Skip to main content Accessibility help
×
Hostname: page-component-848d4c4894-xm8r8 Total loading time: 0 Render date: 2024-06-22T14:20:39.908Z Has data issue: false hasContentIssue false

8 - Combining Genetic and Developmental Methods to Study Musculoskeletal Evolution in Primates

Published online by Cambridge University Press:  25 March 2017

Christopher J. Percival
Affiliation:
University of Calgary
Joan T. Richtsmeier
Affiliation:
Pennsylvania State University
Get access
Type
Chapter
Information
Publisher: Cambridge University Press
Print publication year: 2017

Access options

Get access to the full version of this content by using one of the access options below. (Log in options will check for institutional or personal access. Content may require purchase if you do not have access.)

References

Akiyama, H., Kim, J. E., Nakashima, K., et al. (2005). Osteo-chondroprogenitor cells are derived from Sox9 expressing precursors. Proceedings of the National Academy of Sciences of the USA, 102(41), 1466514670.Google Scholar
Amano, T., Sagai, H., Tanabe, Y., et al. (2009). Chromosomal dynamics at the Shh locus: limb bud-specific differential regulation of competence and active transcription. Developmental Cell, 16(1), 4757.Google Scholar
Auton, A. and McVean, G. (2012). Estimating recombination rates from genetic variation in humans. Methods in Molecular Biology, 856, 217237.Google Scholar
Aziz, A., Liu, Q. C. and Dilworth, F. J. (2010). Regulating a master regulator: establishing tissue-specific gene expression in skeletal muscle. Epigenetics, 5(8), 691695.CrossRefGoogle ScholarPubMed
Barrett, R. D. and Hoekstra, H. E. (2011). Molecular spandrels: tests of adaptation at the genetic level. Nature Reviews Genetics, 12(11), 767780.CrossRefGoogle ScholarPubMed
Bataillon, T., Duan, J., Hvilsom, C., et al. (2015). Inference of purifying and positive selection in three subspecies of chimpanzees (Pan troglodytes) from exome sequencing. Genome Biology and Evolution, 7(4), 11221132.Google Scholar
Behringer, R, Vintersten Nagy, G. M. and Nagy, K. (2014). Manipulating the Mouse Embryo: A Laboratory Manual, Fourth Edition. New York, NY: Cold Spring Harbor Press.Google Scholar
Benito-Sanz, S., Aza-Carmona, M., Rodriguez-Estevez, A., et al. (2012). Identification of the first PAR1 deletion encompassing upstream SHOX enhancers in a family with idiopathic short stature. European Journal of Human Genetics, 20(1), 125127.CrossRefGoogle Scholar
Bernstein, B. E., Kamal, M., Lindblad-Toh, K., et al. (2005). Genomic maps and comparative analysis of histone modifications in human and mouse. Cell, 120(2), 169181.Google Scholar
Bernstein, B. E., Mikkelsen, T. S., Xie, X., et al. (2006). A bivalent chromatin structure marks key developmental genes in embryonic stem cells. Cell, 125(2), 315326.Google Scholar
Bi, W., Deng, J. M., Zhang, Z., Behringer, R. R. and de Crombrugghe, B. (1999). Sox9 is required for cartilage formation. Nature Genetics, 22(1), 8589.Google Scholar
Bowen, M. E., Ayturk, U. M., Kurek, K. C., Yang, W. and Warman, M. L. (2014). SHP2 regulates chondrocyte terminal differentiation, growth plate architecture and skeletal cell fates. PLoS Genetics, 10(5), e1004364.Google Scholar
Buchwalow, I. B. and Böcker, W. (2014). Immunohistochemistry: Basics and Methods. New York, NY: Springer.Google Scholar
Buenrostro, J. D., Giresi, P. G., Zaba, L. C., Chang, H. Y. and Greenleaf, W. J. (2013). Transposition of native chromatin for fast and sensitive epigenomic profiling of open chromatin, DNA-binding proteins and nucleosome position. Nature Methods, 10(12), 12131218.Google Scholar
Campeau, E. and Gobeil, S. (2011). RNA interference in mammals: behind the screen. Briefings in Functional Genomics, 10(4), 215226.CrossRefGoogle ScholarPubMed
Canavez, F., Young, N. T., Guethlein, L. A., et al. (2001). Comparison of chimpanzee and human leukocyte Ig-like receptor genes reveals framework and rapidly evolving genes. Journal of Immunology, 167(10), 57865794.Google Scholar
Capellini, T. D., Di Giacomo, G., Salsi, V., et al. (2006). Pbx1/Pbx2 requirement for distal limb patterning is mediated by the hierarchical control of Hox gene spatial distribution and Shh expression. Development, 133(11), 22632273.Google Scholar
Capellini, T. D., Vaccari, G., Ferretti, E., et al. (2010). Scapula development is governed by genetic interactions of Pbx1 with its family members and with Emx2 via their cooperative control of Alx1. Development, 137(15), 25592569.CrossRefGoogle ScholarPubMed
Carroll, S. B. (2008). Evo–devo and an expanding evolutionary synthesis: a genetic theory of morphological evolution. Cell, 134(1), 2536.Google Scholar
Chan, Y. F., Marks, M. E., Jones, F. C., et al. (2010). Adaptive evolution of pelvic reduction in sticklebacks by recurrent deletion of a Pitx1 enhancer. Science, 327(5963), 302305.CrossRefGoogle ScholarPubMed
