Skip to main content
NIHPA Author Manuscripts logoLink to NIHPA Author Manuscripts
. Author manuscript; available in PMC: 2020 Oct 1.
Published in final edited form as: Curr Osteoporos Rep. 2019 Oct;17(5):301–323. doi: 10.1007/s11914-019-00527-9

Complex Phenotypes: Mechanisms Underlying Variation in Human Stature

Pushpanathan Muthuirulan 1, Terence D Capellini 1,2,*
PMCID: PMC6819265  NIHMSID: NIHMS1538011  PMID: 31441021

Abstract

Purpose of Review

The goal of the review is to provide a comprehensive overview of the current understanding of mechanisms underlying variation in human stature.

Recent Findings

Human height is an anthropometric trait that varies considerably within human populations as well as across the globe. Historically, much research focus was placed on understanding the biology of growth plate chondrocytes and how modifications to core chondrocyte proliferation and differentiation pathways potentially shaped height attainment in normal as well as pathological contexts. Recently, much progress has been made to improve our understanding regarding the mechanisms underlying the normal and pathological range of height variation within as well as between human populations, and today it is understood to reflect complex interactions among a myriad of genetic, environmental, and evolutionary factors. Indeed, recent improvements in genetics (e.g., GWAS) and breakthroughs in functional genomics (e.g., whole exome sequencing, DNA methylation analysis, ATAC-sequencing, and CRISPR) have shed light on previously unknown pathways/mechanisms governing pathological and common height variation. Additionally, use of an evolutionary perspective has also revealed important mechanisms that have shaped height across the planet.

Summary

This review provides an overview of the current knowledge of the biological mechanisms underlying height variation by highlighting new research findings on skeletal growth control with an emphasis on previously unknown pathways/mechanisms influencing pathological and common height variation. In this context, this review also discusses how evolutionary forces likely shaped the genomic architecture of height across the globe.

Keywords: height, complex trait, genetics, heritability, GWAS, environment, evolution, natural selection, GDF5, chondrocyte, gene regulation

Introduction

Among anthropometric traits, human height has been the most extensively studied across the sciences, due in part, to the desire among researchers to understand the mechanisms underlying its marked variation across the globe. Typically, height within any population follows a normal distribution, with a small fraction (3–5%) of individuals exhibiting extreme short and tall height phenotypes and most others hovering around the mean. While the height distributions of any two human populations usually overlap, differences in mean and variance reflect genetic, environmental, and evolutionary factors shaping growth phenotypes. For example, regardless of sampled population, height is 60–70% heritable, indicating a strong role for genetics underlying variation [1]. Yet, environmental factors also significantly influence differences in height attainment around the world; for example, despite its high heritability, there has been a secular increase in height in Western populations arising in part from improvements in nutrition and infectious disease prevention [2]. Importantly, evolutionary forces, such as natural selection, gene flow, genetic drift, and mutation, in concert with environmental and genetic factors, shape height phenotypes and its underlying genomic architecture within and between populations. For example, recent work has revealed that natural selection and other evolutionary forces likely shaped the genomic architecture of heights in different peoples around the globe [3].

What has become evident after hundreds of years of study is that genetic, environmental, and evolutionary factors often impinge directly or indirectly on skeletal growth, and specifically on how chondrocytes, the key cell type responding to growth signals, are regulated during development. During endochondral ossification, mesenchymal cells in the embryonic limb bud or vertebrae respond to signaling cues to condense and differentiate into chondrocytes to form transient models of future calcified bones. At the ends of each model, cartilaginous growth plates form, consisting of distinct zones of gradually differentiating chondrocytes (Fig. 1). Initially, resting zone chondrocytes replicate at a slow rate and give rise to proliferative chondrocytes, which in turn divide relatively quickly into clonally-related daughter cells. These daughter cells typically align into columns along the model’s longitudinal axis and through their expansion and proliferation bones elongate [4, 5]. Proliferative chondrocytes eventually stop dividing and terminally differentiate into hypertrophic chondrocytes, which then increase in size up to twenty-fold, significantly contributing to longitudinal bone growth [6]. Hypertrophic chondrocytes calcify the surrounding extracellular matrix, undergo apoptosis, all the while secreting signaling factors that attract invading osteoblasts and osteoclasts into the model in order to form bone and remove cartilage, respectively [7]. A small proportion of hypertrophic chondrocytes also directly differentiate into osteoblasts [8, 7].

Fig. 1. Growth plates in bones and height variation.

Fig. 1

The growth plate is a cartilaginous structure situated in the ends of long bones. It consists of special cells called chondrocytes that are arranged into distinct zones: resting, proliferative, pre-hypertrophic, and hypertrophic zones. The complex regulation at the growth plate determines length of long bones and human height. See text for details.

While endochondral ossification begins in embryogenesis, in humans, long bone growth plates remain cartilaginous and active centers of proliferation and differentiation through the first two decades of life, and it is at these sites that height variation is chiefly mediated. Indeed, it is the balance of growth plate chondrocyte proliferation, hypertrophy, and senescence as well as the timing of growth plate (epiphyseal) closure that determines bone length over the lifetime, in both normal and pathological phenotypes [9]. Recent improvements in genetics and functional genomics have shed light on the complex genetic architecture underlying this balance but have also revealed additional biological, environmental, and evolutionary factors that drive height variation within and between populations. Here, we discuss novel findings on skeletal growth control and highlight previously unknown pathways influencing growth control in the context of extreme and common height variation. We conclude with a discussion on how evolutionary factors have likely shaped heights’ genetic architecture across the globe.

Genetics and Height Variation

(a). Extreme Height Phenotypes

Among the most commonly studied human traits are extreme tall or short height phenotypes, defined as heights greater than two standard deviations above and below the mean, respectively [10, 11]. Typically, these growth phenotypes are heterogeneous in etiology, resulting from a myriad of different inputs, most notably genetic factors influencing chondrogenesis and growth. In cases where genetic mutations have been identified, they often consist of nucleotide changes localized to coding regions, are extremely rare in populations (i.e., have minor allele frequencies, MAF<0.01), are of strong effect in their influence on growth, and are often evident as dominant negative or recessive genotypes [12] (individual loci involved in extreme height phenotypes are reviewed in Table 1).

Table 1.

