decreasing waiting times-are tentative, pending a study
             of possible biases in the fossil record.},
   Doi = {10.1666/0094-8373(2001)027<0405:THSOOA>2.0.CO;2},
   Key = {fds229090}
}

@article{fds229091,
   Author = {McShea, DW},
   Title = {The minor transitions in hierarchical evolution and the
             question of a directional bias},
   Journal = {Journal of Evolutionary Biology},
   Volume = {14},
   Number = {3},
   Pages = {502-518},
   Publisher = {WILEY},
   Year = {2001},
   Month = {July},
   ISSN = {1010-061X},
   url = {http://dx.doi.org/10.1046/j.1420-9101.2001.00283.x},
   Abstract = {The history of life shows a clear trend in hierarchical
             organization, revealed by the successive emergence of
             organisms with ever greater numbers of levels of nestedness
             and greater development, or 'individuation', of the highest
             level. Various arguments have been offered which suggest
             that the trend is the result of a directional bias, or
             tendency, meaning that hierarchical increases are more
             probable than decreases among lineages, perhaps because
             hierarchical increases are favoured, on average, by natural
             selection. Further, what little evidence exists seems to
             point to a bias: some major increases are known-including
             the origin of the eukaryotic cell from prokaryotic cells and
             of animals, fungi and land plants from solitary eukaryotic
             cells - but no major decreases (except in parasitic and
             commensal organisms), at least at the cellular and
             multicellular levels. The fact of a trend, combined with the
             arguments and evidence, might make a bias seem beyond doubt,
             but here I argue that its existence is an open empirical
             question. Further, I show how testing is
             possible.},
   Doi = {10.1046/j.1420-9101.2001.00283.x},
   Key = {fds229091}
}

@article{fds229046,
   Author = {McShea, DW and Simpson, CG},
   Title = {The miscellaneous transitions in evolution},
   Pages = {19-34},
   Booktitle = {The Major Transitions in Evolution Revisited},
   Publisher = {MIT Press},
   Editor = {Calcott, B and Sterelny, K},
   Year = {2011},
   ISBN = {9780262294539},
   url = {http://dx.doi.org/10.7551/mitpress/9780262015240.003.0002},
   Doi = {10.7551/mitpress/9780262015240.003.0002},
   Key = {fds229046}
}

@article{fds366836,
   Author = {Brandon, RN and McShea, DW},
   Title = {The Missing Two-Thirds of Evolutionary Theory},
   Pages = {1-+},
   Booktitle = {MISSING TWO-THIRDS OF EVOLUTIONARY THEORY},
   Year = {2020},
   url = {http://dx.doi.org/10.1017/9781108591508},
   Doi = {10.1017/9781108591508},
   Key = {fds366836}
}

@article{fds43079,
   Author = {D.W. McShea and C. Anderson},
   Title = {The remodularization of the organism},
   Pages = {185-206},
   Booktitle = {Modularity: Understanding the Development and Evolution of
             Natural Complex Systems},
   Publisher = {The MIT Press},
   Editor = {W. Callebaut and D. Rasskin-Gutman},
   Year = {2005},
   Key = {fds43079}
}

@article{fds229057,
   Author = {McShea, DW},
   Title = {Three provocative patterns in hierarchical
             evolution.},
   Journal = {AMERICAN ZOOLOGIST},
   Volume = {41},
   Number = {6},
   Pages = {1522-1522},
   Publisher = {SOC INTEGRATIVE COMPARATIVE BIOLOGY},
   Year = {2001},
   Month = {December},
   ISSN = {0003-1569},
   url = {http://gateway.webofknowledge.com/gateway/Gateway.cgi?GWVersion=2&SrcApp=PARTNER_APP&SrcAuth=LinksAMR&KeyUT=WOS:000174306500495&DestLinkType=FullRecord&DestApp=ALL_WOS&UsrCustomerID=47d3190e77e5a3a53558812f597b0b92},
   Key = {fds229057}
}