Chau, M., Lui, J. C., Landman, E. B., et al. (2014). Gene expression profiling reveals similarities between the spatial architectures of postnatal articular and growth plate cartilage. PLoS ONE, 9(7), e103061.CrossRefGoogle ScholarPubMed
Chen, Y., Zheng, Y., Kang, Y., et al. (2015). Functional disruption of the dystrophin gene in rhesus monkey using CRISPR/Cas9. Human Molecular Genetics, 24(13), 37643774.Google Scholar
Cheng, Y., Ma, Z., Kim, B. H., et al. (2014). Principles of regulatory information conservation between mouse and human. Nature, 515(7527), 371375.Google Scholar
Chiang, C., Litingtung, Y., Lee, E., et al. (1996). Cyclopia and defective axial patterning in mice lacking Sonic hedgehog gene function. Nature, 383(6599), 407413.CrossRefGoogle ScholarPubMed
Chimpanzee Consortium Analysis. (2005). Initial sequence of the chimpanzee genome and comparison with the human genome. Nature, 437(7055), 6987.Google Scholar
Christian, M., Cermak, T., Doyle, E. L., et al. (2010). Targeting DNA double-strand breaks with TAL effector nucleases. Genetics, 186(2), 757761.Google Scholar
Christians, J. K., de Zwaan, D. R. and Fung, S. H. (2013). Pregnancy associated plasma protein A2 (PAPP-A2) affects bone size and shape and contributes to natural variation in postnatal growth in mice. PLoS ONE, 8(2), e56260.Google Scholar
Church, V., Nohno, T., Linker, C., Marcelle, C. and Francis-West, P. (2002). Wnt regulation of chondrocyte differentiation. Journal of Cell Science, 115(Pt 24), 48094818.Google Scholar
Codina-Sola, M., Rodriguez-Santiago, B., Homs, A., et al. (2015). Integrated analysis of whole-exome sequencing and transcriptome profiling in males with autism spectrum disorders. Molecular Autism, 6, 21.CrossRefGoogle ScholarPubMed
Colosimo, P. F., Hosemann, K. E., Balabhadra, S. G., et al. (2005). Widespread parallel evolution in sticklebacks by repeated fixation of Ectodysplasin alleles. Science, 307(5717), 19281933.Google Scholar
Cooper, K. L., Oh, S., Sung, Y., et al. (2013). Multiple phases of chondrocyte enlargement underlie differences in skeletal proportions. Nature, 495(7441), 375378.Google Scholar
Cotney, J., Leng, J., Yin, J., et al. (2013). The evolution of lineage-specific regulatory activities in the human embryonic limb. Cell, 154(1), 185196.CrossRefGoogle ScholarPubMed
Crawford, G. E., Holt, I. E., Mullikin, J. C., et al. (2004). Identifying gene regulatory elements by genome-wide recovery of DNase hypersensitive sites. Proceedings of the National Academy of Sciences of the USA, 101(4), 992997.Google Scholar
Creyghton, M. P., Cheng, A. W., Welstead, G. G., et al. (2010). Histone H3K27ac separates active from poised enhancers and predicts developmental state. Proceedings of the National Academy of Sciences of the USA, 107(50), 2193121936.Google Scholar
Datta, S., Malhotra, L., Dickerson, R., et al. (2015). Laser capture microdissection: big data from small samples. Histology and Histopathology, 30(11), 12551269.Google Scholar
Davidson, E. (2001). Genomic Regulatory Systems: Development and Evolution. San Diego, CA: Academic Press.Google Scholar
Decker, R. S., Koyama, E., Enomoto-Iwamoto, M., et al. (2014). Mouse limb skeletal growth and synovial joint development are coordinately enhanced by Kartogenin. Developmental Biology, 395(2), 255267.Google Scholar
Dekker, J., Rippe, K., Dekker, M. and Kleckner, N. (2002). Capturing chromosome conformation. Science, 295(5558), 13061311.Google Scholar
Dong, Z. and Chen, Y. (2013). Transcriptomics: advances and approaches. Science in China Series C Life Sciences, 56(10), 960967.Google Scholar
Dostie, J., Richmond, T. A., Arnaout, R. A., et al. (2006). Chromosome Conformation Capture Carbon Copy (5C): a massively parallel solution for mapping interactions between genomic elements. Genome Research, 16(10), 12991309.Google Scholar
Dudley, A. T., Ros, M. A. and Tabin, C. J. (2002). A re-examination of proximodistal patterning during vertebrate limb development. Nature, 418(6897), 539544.Google Scholar
ENCODE Project Consortium. (2012). An integrated encyclopedia of DNA elements in the human genome. Nature, 489(7414), 5774.Google Scholar
Farber, C. R., Kelly, S. A., Baruch, E., et al. (2011). Identification of quantitative trait loci influencing skeletal architecture in mice: emergence of Cdh11 as a primary candidate gene regulating femoral morphology. Journal of Bone and Mineral Research, 26(9), 21742183.CrossRefGoogle ScholarPubMed
Farooq, M., Nakai, H., Fujimoto, A., et al. (2013). Characterization of a novel missense mutation in the prodomain of GDF5, which underlies brachydactyly type C and mild Grebe type chondrodysplasia in a large Pakistani family. Human Genetics, 132(11), 12531264.CrossRefGoogle Scholar
Florio, M., Albert, M., Taverna, E., et al. (2015). Human-specific gene ARHGAP11B promotes basal progenitor amplification and neocortex expansion. Science, 347(6229), 14651470.CrossRefGoogle ScholarPubMed