Pathogenic Coding Variants Underlying Growth Disorders

Growth Disorders Clinical Features Affected Gene(s) Gene Functions
Short Stature Phenotypes:
Microcephalic osteodysplastic primordial dwarfism type 1 (MOPD1) Small head size (microcephaly); abnormal bone growth (skeletal dysplasia); distinctive facial features; brain anomalies, sparse hair and eyebrows; dry skin; short limbs; dislocation of the hips and elbows; seizures; and intellectual disability. RNU4ATAC [153, 154] Small nuclear RNA (snRNA) that is part of the U12- dependent minor spliceosome complex and necessary for proper splicing of U12-dependent introns.
Microcephalic osteodysplastic primordial dwarfism type II (MOPDII) Short stature (dwarfism) with other skeletal abnormalities (osteodysplasia); unusual small head size (microcephaly); high-pitched nasal voice; some have a narrowing of the voicebox (subglottic stenosis); facial features include a prominent nose, full cheeks, a long midface, and a small jaw; small teeth (microdontia). PCNT [155] Pericentrin (PCNT) is important for normal functioning of the centrosomes, cytoskeleton, and cell-cycle progression.
Multiple epiphyseal dysplasia Mild short stature, malformations of the hands, feet, and knees and abnormal curvature of the spine (scoliosis); inward- and upward-turning foot (club foot), an opening in the roof of the mouth (cleft palate), an unusual curving of the fingers or toes (clinodactyly), ear swelling, abnormality of the kneecap (double-layered patella). COL9A1, COL9A2, COL9A3 [156], MATN3, COMP [157], SLC26A2 [158] COL9A1, COL9A2, and COL9A3 proteins strengthen and support connective tissues such as skin, bone, cartilage, tendons, and ligaments. MANT3 protein play a role in the organization of collagen and other cartilage proteins. COMP is an extracellular matrix protein that play role in cell growth, cell division and apoptosis. SLC26A2 protein transports charged molecules (ions), particularly sulfate ions, across cell membranes and is essential for normal cartilage formation.
IMAGe syndrome Short stature, distinctive facial features, such as a prominent forehead, low-set ears, and a short nose with a flat nasal bridge, premature fusion of certain bones of the skull (craniosynostosis), a split in the soft flap of tissue that hangs from the back of the mouth (cleft or bifid uvula), Other possible features include high levels of calcium in the blood (hypercalcemia) or urine (hypercalcuria) and a shortage of growth hormone in childhood. CDKN1C[159] CDKN1C protein encoded by this gene is involved in controlling growth before birth and preventing the developing fetus from becoming too large.
Infantile Nephropathic Cystinosis Severe short stature, blond hair, fair skin, moderate dehydration, generalized impaired proximal tubular reabsorptive capacity with severe fluid-electrolyte balance alterations, hypophosphatemic rickets that causes bone deformities. CTNS [160] The causative gene, CTNS encodes cystinosin, a lysosomal membrane protein that specifically moves the amino acid cystine out of the lysosome.
Dyggve-Melchior-Clausen (DMC) dysplasia and Smith-McCort (SMC) dysplasia Progressive short stature with short trunk dwarfism, microcephaly, protruding sternum, and psychomotor retardation, generalized abnormalities of the epiphyses and metaphyses, and a distinctive lacy appearance of the iliac crest. DYM [161] This gene encodes dymeclin which regulates Golgi-associated secretory pathways that are essential to endochondral bone formation during early development.
Seckel syndrome Short stature (dwarfism); abnormally small head (microcephaly); moderate to severe mental retardation; unusual characteristic of facial features including “beak-like” protrusion of the nose; abnormally large eyes, a narrow face, malformed ears, and/or an unusually small jaw (micrognathia); malformation of the foot in a twisted position (clubfoot), and/or absence of one pair of ribs. PCNT [162]; CENPJ [163]; ATR [164], ATRIP [165], CEP152 [166], CtIP [167]. PCNT is important for centrosomal function. CENPJ, ATR, ATRIP, CEP152 and CtIP genes encoding proteins that control cellular responses to DNA damage.
Wolcott-Rallison syndrome Short stature, walking difficulties, short trunk, excessive lordosis, a short and broad chest, and genu valgum, early infancy with symptoms of diabetes mellitus. EIF2AK3 [168] The protein encoded by this gene phosphorylates the alpha subunit of eukaryotic translation-initiation factor 2 and reduce translational initiation and repression of global protein synthesis.
Fanconi Anemia Short stature, bone marrow failure, skin pigmentation abnormalities, and characteristic malformations of upper extremities, head, eyes, ears, kidneys, developmental defects, predisposition to cancer, GH deficiency (GHD), hypothyroidism, and hypogonadism. FANCA [169, 170] The gene encodes FANCA proteins which involved in Fanconi anemia (FA) pathway, that operates as a post-replication repair or a cell cycle checkpoint.
Achondroplasia, hypochondroplasia, thanatophoric dysplasia, proportionate short stature Short stature, short arms and legs, enlarged head (macrocephaly) with a prominent forehead, fingers are typically short and hand with a trident appearance, normal intelligence, hypochondroplasia features tend to be milder. FGFR3 (gain of function mutation)[171] FGFR3 protein regulates bone growth by limiting the formation of bone from cartilage (a process called ossification), particularly in the long bones. It serves as negative regulator of bone growth.
Small for Gestational Age, Familial Short Stature, Idiopathic Short Stature Advanced bone age, short stature, early growth cessation, midface hypoplasia, flat nasal bridge, prognathism, posteriorly rotated ears, broad forehead, broad great toes, short thumbs, brachydactyly, joint problems, exaggerated lumbar lordosis, and genu valgum. ACAN [172174] Extracellular matrix, aggrecan.
Brachydactyly type A1, A2, C Short limbs, brachydactyly, foot abnormalities. GDF5 [175, 176] This gene encodes a secreted ligand of the TGF-beta (transforming growth factor-beta) superfamily of proteins that bind various TGF-beta receptors leading to recruitment and activation of SMAD family transcription factors that regulate gene expression.
Meier-Gorlin syndrome Short stature, narrow long bones in the arms and legs, a deformity of the knee joint, and delayed bone age, small mouth (microstomia), underdeveloped lower jaw (micrognathia), full lips, and a narrow nose with a high nasal bridge, testes are small or undescended (cryptorchidism) in males, small external genital folds (hypoplasia of the labia majora) and small breasts in affected females. ORC1, ORC4, ORC6, CDT1, CDC6[177, 178] Components of the DNA pre-replication complex which ensure that DNA replication occurs only once per cell division and is required for cells to divide.
Silver-Russell syndrome Low birth weight, postnatal short stature, characteristic facial features, and body asymmetry. IGF2[179, 180] Secreted signaling molecule (IGF2) that promotes growth and division of cells in many tissues, including cartilage development during postnatal long bone growth.
Laron syndrome Dwarfism, facial phenotype, obesity and hypogenitalism, hypoglycemia, hypercholesterolemia and sleep disorders. GHR[181] GHR gene encodes for growth hormone receptor that play critical role in growth hormone signaling.
Brachydactyly type A1 Short limbs, brachydactyly, foot abnormalities. BMPR1B [182] This gene encodes a member of the bone morphogenetic protein (BMP) receptor involved in endochondral bone formation and embryogenesis.
Acrocapitofemoral dysplasia, brachydactyly type A1 Short-limbed dwarfism, relatively large head circumference, short stature becomes more pronounced with age, cone-shaped epiphyses observed in ACFD patient. IHH [183] This gene encodes a member of the hedgehog family of protein that regulate a variety of developmental processes including growth, patterning and morphogenesis.