@article{fds229093,
   Author = {McShea, DW and Changizi, MA},
   Title = {Three puzzles in hierarchical evolution.},
   Journal = {Integrative and comparative biology},
   Volume = {43},
   Number = {1},
   Pages = {74-81},
   Year = {2003},
   Month = {February},
   ISSN = {1540-7063},
   url = {http://www.ncbi.nlm.nih.gov/pubmed/21680411},
   Abstract = {The maximum degree of hierarchical structure of organisms
             has risen over the history of life, notably in three
             transitions: the origin of the eukaryotic cell from
             symbiotic associations of prokaryotes; the emergence of the
             first multicellular individuals from clones of eukaryotic
             cells; and the origin of the first individuated colonies
             from associations of multicellular organisms. The trend is
             obvious in the fossil record, but documenting it using a
             high-resolution hierarchy scale reveals three puzzles: 1)
             the rate of origin of new levels accelerates, at least until
             the early Phanerozoic; 2) after that, the trend may slow or
             even stop; and 3) levels may sometimes arise out of order.
             The three puzzles and their implications are discussed; a
             possible explanation is offered for the first.},
   Doi = {10.1093/icb/43.1.74},
   Key = {fds229093}
}

@article{fds229048,
   Author = {McShea, DW},
   Title = {Three Trends in the History of Life: An Evolutionary
             Syndrome},
   Journal = {Evolutionary Biology},
   Volume = {43},
   Number = {4},
   Pages = {531-542},
   Publisher = {Springer Nature},
   Year = {2016},
   Month = {December},
   ISSN = {0071-3260},
   url = {http://dx.doi.org/10.1007/s11692-015-9323-x},
   Abstract = {The history of life seems to be characterized by three
             large-scale trends in complexity: (1) the rise in complexity
             in the sense of hierarchy, in other words, an increase in
             the number of levels of organization within organisms; (2)
             the increase in complexity in the sense of differentiation,
             that is, a rise in the number of different part types at the
             level just below the whole; and (3) a downward trend, the
             loss of differentiation at the lowest levels in organisms, a
             kind of complexity drain within the parts. Here, I describe
             the three trends, outlining the evidence for each and
             arguing that they are connected with each other, that
             together they constitute an evolutionary syndrome, one that
             has recurred a number times over the history of life.
             Finally, in the last section, I offer an argument connecting
             the third trend to the reduction at lower levels of
             organization in “autonomy”, or from a different
             perspective, to an increase in what might be called the
             “machinification” of the lower levels.},
   Doi = {10.1007/s11692-015-9323-x},
   Key = {fds229048}
}

@article{fds229086,
   Author = {McShea, DW},
   Title = {Trends, tools, and terminology},
   Journal = {PALEOBIOLOGY},
   Volume = {26},
   Number = {3},
   Pages = {330-333},
   Publisher = {Cambridge University Press (CUP)},
   Year = {2000},
   ISSN = {0094-8373},
   url = {http://gateway.webofknowledge.com/gateway/Gateway.cgi?GWVersion=2&SrcApp=PARTNER_APP&SrcAuth=LinksAMR&KeyUT=WOS:000089104700002&DestLinkType=FullRecord&DestApp=ALL_WOS&UsrCustomerID=47d3190e77e5a3a53558812f597b0b92},
   Doi = {10.1666/0094-8373(2000)026<0330:TTAT>2.0.CO;2},
   Key = {fds229086}
}