Fossat, N., Ip, C. K., Jones, V. J., et al. (2015). Context-specific function of the LIM homeobox 1 transcription factor in head formation of the mouse embryo. Development, 142(11), 20692079.Google Scholar
Francis-West, P. H., Richardson, M. K., Bell, E., et al. (1996). The effect of overexpression of BMPs and GDF-5 on the development of chick limb skeletal elements. Annals if the New York Academy of Science, 785, 254255.CrossRefGoogle ScholarPubMed
Fukami, M., Tsuchiya, T., Takada, S., et al. (2012). Complex genomic rearrangement in the SOX9 5' region in a patient with Pierre Robin sequence and hypoplastic left scapula. American Journal of Medical Genetics A, 158A(7), 15291534.Google Scholar
Fullwood, M. J., Han, Y., Wei, C. L., Ruan, X. and Ruan, Y. (2010). Chromatin interaction analysis using paired-end tag sequencing. Current Protocol in Molecular Biology, Chapter 21: Unit 21.1521.25.Google Scholar
Furey, T. S. (2012). ChIP-seq and beyond: new and improved methodologies to detect and characterize protein–DNA interactions. Nature Reviews Genetics, 13(12), 840852.Google Scholar
Furniss, D., Lettice, L. A., Taylor, I. B., et al. (2008). A variant in the Sonic hedgehog regulatory sequence (ZRS) is associated with triphalangeal thumb and deregulates expression in the developing limb. Human Molecular Genetics, 17(16), 24172423.Google Scholar
Gallego Romero, I., Pavlovic, B. J., Hernando-Herraez, I., et al. (2015). A panel of induced pluripotent stem cells from chimpanzees: a resource for comparative functional genomics. Elife 4.Google Scholar
Gehrke, A. R., Schneider, I., de la Calle-Mustienes, E., et al. (2015). Deep conservation of wrist and digit enhancers in fish. Proceedings of the National Academy of Sciences of the USA, 112(3), 803808.Google Scholar
Gennequin, B., Otte, D. M. and Zimmer, A. (2013). CRISPR/Cas-induced double-strand breaks boost the frequency of gene replacements for humanizing the mouse Cnr2 gene. Biochemical and Biophysical Research Commununications, 441(4): 815819.Google Scholar
Gibson, S. V. and Muse, G. (2009). A Primer in Genome Sciences. New York, NY: Sinauer Associates.Google Scholar
Gilad, Y., Rifkin, S. A. and Pritchard, J. K. (2008). Revealing the architecture of gene regulation: the promise of eQTL studies. Trends in Genetics, 24(8), 408415.Google Scholar
Gilbert, S. (2013). Developmental Biology. New York, NY: Sinauer Associates.Google Scholar
Giresi, P. G., Kim, J., McDaniell, R. M., Iyer, V. R. and Lieb, J. D. (2007). FAIRE (formaldehyde-assisted isolation of regulatory elements) isolates active regulatory elements from human chromatin. Genome Research, 17(6), 877885.Google Scholar
Gordon, C. T., Rodda, F. A. and Farlie, P. G. (2009a). The RCAS retroviral expression system in the study of skeletal development. Developmental Dynamics, 238(4), 797811.Google Scholar
Gordon, C. T., Tan, T. Y., Benko, S., et al. (2009b). Long-range regulation at the SOX9 locus in development and disease. Journal of Medical Genetics, 46(10), 649656.Google Scholar
Gross, D. S. and Garrard, W. T. (1988). Nuclease hypersensitive sites in chromatin. Annual Review of Biochemistry, 57, 159197.Google Scholar
Grossman, S. R., Andersen, K. G., Shlyakhter, I., et al. (2013). Identifying recent adaptations in large-scale genomic data. Cell, 152(4), 703713.Google Scholar
Guenther, C., Pantalena-Filho, L. and Kingsley, D. M. (2008). Shaping skeletal growth by modular regulatory elements in the Bmp5 gene. PLoS Genetics, 4(12), e1000308.Google Scholar
Gupta, R. M. and Musunuru, K. (2014). Expanding the genetic editing tool kit: ZFNs, TALENs, and CRISPR-Cas9. Journal of Clinical Investigations, 124(10), 41544161.Google Scholar
Gurnett, C. A., Bowcock, A. M., Dietz, F. R., et al. (2007). Two novel point mutations in the long-range SHH enhancer in three families with triphalangeal thumb and preaxial polydactyly. American Journal of Medical Genetics A, 143A(1) , 2732.Google Scholar
Harfe, B. D., Scherz, P. J., Nissim, S., et al. (2004). Evidence for an expansion-based temporal Shh gradient in specifying vertebrate digit identities. Cell, 118(4), 517528.Google Scholar
Hartl, D. and Ruvolo, M. (2011). Genetics: Analysis of Genes and Genomes. New York, NY: Jones & Bartlett Learning.Google Scholar
Hartmann, C. and Tabin, C. J. (2000). Dual roles of Wnt signaling during chondrogenesis in the chicken limb. Development, 127(14), 31413159.Google Scholar
Hauptmann, G. (2015). In Situ Hybridization Methods (Neuromethods). New York, NY: Humana Press.Google Scholar
Havill, L. M., Mahaney, M. C., Cox, L. A., et al. (2005). A quantitative trait locus for normal variation in forearm bone mineral density in pedigreed baboons maps to the ortholog of human chromosome 11q. Journal of Clinical Endocrinology and Metabolism, 90(6), 36383645.Google Scholar
Heintzman, N. D., Hon, G. C., Hawkins, R. D., et al. (2009). Histone modifications at human enhancers reflect global cell-type-specific gene expression. Nature, 459(7243), 108112.Google Scholar