Rainbow syndrome Shortening of long bones in arms and legs, particularly the forearms, brachydactyly, wedge-shaped spinal bones, fused or missing ribs and short stature, distinctive facial features, such as a broad forehead, widely spaced eyes, short nose, triangle shaped mouth, increased bone mineral density (osteosclerosis) affecting the bones of the skull ROR2 [184], FZD2 [185], WNT5A [186], DVL1 [187], DVL3 [185] These genes encodes proteins that involved in chemical signaling pathways called Wnt signaling, which affect many aspects of development
Leri-Weill dyschondrosteosis Shortening of the long bones in the arms and legs (mesomelia), short stature, abnormality of the wrist and forearm bones (Madelung deformity), increased muscle mass (muscle hypertrophy), bowing of a bone in the lower leg called the tibia, a greater-than-normal angling of the elbow away from the body, and a high arched palate. SHOX [188] SHOX gene is essential for the development of the skeleton. It plays a particularly important role in the growth and maturation of bones in the arms and legs.
Noonan syndrome Mildly unusual facial features, short stature, heart defects, bleeding problems, skeletal malformations, delayed puberty, short neck, excess neck skin, abnormal side-to-side curvature of the skin. PTPN11 [189], SOS1 [190],RAF1[191], and RIT1[192] These genes encodes proteins important in the RAS/MAPK cell signaling pathway, which is needed for cell division and growth.
Costello syndrome Characteristic coarse facies, short stature, distinctive hand posture and appearance, severe feeding difficulty, and failure to thrive. HRAS [193] H-Ras protein is part of a RAS-MAP kinase pathway that helps control cell growth and division.
Coffin-Lowry syndrome Short stature, facial dysmorphism, development retardation and hearing defect RPS6KA3 [194] This gene encodes RSK proteins that play a role in several important cellular processes including cell growth and division, cell differentiation and apoptosis.
Bardet-Biedl syndrome Short stature and brachydactyly, stemming, intellectual disability and retinitis pigmentosa. SCAPER [195] This gene encodes SCAPER protein, which is associated with mitotic progression.
Peters plus syndrome Short stature, an opening in the lip (cleft lip) with or without an opening in the roof of the mouth (cleft palate), eye abnormality, distinctive facial features, and intellectual disability. B3GALTL [196] This gene encodes an enzyme called beta 3-glucosyltransferase (B3Glc-T), which is involved in the glycosylation.
Acromesomelic dysplasia, Maroteaux type (AMDM) Short stature and disproportionate shortening of limbs NPR2 [197] This gene encodes an NPR protein, which is the primary receptor for C-type natriuretic peptide (CNP) and it play a role in regulation of skeletal growth.
3-M syndrome Severe growth retardation, delayed bone age, distinctive facial features, normal mental development. CUL7 [198], OBSL1 [199], CCDC8 [200] These genes play role in microtubule stabilization and genome stability
Campomelic dysplasia Short stature, distinctive facial features, including a small chin, prominent eyes, and a flat face, weakened cartilage and difficulty in breathing. SOX9 [201] This gene encodes a protein SOX9, which is critical for the formation of many different tissues and organs during embryonic development. The SOX9 protein also regulates the activity of other genes, particularly genes involved in the development of the skeleton and reproductive organs.
Tall Stature Phenotypes:
Peutz-Jeghers syndrome (PJS) Delayed development, hypermobility, tall stature and advanced bone age STK11 [202] This gene encodes an enzyme, serine/threonine kinase 11, which is tumor suppressor and it helps to keep cells from growing and dividing too fast or in an uncontrolled way
Extreme tall stature Extreme tall stature and increased expression of androgen receptor IGF1RR1353H (activating mutation)[203] This gene encodes for IGF1 receptor, which is activated by hormone called insulin-like growth factor 1 and it play essential role in insulin signaling
Weaver syndrome Tall stature, intellectual disability and distinct facial features EZH2 [204] This gene encodes an enzyme called histone methyltransferase that play role in histone methylation
Marfan syndrome Tall stature, abnormalities in the heart, blood vessels, eyes, bones, and joints FBN1 [205] This gene encodes a protein called fibrillin-1 that provide strength and flexibility to connective tissues.
Tatton-Brown-Rahman syndrome (TBRS) Tall stature, intellectual disability and a distinctive facial appearance
DNMT3A [206] This gene encode an enzyme called DNA methyltransferase 3 alpha that involved in DNA methylation
Malan syndrome Tall stature, wide spectrum of malformations, intellectual disability and/or macrocephaly. NFIX [207] NFIX are transcription factors plays a pivotal role during the development of brain and skeleton.
Sotos syndrome Tall stature, prominent forehead, coarse facial features, macrocephaly, large ears and learning disability. NSD1 [208] This gene encodes an enzyme called histone methyltransferase that play an important role in histone methylation.
CATSHL syndrome Camptodactyly, tall stature, abnormality of lower limb joint, scoliosis, and hearing loss FGFR3 (Partial loss of function mutation) [209] FGFR3 protein regulates bone growth by limiting the formation of bone from cartilage (a process called ossification), particularly in the long bones. It serves as negative regulator of bone growth.
Estrogen deficiency Tall stature, lack of pubertal development and osteopenia (decrease in the mineral density of bone). CYP19A1 [210] The CYP19A1 gene provides instructions for making an enzyme called aromatase, which is responsible for the aromatization of androgen into estrogens.
Estrogen resistance Tall stature, elevated serum estrogen, abnormal serum gonadotropin, failure of epiphyseal fusion, and possibly insulin resistance ESR1 (loss of function) [30] This gene encodes for estrogen receptor that play important role in regulation of eukaryotic gene expression, cellular proliferation and differentiation in target tissues
Beckwith-Wiedemann syndrome (BWS) Affected infants are considerably larger than normal (macrosomia) and tend to be taller than their peers during childhood, abnormally large abdominal organs (visceromegaly), creases in the skin near the ears, low blood sugar (hypoglycemia) in infancy, and kidney abnormalities H19, CDKN1C, IGF2, and KCNQ1OT1 [80] H19 is a long noncoding RNA that has a role in the negative regulation (or limiting) of body weight and cell proliferation. CDKN1C protein is involved in controlling growth before birth. IGF2 promotes growth and division of cells in many tissues, including cartilage development during postnatal long bone growth. KCNQ1OT1 is a non-coding RNA, which regulate genes that are essential for normal growth and development before birth.
Childhood-onset obesity Tall stature/increased growth velocity, development of severe obesity, persistent food-seeking behaviour. MC4R [211] The protein encoded by this gene is a member of the melanocortin receptor family. The encoded protein interacts with adrenocorticotropic and MSH hormones and is mediated by G proteins
Multiple synostosis syndrome 1 Progressive symphalangism, multiple joint fusions, conductive deafness, mild facial dysmorphism, delayed puberty bone age, and closure of the epiphyseal lines of long bones with tall stature. NOG [212] The protein encoded by this gene is called noggin, which involved in the development of many body tissues, including nerve tissue, muscles, and bones.
Isolated ACTH deficiency (IAD) Tall Stature, secondary adrenocortical insufficiency with low or absent cortisol production, normal secretion of pituitary hormones other than adrenocorticotropic hormone (ACTH) and the absence of structural pituitary defects TBX19 [213] This gene encodes for a transcriptional regulator involved in developmental processes.