@article{fds229076,
   Author = {Payne, JL and Boyer, AG and Brown, JH and Finnegan, S and Kowalewski, M and Krause, RA and Lyons, SK and McClain, CR and McShea, DW and Novack-Gottshall, PM and Smith, FA and Stempien, JA and Wang,
             SC},
   Title = {Two-phase increase in the maximum size of life over 3.5
             billion years reflects biological innovation and
             environmental opportunity.},
   Journal = {Proceedings of the National Academy of Sciences of the
             United States of America},
   Volume = {106},
   Number = {1},
   Pages = {24-27},
   Year = {2009},
   Month = {January},
   url = {http://www.ncbi.nlm.nih.gov/pubmed/19106296},
   Abstract = {The maximum size of organisms has increased enormously since
             the initial appearance of life >3.5 billion years ago (Gya),
             but the pattern and timing of this size increase is poorly
             known. Consequently, controls underlying the size spectrum
             of the global biota have been difficult to evaluate. Our
             period-level compilation of the largest known fossil
             organisms demonstrates that maximum size increased by 16
             orders of magnitude since life first appeared in the fossil
             record. The great majority of the increase is accounted for
             by 2 discrete steps of approximately equal magnitude: the
             first in the middle of the Paleoproterozoic Era
             (approximately 1.9 Gya) and the second during the late
             Neoproterozoic and early Paleozoic eras (0.6-0.45 Gya). Each
             size step required a major innovation in organismal
             complexity--first the eukaryotic cell and later eukaryotic
             multicellularity. These size steps coincide with, or
             slightly postdate, increases in the concentration of
             atmospheric oxygen, suggesting latent evolutionary potential
             was realized soon after environmental limitations were
             removed.},
   Doi = {10.1073/pnas.0806314106},
   Key = {fds229076}
}

@article{fds229051,
   Author = {McShea, DW},
   Title = {Unnecessary Complexity  Complexity and the Arrow of Time
             Charles H. Lineweaver, Paul C. W. Davies, and Michael
             Ruse, Eds.  Cambridge University Press, Cambridge,
             2013. 369 pp. $30, £21.99. ISBN 9781107027251.},
   Journal = {Science},
   Volume = {342},
   Number = {6164},
   Pages = {1319-1320},
   Publisher = {American Association for the Advancement of Science
             (AAAS)},
   Year = {2013},
   Month = {December},
   ISSN = {0036-8075},
   url = {http://gateway.webofknowledge.com/gateway/Gateway.cgi?GWVersion=2&SrcApp=PARTNER_APP&SrcAuth=LinksAMR&KeyUT=WOS:000328196000028&DestLinkType=FullRecord&DestApp=ALL_WOS&UsrCustomerID=47d3190e77e5a3a53558812f597b0b92},
   Abstract = {<jats:p>The contributors examine the nature of complexity
             and its changes over time as well as their
             causes.</jats:p>},
   Doi = {10.1126/science.1245386},
   Key = {fds229051}
}

@article{fds229059,
   Author = {Mcshea, DW},
   Title = {Unpredictability! and the Function of Mind in
             Nature},
   Journal = {Adaptive Behavior},
   Volume = {4},
   Number = {4},
   Pages = {466-470},
   Publisher = {SAGE Publications},
   Year = {1996},
   Month = {January},
   ISSN = {1059-7123},
   url = {http://gateway.webofknowledge.com/gateway/Gateway.cgi?GWVersion=2&SrcApp=PARTNER_APP&SrcAuth=LinksAMR&KeyUT=WOS:A1996VY98400009&DestLinkType=FullRecord&DestApp=ALL_WOS&UsrCustomerID=47d3190e77e5a3a53558812f597b0b92},
   Doi = {10.1177/105971239600400309},
   Key = {fds229059}
}

@article{fds229069,
   Author = {McShea, DW},
   Title = {Untangling the morass},
   Journal = {American Scientist},
   Volume = {99},
   Number = {2},
   Pages = {154-156},
   Publisher = {Sigma Xi},
   Year = {2011},
   Month = {January},
   ISSN = {0003-0996},
   url = {http://dx.doi.org/10.1511/2011.89.154},
   Abstract = {Daniel W. McShea reviews the book 'The mirage of a space
             between nature and nurture,' by Evelyn Fox Keller. Keller
             argues that much of the trouble has to do with linguistic
             practice, with slippages in usage and concepts. In her apt
             words, the nature- nurture debate is a 'morass of linguistic
             and conceptual vegetation grown together in ways that seem
             to defy untangling.' The address on an envelope makes a huge
             difference in where the letter goes but has little to do
             with generating the process that actually gets it there. A
             mutant allele associated with a speech and language disorder
             still gets labeled a speech and language gene. Most
             behavioral geneticists would agree that a mistake has been
             made when explicit claims about the genetic basis of
             individual traits are inferred from measures of technical
             heritability.},
   Doi = {10.1511/2011.89.154},
   Key = {fds229069}
}