Hernando-Herraez, I., Prado-Martinez, J., Garg, P., et al. (2013). Dynamics of DNA methylation in recent human and great ape evolution. PLoS Genetics, 9(9), e1003763.Google Scholar
Hiller, M., Schaar, B. T., Indjeian, V. B., et al. (2012). A forward genomics approach links genotype to phenotype using independent phenotypic losses among related species. Cell Reports, 2(4), 817823.Google Scholar
Horvat-Gordon, M., Praul, C. A., Ramachandran, R., Bartell, P. A. and Leach, R. M. Jr. (2010). Use of microarray analysis to study gene expression in the avian epiphyseal growth plate. Comparative Biochemistry and Physiology Part D Genomics and Proteomics, 5(1), 1223.Google Scholar
Huang, J., Chen, J., Esparza, J., et al. (2015). eQTL mapping identifies insertion- and deletion-specific eQTLs in multiple tissues. Nature Communications, 6, 6821.Google Scholar
Huang, R., Christ, B. and Patel, K. (2006). Regulation of scapula development. Anatomy and Embryology, 211(Suppl 1), 6571.Google Scholar
Infante, C. R., Park, S., Mihala, A. G., Kingsley, D. M. and Menke, D. B. (2013). Pitx1 broadly associates with limb enhancers and is enriched on hindlimb cis-regulatory elements. Developmental Biology, 374(1), 234244.Google Scholar
Inoue, K. and Imai, Y. (2014). Identification of novel transcription factors in osteoclast differentiation using genome-wide analysis of open chromatin determined by DNase-seq. Journal of Bone and Mineral Research, 29(8), 18231832.Google Scholar
James, C. G., Stanton, L. A., Agoston, H., et al. (2010). Genome-wide analyses of gene expression during mouse endochondral ossification. PLoS ONE, 5(1), e8693.Google Scholar
Jasinska, A. J., Lin, M. K., Service, S., et al. (2012). A non-human primate system for large-scale genetic studies of complex traits. Human Molecular Genetics, 21(15), 33073316.Google Scholar
Jinek, M., Chylinski, K., Fonfara, I., et al. (2012). A programmable dual-RNA-guided DNA endonuclease in adaptive bacterial immunity. Science, 337(6096), 816821.Google Scholar
Jones, F. C., Grabherr, M. G., Chan, Y. F., et al. (2012). The genomic basis of adaptive evolution in threespine sticklebacks. Nature, 484(7392), 5561.Google Scholar
Kamberov, Y. G., Wang, S., Tan, J., et al. (2013). Modeling recent human evolution in mice by expression of a selected EDAR variant. Cell, 152(4), 691702.Google Scholar
Kenney-Hunt, J. P., Vaughn, T. T., Pletscher, L. S., et al. (2006). Quantitative trait loci for body size components in mice. Mammalian Genome, 17(6), 526537.Google Scholar
Khaitovich, P., Hellmann, I., Enard, W., et al. (2005). Parallel patterns of evolution in the genomes and transcriptomes of humans and chimpanzees. Science, 309(5742), 18501854.Google Scholar
Khaitovich, P., Muetzel, B., She, X., et al. (2004). Regional patterns of gene expression in human and chimpanzee brains. Genome Research, 14(8), 14621473.Google Scholar
Khan, Z., Ford, M. J., Cusanovich, D. A., et al. (2013). Primate transcript and protein expression levels evolve under compensatory selection pressures. Science, 342(6162), 11001104.Google Scholar
King, M. C. and Wilson, A. C. (1975). Evolution at two levels in humans and chimpanzees. Science, 188(4184), 107116.Google Scholar
Koyama, E., Shibukawa, Y., Nagayama, M., et al. (2008). A distinct cohort of progenitor cells participates in synovial joint and articular cartilage formation during mouse limb skeletogenesis. Developmental Biology, 316(1), 6273.Google Scholar
Kozhemyakina, E., Lassar, A. B. and Zelzer, E. (2015). A pathway to bone: signaling molecules and transcription factors involved in chondrocyte development and maturation. Development, 142(5), 817831.Google Scholar
Kraft, K., Geuer, S., Will, A. J., et al. (2015). Deletions, inversions, duplications: engineering of structural variants using CRISPR/Cas in mice. Cell Reports.Google Scholar
Kraus, P., Xing, S. V. and Lufkin, T. (2014). Generating mouse lines for lineage tracing and knockout studies. In: Singh, S. R. and Coppola, V. (eds.) Mouse Genetics: Methods and Protocols, Methods in Molecular Biology. New York, NY: Springer Science+Business Media.Google Scholar
Krebs, J. E., Gi, E. and Kilpatrick, S. T. (2014). Lewin’s Genes XI. Burlington, MA: Jones & Bartlett Learning.Google Scholar
Lanctôt, C., Moreau, A., Chamberland, M., Tremblay, M. L. and Drouin, J. (1999). Hindlimb patterning and mandible development require the Ptx1 gene. Development, 126(9), 18051810.Google Scholar
Lappalainen, T., Sammeth, M., Friedlander, M. R., et al. (2013). Transcriptome and genome sequencing uncovers functional variation in humans. Nature, 501(7468), 506511.Google Scholar
Lemay, P., Guyot, M. C., Tremblay, E., et al. (2015). Loss-of-function de novo mutations play an important role in severe human neural tube defects. Journal of Medical Genetics, 52(7), 493497.Google Scholar
Lettice, L. A., Heaney, S. J., Purdie, L. A., et al. (2003). A long-range Shh enhancer regulates expression in the developing limb and fin and is associated with preaxial polydactyly. Human Molecular Genetics, 12(14), 17251735.Google Scholar