Recently, whole exome sequencing (WES) on affected and unaffected siblings has permitted the rapid identification of relatively large effect, rare variant loci underlying Mendelian inherited height phenotypes. For example, WES of three families with autosomal-dominant short stature, advanced bone age, and premature growth cessation identified heterozygous mutations in ACAN (Aggrecan) [13], as did a larger study on 200 short stature patients, which identified variants in about 16.5% of affected individuals, with ACAN mutations being most prevalent (2.5%) [14]. ACAN is an integral part of the extracellular matrix (ECM) in cartilaginous tissue, and disruptions to it influence growth plate organization and chondrocyte differentiation ultimately impacting longitudinal growth [15]. WES on other severe short stature patients also identified mutations in B4GALT7, CUL7, FAM111A, OBSL1, and SRCAP [16], while exome sequencing on a family with hypochondroplasia identified a novel missense mutation in FGFR3 [17], a negative regulator of bone growth [18]. Most recently, WES on 20 short stature patients lead to the potential cause of short stature in half with pathological alleles at ACAN, BRAF, COL2A1, HRAS, LMNA, PRKAR1A, PTPN11, and SLC7A7 [19].

WES and Sanger sequencing methods are efficient for identifying mutations in coding regions, yet patients can also have large effect non-coding alterations, especially those influencing chromosome structure and organization. Scanning patient chromosomes for large structural modifications using karyotyping and fluorescence in situ hybridization [20] has been somewhat effective in identifying these larger regions, which at times can be quite broad (i.e.., mega base intervals) and contain many affected coding and non-coding features making the specific causative mutation difficult to pinpoint. By altering large swaths of sequence, chromosomal modifications also can have detectable impacts on the regulatory control of gene expression further obscuring causal variant discovery. Such modifications can impact topologically associating domains (TAD) that are believed to stabilize regulatory element-gene promoter interactions that initialize and maintain transcription [21]. For example, alterations to a TAD boundary spanning the WNT6/IHH/EPHA4/PAX3 locus is implicated in human limb and digit developmental defects [22], whereas chromosomal rearrangements in the regulatory landscape of PITX1, a major hind limb transcription factor, causes Liebenberg syndrome, an autosomal-dominant upper-limb malformation in which arms acquire morphological characteristics of legs [23, 24]. Interestingly, in these cases, rare coding mutations specific to IHH and PITX1 also underlie extreme short stature [25], likely due to their disruption of chondrocyte differentiation, whereas higher frequency genetic variants located in non-coding regions have been implicated in normal height variation, likely due to more subtle cis-regulatory effects on growth plate dynamics (see below).

There are more classic examples of chromosomal alterations influencing specific coding regions underlying extreme stature phenotypes including Turner syndrome (missing or incomplete X chromosome), manifesting as short height, and Klinefelter syndrome (47, XXY), manifesting as tall height [26, 27]. For example, one major locus, SHOX (short stature homeobox), present at Xp22.3, encodes a transcription factor expressed in the developing limbs that regulates chondrocyte differentiation and proliferation [28]. Increased copy number variation (CNV) of SHOX may be in part responsible for the taller height phenotypes of Klinefelter patients, whereas chromosomal deletions encompassing SHOX have been linked to Leri-Weill dyschondrosteosis (LWD), a rare genetic disorder characterized by forearm and leg shortening, and associated short stature [29]. For more detailed reviews on extreme height genetics, please see [30, 31], and Table 1.