@article{fds229074,
   Author = {McShea, DW},
   Title = {Upper-directed systems: A new approach to teleology in
             biology},
   Journal = {Biology and Philosophy},
   Volume = {27},
   Number = {5},
   Pages = {663-684},
   Publisher = {Springer Nature},
   Year = {2012},
   Month = {September},
   ISSN = {0169-3867},
   url = {http://dx.doi.org/10.1007/s10539-012-9326-2},
   Abstract = {How shall we understand apparently teleological systems?
             What explains their persistence (returning to past
             trajectories following errors) and their plasticity (finding
             the same trajectory from different starting points)? Here I
             argue that all seemingly goal-directed systems-e. g., a
             food-seeking organism, human-made devices like thermostats
             and torpedoes, biological development, human goal seeking,
             and the evolutionary process itself-share a common
             organization. Specifically, they consist of an entity that
             moves within a larger containing structure, one that directs
             its behavior in a general way without precisely determining
             it. If so, then teleology lies within the domain of the
             theory of compositional hierarchies. © 2012 Springer
             Science+Business Media B.V.},
   Doi = {10.1007/s10539-012-9326-2},
   Key = {fds229074}
}

@article{fds355922,
   Author = {McShea, DW and Venit, EP},
   Title = {What is a Part?},
   Pages = {259-284},
   Booktitle = {The Character Concept in Evolutionary Biology},
   Publisher = {Elsevier},
   Editor = {G.P. Wagner},
   Year = {2001},
   ISBN = {9780127300559},
   url = {http://dx.doi.org/10.1016/b978-012730055-9/50022-7},
   Doi = {10.1016/b978-012730055-9/50022-7},
   Key = {fds355922}
}

@article{fds366844,
   Author = {Rosenberg, A and McShea, DW},
   Title = {What is the philosophy of biology? Introduction},
   Pages = {1-+},
   Booktitle = {PHILOSOPHY OF BIOLOGY: A CONTEMPORARY INTRODUCTION},
   Year = {2008},
   Key = {fds366844}
}


%% Book Reviews   
@article{fds197562,
   Author = {D.W. McShea},
   Title = {(Review of The Tangled Web, by Carl Zimmer)},
   Journal = {Quarterly Review of Biology},
   Volume = {86},
   Pages = {47},
   Year = {2011},
   Key = {fds197562}
}

@article{fds16144,
   Author = {D.W. McShea},
   Title = {Adaptive glory},
   Journal = {American Scientist},
   Volume = {91},
   Pages = {567-569},
   Year = {2003},
   Month = {November},
   Key = {fds16144}
}


%% Book Chapter   
@misc{fds219847,
   Author = {D.W. McShea},
   Title = {Freedom and purpose in biology},
   Booktitle = {Contingency and Order in History and the Sciences (working
             title)},
   Editor = {Peter Harrison},
   Year = {2013},
   Key = {fds219847}
}


%% Published Abstracts   
@misc{fds219848,
   Author = {Finnegan, Seth and Steve C. Wang and John Alroy and Alison G. Boyer and Matthew E. Clapham and Zoe V. Finkel and Matthew A. Kosnik and Michał Kowalewski and Richard A. Krause, Jr. and S. Kathleen
             Lyons and Craig R. McClain and Daniel W. McShea and Philip M.
             Novack- Gottshall and Rowan Lockwood and Jonathan L. Payne and Felisa
             A. Smith and Paula A. Spaeth and Jennifer A.
             Stempien},
   Title = {No consistent relationship between body size and extinction
             risk in the marine fossil record},
   Journal = {GSA Abstracts},
   Year = {2009},
   Key = {fds219848}
}