Lettice, L. A., Hill, A. E., Devenney, P. S. and Hill, R. E. (2008). Point mutations in a distant sonic hedgehog cis-regulator generate a variable regulatory output responsible for preaxial polydactyly. Human Molecular Genetics, 17(7), 978985.Google Scholar
Li, Y. R. and Keating, B. J. (2014). Trans-ethnic genome-wide association studies: advantages and challenges of mapping in diverse populations. Genome Medicine, 6(10), 91.CrossRefGoogle ScholarPubMed
Liang, P., Xu, Y., Zhang, X., et al. (2015). CRISPR/Cas9-mediated gene editing in human tripronuclear zygotes. Protein and Cell, 6(5), 363372.Google Scholar
Lieberman-Aiden, E., van Berkum, N. L., Williams, L., et al. (2009). Comprehensive mapping of long-range interactions reveals folding principles of the human genome. Science, 326(5950), 289293.Google Scholar
Lizio, M., Harshbarger, J., Shimoji, H., et al. (2015). Gateways to the FANTOM5 promoter level mammalian expression atlas. Genome Biology, 16, 22.Google Scholar
Logan, M. and Tabin, C. J. (1999). Role of Pitx1 upstream of Tbx4 in specification of hindlimb identity. Science, 283(5408), 17361739.Google Scholar
Lui, J. C., Andrade, A. C., Forcinito, P., et al. (2010). Spatial and temporal regulation of gene expression in the mammalian growth plate. Bone, 46(5), 13801390.Google Scholar
Lupiáñez, D. G., Kraft, K., Heinrich, V., et al. (2015). Disruptions of topological chromatin domains cause pathogenic rewiring of gene-enhancer interactions. Cell, 161(5), 10121025.Google Scholar
MacCabe, A. B., Gasseling, M. T. and Saunders Jr, J. W. (1973). Spatiotemporal distribution of mechanisms that control outgrowth and anteroposterior polarization of the limb bud in the chick embryo. Mechanisms of Ageing and Development, 2(1), 112.Google Scholar
Macosko, E. Z., Basu, A., Satija, R., et al. (2015). Highly parallel genome-wide expression profiling of individual cells using nanoliter droplets. Cell, 161(5), 12021214.Google Scholar
Madrigal, P. and Krajewski, P. (2012). Current bioinformatic approaches to identify DNase I hypersensitive sites and genomic footprints from DNase-seq data. Frontiers in Genetics, 3, 230.Google Scholar
Majewski, J. and Pastinen, T. (2011). The study of eQTL variations by RNA-seq: from SNPs to phenotypes. Trends in Genetics, 27(2), 7279.Google Scholar
Marinić, M., Aktas, T., Ruf, S. and Spitz, F. (2013). An integrated holo-enhancer unit defines tissue and gene specificity of the Fgf8 regulatory landscape. Developmental Cell, 24(5), 530542.Google Scholar
McLean, C. Y., Reno, P. L., Pollen, A. A., et al. (2011). Human-specific loss of regulatory DNA and the evolution of human-specific traits. Nature, 471(7337), 216219.Google Scholar
Menke, D. B. (2013). Engineering subtle targeted mutations into the mouse genome. Genesis, 51(9), 605618.CrossRefGoogle ScholarPubMed
Merino, R., Macias, D., Ganan, Y., et al. (1999). Expression and function of Gdf-5 during digit skeletogenesis in the embryonic chick leg bud. Developmental Biology, 206(1), 3345.CrossRefGoogle ScholarPubMed
Meyer, C. A. and Liu, X. S. (2014). Identifying and mitigating bias in next-generation sequencing methods for chromatin biology. Nature Reviews Genetics, 15(11), 709721.Google Scholar
Mori, Y., Chung, U. I., Tanaka, S. and Saito, T. (2014). Determination of differential gene expression profiles in superficial and deeper zones of mature rat articular cartilage using RNA sequencing of laser microdissected tissue specimens. Biomedical Research, 35(4), 263270.Google Scholar
Mortlock, D. P., Guenther, C. and Kingsley, D. M. (2003). A general approach for identifying distant regulatory elements applied to the Gdf6 gene. Genome Research, 13(9), 20692081.Google Scholar
Nagai, H. and Aoki, M. (2002). Inhibition of growth plate angiogenesis and endochondral ossification with diminished expression of MMP-13 in hypertrophic chondrocytes in FGF-2-treated rats. Journal of Bone and Mineral Metabolism, 20(3), 142147.Google Scholar
Nam, K., Munch, K., Hobolth, A., et al. (2015). Extreme selective sweeps independently targeted the X chromosomes of the great apes. Proceedings of the National Academy of Sciences of the USA, 112(20), 64136418.Google Scholar
Nikolskiy, I., Conrad, D. F., Chun, S., et al. (2015). Using whole-genome sequences of the LG/J and SM/J inbred mouse strains to prioritize quantitative trait genes and nucleotides. BMC Genomics, 16, 415.Google Scholar
Nilsson, O., Guo, M. H., Dunbar, N., et al. (2014). Short stature, accelerated bone maturation, and early growth cessation due to heterozygous aggrecan mutations. Journal of Clinical Endocrinology and Metabolism, 99(8), E1510–1518.Google Scholar
Niu, Y., Shen, B., Cui, Y., et al. (2014). Generation of gene-modified cynomolgus monkey via Cas9/RNA-mediated gene targeting in one-cell embryos. Cell, 156(4), 836843.Google Scholar