(b). Common Height Variation

In general, it is believed that rare (MAF<0.01) variants of large effect (i.e., mutations) that typically underlie pathological height, do not individually reach high enough allele frequencies within populations to account for the observed patterns of non-pathological, heritable variation in height. As a result, geneticists have sought to identify higher frequency - i.e., common (MAF>0.01) variants, using Genome-wide Association Studies (GWAS). GWAS involve revealing statistical correlations between common single nucleotide polymorphisms (SNPs), densely genotyped across the genome using a SNP array or chip, and height, measured using standardized protocols. SNPs are considered significantly associated if they fall below a standardized GWAS association p-value of 5×10−8. Often significantly detected loci contain numerous SNPs that are each roughly correlated with height because they are inherited in sequence blocks (i.e., haplotypes), are in modest to strong linkage disequilibrium (LD), and are generally at similar frequencies in sampled populations. This issue of having numerous linked associated SNPs makes identifying the causal variant extremely difficult. Recent improvements in genomic sequencing, haplotype imputation methods, statistical approaches, functional genomics assays, and the formation of large consortia (e.g., Genetic Investigation of Anthropometric Traits (GIANT) focusing on hundreds of thousands of people) as well as more focused investigations (e.g., population-specific GWAS) has led to increases in variant discovery and the whittling-down of haplotypes to smaller lists of putatively causal variants. While the geographic region of focus of GWAS has been predominantly Europe, those conducted on Asians [3237] and African/African-Americans [3840] has revealed shared as well as novel height signals. More recently, population-specific GWAS on Greenland Inuits [41], Sardinians [42], and Peruvians [43] have also contributed novel loci driving height variation (see below).

Of the many height GWAS [e.g., 4449, 40, 39, 50, 37, 41, 42, 51, 43], currently, the largest study by Yengo and colleagues [51] is on ~700,000 people predominantly of European descent. At the GWAS association p-value, they revealed 3,290 associated common variants found across 712 genomic loci, revealing loci with multiple independent signals. For example, nineteen significant signals were found within a 1.05 Mb locus containing the IGF1 gene involved in chondrogenesis. Independent signals within this locus may reflect distinct functional haplotypes harboring separate causative variants (i.e., those in very weak LD) each within some functionally relevant sequence, supporting the notion that genes have complex regulatory systems that both qualitatively (i.e., spatially) and quantitatively (i.e., level) mediate gene expression and phenotypic diversity.

Interestingly, these 3,290 variants explain only ~24.6% of the predicted genetic variance in height, reflecting the well-known issue of “missing heritability” [52]. Several studies have offered insight into this issue by trying to find variants that account for the remaining unexplained heritability. Recent studies [5358] used improved statistical modeling of a larger set of common variants to demonstrate that by factoring in those variants just failing significance, the missing heritability may be explained. However, a most recent study [59] approached this issue from a different angle in which they used whole genome sequencing on 21,620 unrelated Europeans to partition variants into those that are rare versus more common and at varying levels of LD. When they applied these data in concert with height data, they identified that rare variants in low LD also significantly contribute to and largely explain this missing heritability. Indeed, while no individual rare variant explains any population level associations with height, many rare variants in low to modest LD with common variants (i.e., even with those falling well above statistical thresholds) en masse tentatively appear to influence heritable variation in height [59]. These findings also indicate that narrowing down association intervals to true causal variants may be much harder than previously thought.

The sheer number of independent associations, as well as the preliminary finding that rare variants significantly explain a portion of inherited variation, experimentally demonstrate that height is extremely polygenic. Analyses of variant effect size distributions reveal that only ~1% of variants have moderate effect sizes (i.e., ~0.5cm/allele). For example, in three studies [4951], the ZBTB38 locus (rs2871960) was the most significant locus identified with each allele increasing height by ~0.4cm; similar to the effects of previously identified variants at GDF5-UQCC (rs6060369) [47] and HMGA2 (rs1042725) [46]. Recently, GWAS on more localized populations reveal that common variants can have a much larger effect previously thought. In Sardinians, an intronic variant (rs150199504) in KCNQ1, a voltage-gated potassium channel encoding gene reduces height by ~1.8cm [42]; in Greenland Inuits, an intronic variant (rs7115739) in FADS3, a gene involved in fatty acid metabolism reduces height by ~1.9cm [41] and in Peruvians, a missense variant (rs200342067) in FBN1, a gene encoding ECM glycoprotein that may interact with TGF-β in tissue homeostasis reduces height by ~2.2cm [43]. Despite these findings, it is important to point out that greater than 99% of common GWAS and rare variants have extremely minor effects on height variation (i.e., <5mm/allele) [51]. These findings strongly support insights made in the early 1900’s by R.A Fisher [60], who modeled height polygenicity and the infinitesimal effects of an extremely large number of Mendelian loci on height variation.

Given this extreme polygenicity and findings that height loci overlap with associated loci for other complex traits and diseases (reviewed in [6164]), Boyle and Pritchard [64] have suggested that height is an omnigenic trait, where the actions of large portions of the genome influence normal phenotypic variation and heritability. In this model, a sizable portion of overall allelic variation (~4%), overlapping a series of core regions (e.g., possibly growth/cartilage relevant pathways) and many peripherally-acting ones (e.g., biological processes that indirectly impinge of height), drive height variation; with the largest effect variants only modestly enriched in sequences that act directly on height and the vast majority of variants instead distributed ubiquitously across the genome (e.g., every 100 kb on average) each contributing minimal amounts to height variation. These common variants, with effects well below 5 mm, are enriched in active chromatin regions, with little signal of variants influencing heritability in inactive sequences. Furthermore, while there are specific enrichment signals in regulatory elements for chondrocytes/cartilage (see below), there are nonetheless quite strong signals of enrichment in regulatory elements active across a number of cell types, and therefore, these more pleiotropic variants likely also mediate other observed trait and disease associations. Overall, it suggests that the effects of evolution in shaping other aspects of human biology may often indirectly impinge on growth altering pathways influencing height (see below).

As alluded to above, studies using GWAS data reveal that height-associated loci are enriched near genes involved in a number of different biological processes. Lango Allen and colleagues [49] discovered 180 height loci in a GWAS on 183,727 individuals and found that variants were enriched near genes involved in Hedgehog, Transforming Growth Factor (TGF)-β , and Growth Hormone (GH) pathways, while Wood and colleagues [50] discovered 697 variants in a GWAS on 253,288 individuals and found additional enrichments in signaling by Fibroblast Growth Factors (FGFs), WNT/Beta-Catenin, chondroitin sulfate-related genes, mechanistic target of rapamycin (mTOR), osteoglycin, and glycosaminoglycans. Yengo and colleagues [51] identified the strongest enrichments near genes contributing to skeletal growth, chondrocyte biology, and cartilage/bone ECM (ACAN, ADAMTS17, EFEMP1, FBLN5), but also in Bone Morphogenetic Protein (BMP) pathways (BMP2, BMP6, GDF5, and NOG) and Hedgehog pathways (HHIP, IHH, PTCH1). However, they also revealed signals near genes involved in chromatin remodeling (DOT1L, HMGA1, HMGA2, SCHM1), cell cycle regulation (ANAPC13, CABLES1, CDK6, NCAPG), GPI-anchor protein synthesis (PIGP), fatty acid elongation (HSD17B12), among many others processes [65, 51]. Given the role of the Hippo-Yap pathway in size control in animals, GWAS has also identified signals near LATS2, TEAD1, VGLL2-4, and YAP1, with variants in VGLL3 associated with height only in women and in YAP1 and VGLL3 associated with shorter stature during pubertal growth [66].