Oh, C. D., Lu, Y., Liang, S., et al. (2014). SOX9 regulates multiple genes in chondrocytes, including genes encoding ECM proteins, ECM modification enzymes, receptors, and transporters. PLoS ONE, 9(9), e107577.Google Scholar
Ohba, S., He, X., Hojo, H. and McMahon, A. P. (2015). Distinct transcriptional programs underlie Sox9 regulation of the mammalian chondrocyte. Cell Report, 12(2), 229243.Google Scholar
1000 Genomes Project Consortium. (2012). An integrated map of genetic variation from 1,092 human genomes. Nature, 491(7422), 5665.Google Scholar
Orlando, V. (2000). Mapping chromosomal proteins in vivo by formaldehyde-crosslinked-chromatin immunoprecipitation. Trends in Biochemical Science, 25(3), 99104.Google Scholar
Ozsolak, F. and Milos, P. M. (2011). RNA sequencing: advances, challenges and opportunities. Nature Reviews Genetics, 12(2), 8798.Google Scholar
Pacifici, M., Koyama, E., Shibukawa, Y., et al. (2006). Cellular and molecular mechanisms of synovial joint and articular cartilage formation. Annals of the New York Academy of Science, 1068, 7486.Google Scholar
Pardo-Diaz, C., Salazar, C. and Jiggins, C. D. (2015). Towards the identification of the loci of adaptive evolution. Methods in Ecology and Evolution, 6(4), 445464.Google Scholar
Parry, D. A., Logan, C. V., Stegmann, A. P., et al. (2013). SAMS, a syndrome of short stature, auditory-canal atresia, mandibular hypoplasia, and skeletal abnormalities is a unique neurocristopathy caused by mutations in Goosecoid. American Journal of Human Genetics, 93(6), 11351142.Google Scholar
Pazin, D. E., Gamer, L. W., Cox, K. A. and Rosen, V. (2012). Molecular profiling of synovial joints: use of microarray analysis to identify factors that direct the development of the knee and elbow. Developmental Dynamics, 241(11), 18161826.Google Scholar
Perry, G. H., Melsted, P., Marioni, J. C., et al. (2012). Comparative RNA sequencing reveals substantial genetic variation in endangered primates. Genome Research, 22(4), 602610.Google Scholar
Pokholok, D. K., Harbison, C. T., Levine, S., et al. (2005). Genome-wide map of nucleosome acetylation and methylation in yeast. Cell, 122(4), 517527.Google Scholar
Porcu, E., Sanna, S., Fuchsberger, C. and Fritsche, L. G. (2013). Genotype imputation in genome-wide association studies. Current Protocol in Human Genetics, Chapter 1: Unit 1.25.Google Scholar
Prabhakar, S., Noonan, J.P., Paabo, S. and Rubin, E.M. (2006). Accelerated evolution of conserved noncoding sequences in humans. Science, 314(5800), 786.Google Scholar
Prabhakar, S., Visel, A., Akiyama, J.A., et al. (2008). Human-specific gain of function in a developmental enhancer. Science, 321(5894), 13461350.Google Scholar
Prado-Martinez, J., Sudmant, P. H., Kidd, J. M., et al. (2013). Great ape genetic diversity and population history. Nature, 499(7459), 471475.Google Scholar
Provot, S., Kempf, H., Murtaugh, L. C., et al. (2006). Nkx3.2/Bapx1 acts as a negative regulator of chondrocyte maturation. Development, 133(4), 651662.Google Scholar
Prüfer, K., Munch, K., Hellmann, I., et al. (2012). The bonobo genome compared with the chimpanzee and human genomes. Nature, 486(7404), 527531.Google Scholar
Rada-Iglesias, A., Bajpai, R., Swigut, T., et al. (2011). A unique chromatin signature uncovers early developmental enhancers in humans. Nature, 470(7333), 279283.Google Scholar
Reilly, S. K., Yin, J., Ayoub, A. E., et al. (2015). Evolutionary genomics. Evolutionary changes in promoter and enhancer activity during human corticogenesis. Science, 347(6226), 11551159.Google Scholar
Riddle, R. D., Johnson, R. L., Laufer, E. and Tabin, C. (1993). Sonic hedgehog mediates the polarizing activity of the ZPA. Cell, 75(7), 14011416.Google Scholar
Rivera, C. M. and Ren, B. (2013). Mapping human epigenomes. Cell, 155(1), 3955.Google Scholar
Roadmap Epigenomics Project Consortium. (2015). Integrative analysis of 111 reference human epigenomes. Nature, 518(7539), 317330.Google Scholar
Rock, J. R., Lopez, M. C., Baker, H. V. and Harfe, B. D. (2007). Identification of genes expressed in the mouse limb using a novel ZPA microarray approach. Gene Expression Patterns, 8(1), 1926.Google Scholar
Rodriguez-Esteban, C., Tsukui, T., Yonei, S., et al. (1999). The T-box genes Tbx4 and Tbx5 regulate limb outgrowth and identity. Nature, 398(6730), 814818.Google Scholar
Roux, J., Rosikiewicz, M. and Robinson-Rechavi, M. (2015). What to compare and how: comparative transcriptomics for Evo–Devo. Journal of Experimental Zoology Series B Molecular Development and Evolution, 324(4), 372382.Google Scholar
Sabo, P. J., Humbert, R., Hawrylycz, M., et al. (2004). Genome-wide identification of DNaseI hypersensitive sites using active chromatin sequence libraries. Proceedings of the National Academy of Science of the USA, 101(13), 45374542.Google Scholar
Sagai, T., Hosoya, M., Mizushina, Y., Tamura, M. and Shiroishi, T. (2005). Elimination of a long-range cis-regulatory module causes complete loss of limb-specific Shh expression and truncation of the mouse limb. Development, 132(4), 797803.Google Scholar
Scally, A., Dutheil, J. Y., Hillier, L. W., et al. (2012). Insights into hominid evolution from the gorilla genome sequence. Nature, 483(7388), 169175.Google Scholar
Scherz, P. J., McGlinn, E., Nissim, S. and Tabin, C. J. (2007). Extended exposure to Sonic hedgehog is required for patterning the posterior digits of the vertebrate limb. Developmental Biology, 308(2), 343354.CrossRefGoogle ScholarPubMed
Schreiner, C. M., Bell, S. M. and ScottJr., W. J. (2009). Microarray analysis of murine limb bud ectoderm and mesoderm after exposure to cadmium or acetazolamide. Birth Defects Research Series A Clinical and Molecular Teratology, 85(7), 588598.Google Scholar
Shao, Y., Guan, Y., Wang, L., et al. (2014). CRISPR/Cas-mediated genome editing in the rat via direct injection of one-cell embryos. Nature Protocols, 9(10), 24932512.Google Scholar
Shapiro, M. D., Marks, M. E., Peichel, C. L., et al. (2004). Genetic and developmental basis of evolutionary pelvic reduction in threespine sticklebacks. Nature, 428(6984), 717723.Google Scholar
Sherwood, R. J., Duren, D. L., Havill, L. M., et al. (2008). A genomewide linkage scan for quantitative trait loci influencing the craniofacial complex in baboons (Papio hamadryas spp.). Genetics, 180(1), 619628.CrossRefGoogle ScholarPubMed
Shou, S., Scott, V., Reed, C., Hitzemann, R. and Stadler, H. S. (2005). Transcriptome analysis of the murine forelimb and hindlimb autopod. Developmental Dynamics, 234(1), 7489.Google Scholar
Soriano, P. (1999). Generalized lacZ expression with the ROSA26 Cre reporter strain. Nature Genetics, 21(1), 7071.Google Scholar
Stephens, T. D. and McNulty, T. R. (1981). Evidence for a metameric pattern in the development of the chick humerus. Journal of Embryology and Experimental Morphology, 61, 191205.Google Scholar
Stergachis, A. B., Neph, S., Sandstrom, R., et al. (2014). Conservation of trans-acting circuitry during mammalian regulatory evolution. Nature, 515(7527), 365370.Google Scholar
Sudmant, P. H., Huddleston, J., Catacchio, C. R., et al. (2013). Evolution and diversity of copy number variation in the great ape lineage. Genome Research, 23(9), 13731382.Google Scholar
Summerbell, D. (1979). The zone of polarizing activity: evidence for a role in normal chick limb morphogenesis. Journal of Embryology and Experimental Morphology, 50, 217233.Google Scholar
Summerbell, D., Ashby, P. R., Coutelle, O., et al. (2000). The expression of Myf5 in the developing mouse embryo is controlled by discrete and dispersed enhancers specific for particular populations of skeletal muscle precursors. Development, 127(17), 37453757.Google Scholar
Sun, M., Ma, F., Zeng, X., et al. (2008). Triphalangeal thumb-polysyndactyly syndrome and syndactyly type IV are caused by genomic duplications involving the long range, limb-specific SHH enhancer. Journal of Medical Genetics, 45(9), 589595.Google Scholar
Tai, P. W., Wu, H., Gordon, J. A., et al. (2014). Epigenetic landscape during osteoblastogenesis defines a differentiation-dependent Runx2 promoter region. Gene, 550(1), 19.Google Scholar
Takeuchi, J. K., Koshiba-Takeuchi, K., Matsumoto, K., et al. (1999). Tbx5 and Tbx4 genes determine the wing/leg identity of limb buds. Nature, 398(6730), 810814.Google Scholar
Tetreault, M., Bareke, E., Nadaf, J., Alirezaie, N. and Majewski, J. (2015). Whole-exome sequencing as a diagnostic tool: current challenges and future opportunities. Expert Review of Molecular Diagnostics, 15(6), 749760.Google Scholar
Tickle, C., Shellswell, G., Crawley, A. and Wolpert, L. (1976). Positional signalling by mouse limb polarising region in the chick wing bud. Nature, 259(5542), 396397.Google Scholar
Tung, J. W., Heydari, K., Tirouvanziam, R., et al. (2007). Modern flow cytometry: a practical approach. Clinical Laboratory Medicine, 27(3), 453468, v.Google Scholar
Urnov, F. D., Rebar, E. J., Holmes, M. C., Zhang, H. S. and Gregory, P. D. (2010). Genome editing with engineered zinc finger nucleases. Nature Reviews Genetics, 11(9), 636646.Google Scholar
Valasek, P., Evans, D. J., Maina, F., Grim, M. and Patel, K. (2005). A dual fate of the hindlimb muscle mass: cloacal/perineal musculature develops from leg muscle cells. Development, 132(3), 447458.Google Scholar
Vargesson, N., Clarke, J. D., Vincent, K., et al. (1997). Cell fate in the chick limb bud and relationship to gene expression. Development, 124(10), 19091918.Google Scholar
Vitti, J. J., Grossman, S. R. and Sabeti, P. C. (2013). Detecting natural selection in genomic data. Annual Review of Genetics, 47, 97120.Google Scholar
Vortkamp, A., Lee, K., Lanske, B., et al. (1996). Regulation of rate of cartilage differentiation by Indian hedgehog and PTH-related protein. Science, 273(5275), 613622.CrossRefGoogle ScholarPubMed