The finding that some loci have effects on height via their roles in pubertal growth and chondrocyte development is not surprising considering that comparisons of Mendelian genetics studies on extreme height phenotypes and GWAS meta-analyses have revealed that many loci harboring common height variants can also have rarer coding mutations involved in monogenic, often congenital growth and height disorders. Indeed, these studies reveal the importance of pre- and early post-natal mechanisms in height control [65] (see Table.1). For example, Eurasian individuals carrying a common high-frequency 130 kb haplotype spanning GDF5-UQCC1 are ~0.4cm shorter [47], the likely causal variant (rs4911178) in a GDF5 growth plate enhancer [67]. Conversely, Hunter-Thompson type chondrodysplasia [68], DuPan syndrome [69], and Acromesomelic Dysplasia type – Grebe [70] patients possess missense GDF5 coding mutations, resulting in severe short stature and joint phenotypes; syndromes which phenocopy brachypodism mice harboring Gdf5 coding mutations [71]. Likewise, a common variant at HMGA2 (rs1042725) increases adult height by ~0.4cm, whereas rare, severe coding mutations in HMGA2 markedly alter body size in humans and “pygmy” mice [46]. A number of other height GWAS loci (ACAN, C6orf173, HHIP, HLA, PTCH1, SF3B4, SPAG17, and ZNFX1) are associated with infant length [72], whereas others (CABLES1, CDK10, LCORL, TSEN15, ZBTB38, and ZNF638) are associated with familial short stature [73], indicating the effects of variants during early development.

Given the relationships between common and rare variants and their respective effect sizes at a number of developmental loci, it has been suggested that the strongest phenotypic effects on height are most likely caused by rare variants, which unfortunately, may also be deleterious due to their generally pleiotropic nature. Marouli and colleagues [74] tested the association between 241,453 variants (83% coding with MAF≤0.05) and adult height variation in 711,428 individuals, and observed that the largest effect sizes were for four rare missense variants located in AR (rs137852591), CRISPLD2 (rs148934412), IHH (rs142036701), and STC2 (rs148833559). Carriers with the most significant rare variant at STC2, a regulator of postnatal growth [75], were ~2 cm taller than non-carriers, while carriers with rare variants at AR, CRISPLD2, and IHH were 2 cm shorter than non-carriers. Importantly in all four cases, the genes within these loci also result in growth disorders when mutated. Given the recent initial findings [59] on rare variants and their roles in height heritability, it is also likely that individual rare variants of modest effect will be found after more improved statistical modelling and by examining their roles in functional non-coding annotations.

With regard to the non-coding genome, GWAS has revealed an important functional impact that the non-coding genome has on height variation; evident by the finding that greater than 95% of GWAS height variants reside in non-coding regions. While many of these variants likely impact transcriptional regulation (see below), some signals likely impact RNA biology. Variants located within miRNA target sites at genes within associated loci likely contribute to variation in human stature. For example, Let-7 miRNA-binding site polymorphisms in the 3’UTR of CDK6, DOT1L, HMGA2, LIN28B, and PAPPA alter miRNA binding and therefore influence miRNA-mediated gene regulation. Because miRNAs are critical post-translational regulators of gene expression, it would be interesting to test whether the other Let-7 targets or Let-7 itself are regulators of adult height [76, 65, 48]. Likewise, hsa-miR-140–5p plays an important role in skeletal development in vivo [77] and it has a target site polymorphism in FGFRL1 3′ UTR that could modulate FGFRL1 expression levels affecting bone formation and height variation [78]. Other miRNAs enriched in GWAS genes that may regulate adult height are described in [79]. Variants in long non-coding RNA (lncRNA) may also shape height variation. H19 is a lncRNA with roles in the negative regulation of body weight and cell proliferation. Similarly, KCNQ1OT1 is another that regulates genes that are essential for normal prenatal development and growth. The aberrant expression of H19 and KCNQ1OT1 appear involved in Beckwith-Wiedemann Syndrome (BWS), where affected individuals are considerably larger than normal at birth and childhood [80]. Three GWAS variants, two (rs147239461 and rs7482510) at IGF2-H19 and one (rs143840904) at KCNQ1, a gene encoded in the locus, are located within an imprinted 1.8 Mb region on chromosome 11p15 and these may affect growth by paternal imprinting [81]. In a separate study on Sardinians, a different variant (rs150199504), one in moderate LD with rs143840904, may act via maternal imprinting and appears associated with shorter height [42], further obscuring variant causality at the locus.

While GWAS have predominantly focused on overall (aka standing) height, the identified loci cannot easily be partitioned into those influencing hind limb length, vertebral height (trunk height), or head height. Understanding the different mechanisms controlling regional growth is important especially as these anatomical regions have different embryonic origins, respond differently to growth hormones pre- and post-natally, as well as markedly differ in proportion across human groups. GWAS on sitting height ratio (SHR), the ratio of sitting height (measured from top of skull to surface upon which a person is sitting) to standing height was performed on 3,545 African Americans and 21,590 Europeans [82]. Chan and colleagues not only validated SHR differences between these groups, with African Americans having lower SHR, but found that SHR was quite heritable with common variants explaining 26% and 39% of the total variance in European and African Americans, respectively. Their GWAS on African Americans identified one locus (rs10736877) near C10orf90, whose function unfortunately remains unknown. Conversely, GWAS on Europeans identified several loci: (1) two novel loci at PTPRM (rs140449984), potentially involved in growth control and osteocyte biology [83], and NFATC2 (rs228836), involved in growth plate chondrocyte differentiation [84]; (2) two loci previously reported as standing height loci, Tbx2 (rs882367), whose loss of function results in skeletal malformations [85, 86] and BCKDHB (rs6931421), expressed in embryonic limb buds [87]; and (3) one locus containing IGFBP3 (rs1722141), involved in IGF signaling and skeletal growth [88], where other SNPs not in LD independently correlate with standing height, suggesting different mechanisms controlling height at this locus. They also observed enrichments of standing height-associated variants [50] with SHR variants. For example, 71 (of 130) loci had height-increasing alleles predicted to decrease SHR, indicating that these act disproportionately on long bone growth, whereas 59 (of 130) loci had the height increasing allele predicted to increase the SHR, indicating they act to affect head or spine length. Like overall height, most SHR variants reside in the non-coding genome and remain understudied.