Wan, H., Feng, C., Teng, F., et al. (2015). One-step generation of p53 gene biallelic mutant Cynomolgus monkey via the CRISPR/Cas system. Cell Research, 25(2), 258261.Google Scholar
Wang, Y., Middleton, F., Horton, J. A., et al. (2004). Microarray analysis of proliferative and hypertrophic growth plate zones identifies differentiation markers and signal pathways. Bone, 35(6), 12731293.Google Scholar
Wang, H., Yang, H., Shivalila, C. S., et al. (2013). One-step generation of mice carrying mutations in multiple genes by CRISPR/Cas-mediated genome engineering. Cell, 153(4), 910918.Google Scholar
Weir, B. S., Cardon, L. R., Anderson, A. D., Nielsen, D. M. and Hill, W. G. (2005). Measures of human population structure show heterogeneity among genomic regions. Genome Research, 15(11), 14681476.Google Scholar
Widdig, A., Kessler, M. J., Bercovitch, F. B., et al. (2016). Genetic studies on the Cayo Santiago rhesus macaques: a review of 40 years of research. American Journal of Primatology, 78(1), 4462.Google Scholar
Wood, A. R., Esko, T., Yang, J., et al. (2014). Defining the role of common variation in the genomic and biological architecture of adult human height. Nature Genetics, 46(11), 11731186.Google Scholar
Wu, C., Bingham, P. M., Livak, K. J., Holmgren, R. and Elgin, S. C. (1979a). The chromatin structure of specific genes: I. Evidence for higher order domains of defined DNA sequence. Cell, 16(4), 797806.Google Scholar
Wu, C., Wong, Y. C. and Elgin, S. C. (1979b). The chromatin structure of specific genes: II. Disruption of chromatin structure during gene activity. Cell, 16(4), 807814.Google Scholar
Wu, S. and De Luca, F. (2006). Inhibition of the proteasomal function in chondrocytes down-regulates growth plate chondrogenesis and longitudinal bone growth. Endocrinology, 147(8), 37613768.Google Scholar
Xue, Y., Prado-Martinez, J., Sudmant, P. H., et al. (2015). Mountain gorilla genomes reveal the impact of long-term population decline and inbreeding. Science, 348(6231), 242245.Google Scholar
Yang, Y., Drossopoulou, G., Chuang, P. T., et al. (1997). Relationship between dose, distance and time in Sonic Hedgehog-mediated regulation of anteroposterior polarity in the chick limb. Development, 124(21), 43934404.Google Scholar
Young, N. M. (2006). Function, ontogeny and canalization of shape variance in the primate scapula. Journal of Anatomy, 209(5), 623636.Google Scholar
Yue, F., Cheng, Y., Breschi, A., et al. (2014). A comparative encyclopedia of DNA elements in the mouse genome. Nature, 515(7527), 355364.Google Scholar
Zhang, F., Wen, Y. and Guo, X. (2014). CRISPR/Cas9 for genome editing: progress, implications and challenges. Human Molecular Genetics, 23(R1), R40–46.Google Scholar
Zhang, M., Pritchard, M. R., Middleton, F. A., Horton, J. A. and Damron, T. A. (2008). Microarray analysis of perichondral and reserve growth plate zones identifies differential gene expressions and signal pathways. Bone, 43(3), 511520.Google Scholar
Zhao, Z., Tavoosidana, G., Sjolinder, M., et al. (2006). Circular chromosome conformation capture (4C) uncovers extensive networks of epigenetically regulated intra- and interchromosomal interactions. Nature Genetics, 38(11), 13411347.Google Scholar
Zhou, H. Y., Katsman, Y., Dhaliwal, N. K., et al. (2014). A Sox2 distal enhancer cluster regulates embryonic stem cell differentiation potential. Genes and Development, 28(24), 26992711.Google Scholar
Zou, H., Wieser, R., Massague, J. and Niswander, L. (1997). Distinct roles of type I bone morphogenetic protein receptors in the formation and differentiation of cartilage. Genes and Development, 11(17), 21912203.Google Scholar

Save book to Kindle

To save this book to your Kindle, first ensure coreplatform@cambridge.org is added to your Approved Personal Document E-mail List under your Personal Document Settings on the Manage Your Content and Devices page of your Amazon account. Then enter the ‘name’ part of your Kindle email address below. Find out more about saving to your Kindle.

Note you can select to save to either the @free.kindle.com or @kindle.com variations. ‘@free.kindle.com’ emails are free but can only be saved to your device when it is connected to wi-fi. ‘@kindle.com’ emails can be delivered even when you are not connected to wi-fi, but note that service fees apply.

Find out more about the Kindle Personal Document Service.

Available formats
×

Save book to Dropbox

To save content items to your account, please confirm that you agree to abide by our usage policies. If this is the first time you use this feature, you will be asked to authorise Cambridge Core to connect with your account. Find out more about saving content to Dropbox.

Available formats
×

Save book to Google Drive

To save content items to your account, please confirm that you agree to abide by our usage policies. If this is the first time you use this feature, you will be asked to authorise Cambridge Core to connect with your account. Find out more about saving content to Google Drive.

Available formats
×