To date, GWAS studies have predominantly focused on identifying SNP associations with height, although other variant types may also underlie normal height variation. While novel chromosomal-level alterations are often incompatible with life [89] and/or cause syndromic phenotypes (see above), most humans possess several large structural modifications that do not lead to overt, deleterious insults but instead may underlie normal variation in physical traits [90]. These features may increase or decrease gene CNV, but they could also have regulatory effects on gene expression in target tissues [91]. A genome-wide CNV association study for height in 618 Chinese subjects found four CNV at chromosomes 6p21.3, 8p23.3–23.2, 9p23, and 16p12.1, with a CNV gain at 8p23.3–23.2 associated with lower height, and a CNV loss at 6p21.3 associated with lower height. Likewise, another CNV analysis on a clinical cohort of children found that subjects with short stature had an increased CNV burden suggesting they might contribute to variation in stature in the general population [92]. In another study [93], CNV regions nearby genes from height GWAS were examined and a 17.7 kb deletion was identified at chromosomal position 12q24.33 downstream of GPR133, a locus linked with human height.

Transcriptomics, Epigenomics, and Height Variation

A recent focus has been on understanding how GWAS variants en masse influence functional aspects of height biology. In these studies, often the top GWAS variants and those in moderate to strong LD are examined for how often they overlap functionally annotated sequences in the genome. The functional annotations (e.g., marking gene bodies, promoters, enhancers, CpG sites, etc.) derive from transcriptomic and functional epigenomics datasets from the ENCODE Project [94], the Roadmap Epigenomics Project [95], and the FANTOM5 Project [96] or individual research labs focusing on chondrocyte biology (e.g., [97, 98]). In most cases, proximity-based approaches (i.e., the assignment of a variant in a genomic feature to its closest gene) are used to assess whether height variants are enriched near a particular functional class of genes relative to background. Consequently, these analyses also help whittle-down non-coding signals to potential causal regulatory variants for functional follow-up, albeit there are only a couple of instances where follow-up functional validation experiments have been performed (see below). One of the earliest proximity-based enrichment study, by Lui et al. [97], focused on the proximity of 207 GWAS height variants ([49] and others) to genes differentially expressed in the murine growth plate. After identifying 427 differentially growth plate expressed genes, they saw specific enrichments for height loci near these genes, indicating that variants often target chondrocyte gene expression.

A number of studies on common and pathological height variation have focused on whether genetic variants influence gene expression and/or function via gene promoter CpG methylation [99]. CpG methylation analysis revealed that 72 of 87 (82.8%) genes previously shown to be most associated with height contained CpG islands upstream of their transcription start sites and correlated with gene regulation. Yengo and colleagues [51] used methylation quantitative trait loci derived from blood [100] to identify 775 methylation sites showing pleiotropic associations with height, and in the process revealing potential mechanisms. For example, at one CpG site (cg19825988) within ZBTB38, a zinc finger transcriptional activator that binds methylated DNA, they found that increased methylation had the largest mediation effect on height compared to other sites. Methylation has also been examined in the context of pathological phenotypes, but at candidate genes. For example, hypomethylation at CpG sites in the HOXA4 promoter has been associated with Russell-Silver syndrome and growth restrictions in children [101], whereas abnormal methylation at the maternal GNAS promoter has been associated with severe obesity and short stature [102]. Moreover, Ouni and colleagues [103] found that CG methylation of the P2 promoter of IGF1 gene plays a role in idiopathic short stature. Interestingly, height GWAS also identified significant associations at IGF1 (rs17032362, rs1520223, rs5742692, rs35767 and rs1457595) [103].

Some studies have focused on height variation in the context of genome-wide regulatory functionality and chromatin accessibility. Trynka and colleagues [104] examined whether 697 associated variants (and those in LD) from Wood et al. [50] were enriched in DNase I open chromatin regions from 217 ENCODE cell types, and found the strongest enrichments for height variants overlapping sites in embryonic stem cells (H1-hESCs), indicating that the regulatory control of embryonic development may in part underlie height variation. Chan and colleagues [82] examined whether SHR variants were enriched in FANTOM enhancer annotations, and of seven significant loci, six showed overlaps with enhancer elements, albeit FANTOM did not examine tissues directly relevant to cartilage growth. Furthermore, of the 130 height variants overlapping SHR variants, this group also had stronger enrichments for enhancer overlap than non-overlapping height variants, and when the 130 variants were divided into those likely influencing long bones (71 loci) versus head/spine (59 loci), stronger enrichment for long bone influencing variants was observed. In support of their omnigenic model, Boyle and Pritchard [64] also performed enrichment studies on height variants as well as general common variants and found that common variants are enriched in active chromatin regions across many cell types, indicative of pleiotropy, with little signal of variants influencing heritability in inactive sequences.

Capellini, Guo and colleagues [98] profiled open chromatin regions in chondrocytes of the developing femur in mice using ATAC-seq [105, 106]. After identifying thousands of orthologous human sites and intersecting them with GWAS height variants, they found substantial enrichment above genomic background levels. The specificity of their findings was supported by the lack of enrichments in these femur datasets for GWAS variants from other complex, polygenic traits and diseases, indicating that not all complex traits share the same underlying architectures. They also used transcriptomic data on the growth plate [97], and found height variants in open chromatin regions were enriched near genes that were also differentially expressed in the growth plate. To move closer towards causal variant discovery, they also performed follow-up functional tests in human chondrocytes on variants in the Chondroitin Sulfate Synthase 1 (CHSY1) locus. CHSY1 coding mutations in humans and mice cause severe skeletal phenotypes including Temtamy preaxial brachydactyly syndrome, characterized by short stature and preaxial brachydactyly [107109, 91]. They identified that a C/T base-pair change at rs9920291 modified a repressor sequence which impacted gene expression in vitro, with the height-increasing allele (T) making the regulatory element a weaker repressor. In unpublished data, this variant position also mediates HOXD13 binding while overexpression of HOXD13 led to CHSY1 upregulation in human chondrocytes. None of the remaining three ATAC-seq variants in CHSY1 influenced reporter expression, pointing strongly towards rs9920291 as putatively causal. These in-depth studies reveal that not all variants in open chromatin regions need to effect expression. While all of these studies are compelling, much more focus should be spent on causal variant discovery especially at GWAS loci of large effect, those involved in extreme height phenotypes, as well as those shaped by natural selection and other evolutionary forces (see below).

Evolutionary Mechanisms and Conclusions

While reinforcing key roles that chondrocyte and skeletal pathways have in mediating height variation, GWAS have also revealed the sheer number of independent inputs controlling height as well as the importance of the non-coding genome. These two features should be thought of in the context of how evolution has shaped height phenotypes world-wide, given that it shapes the effect sizes, allelic frequencies, and types of loci mediating height attainment. The non-coding genome is an especially important place to investigate further because regulatory variants influencing phenotypes can increase in frequency within populations without the extreme pleiotropic consequences typically associated with coding variants. With this in mind, we delve into a brief discussion on how evolution shaped height loci, and we highlight areas of future inquiry.

There is a common (mis-)belief that in dispersing out-of-Africa (OoA), human height became extremely diverse as migratory populations adapted to new ecological settings in the Old World. However, prior to this OoA dispersal, which occurred between 130,000–50,000 years ago (reviewed in [110]), heights within Africa were already likely quite varied given the marked ecological diversity across the continent, and that modern Africans are markedly genetically diverse and varied in height [62, 111, 112]. In ancient Africa, the diversity of functional genetic haplotypes would have served as the standing genetic variation for which adaptive and non-adaptive regulatory evolution occurred within Africa as well as during OoA colonization. Future efforts must characterize the genetic architecture of height in Africa through GWAS on more localized populations, and how it relates to ancient genetic and phenotypic diversity across the continent.

Early OoA populations experienced a substantial genetic bottleneck reducing genetic variation [113, 114], but upon which a number of evolutionary forces acted. When groups entered Eurasia they became quite dispersed, smaller in population size, and exposed to new interactions with ancient populations that exited Africa hundreds of thousands of years earlier (reviewed in [110]). For example, in living Eurasians there are introgressed genomic sequences from Neanderthals and Denisovans, each of which resided and ecologically adapted to different environments well before modern human contact [115120]. Neandertal alleles in human loci associated with sitting and standing height appear to influence modern human height attainment, an effect possibly on developmental growth [121]. Some of this introgressed variation consisted of novel sequences arising in these archaic hominins but some were older variants that became reintroduced into OoA groups that lost them in the bottleneck [122]. Regardless, these archaic haplotypes, and the novel epistatic interactions that arose via their introgression in modern human genomes, served as part of the genetic fodder upon which natural selection could shape modern height variation. Future efforts should be made to understand regulatory diversity in sitting and standing height, and the effects of archaic introgressed sequences on height biology.

Some of the possible natural selection pressures often invoked in early OoA populations and their descendants include thermoregulation, energy conservation, positive assortative mating [123], among others [62], although, their effects have been examined mainly at individual loci. For example, at the GDF5-UQCC1 locus, selection on a chondrocyte growth plate enhancer variant “A” (rs4911178) for its height- or growth-reducing effects, increased a 130 kb haplotype, and likely contributed to shorter height in southern Europeans and in East Asians [67, 124126]. Here, the selection pressures may have related to thermoregulation and/or energy conservation in part because the specific genetic alterations appear to shorten limb lengths, and this may have been a mechanism to conserve body heat [127] and/or reduce overall body size for energy conservation during resource scarcity [124]. Analyses of the frequencies of the shorter height allele across parts of Asia suggests evidence of a geographic cline, such that populations living in colder climates (e.g., Siberians) appear to have increased frequencies of the “A” variant especially compared to southern Asians (unpublished data; [67]), matching cold adaptation expectations. In Europe, the “A” variant is high in frequency in southern Europeans, a possible influence of gene flow and migration from earlier first farmer populations sharing genetic signals with these populations [128, 129] and potentially serving to maintain smaller body size during energetic scarcities. Interestingly, the short height “A” variant is also present in Neandertals and Denisovans, who lived in Eurasia at times when it was markedly colder, albeit it occurs on a related haplotype that may have been independently selected for its effects on growth [67]. When considering body size more broadly across mammals, the GDF5-UQCC1 locus has repeatedly been shown to act as a key quantitative trait locus and under selection in species where body size has been artificially selected [47, 130132], indicating that the locus presents with functional genetic diversity and may be evolvable, subject to repeated selection.

Other loci have been found associated with shorter height and likely under positive selection, although their signals were detected in geographically localized populations, and likely arose more recently. In Sardinians, a small stature population in Europe, selection on the short height intronic variant in KCNQ1 potentially reflected an island biogeography response in which smaller body size aided in lowering energy requirements to accommodate limited caloric availability [42]. In Greenland Inuits, selection on the short height intronic variant in FADS3 may have been in response to their fat-rich diet, which indirectly decreased height, although it could be due to cold stress and the need for shortened extremities, as Inuit body and appendage sizes are often cited examples of cold-adaptation [41]. Most recently, in Peruvians, selection on a missense variant in FBN1 may have reflected coastal lifestyles, which we are assuming to mean potentially lower energetic resource availabilities, although this remains unclear [43]. Future GWAS efforts focused on more localized populations (as well as comparisons between closely related populations that differ in phenotype) will help to determine which other height loci appear under selection, as this should aid in understanding the diversity of selective pressures driving height variation across similar and different ecologies. To this end, genetic studies on different short stature “pygmy” populations in West and Central Africa [133138] and Asia [139, 140] have revealed a number of genes involved in height variation (IGF-1; GH, and other growth factors) and which show the effects of recent selection and convergence on growth control mechanisms.

There has been a recent desire to go beyond individual loci and ask whether GWAS height loci en masse show evidence of natural selection. In this regard, height has often been considered the quintessential example of a polygenic adaptation - i.e., reflecting directional (positive selective) changes via small allele frequency shifts at many phenotypically relevant sites [141]. A number of studies [42, 61, 140, 142, 143] have used genome-wide selection tests and polygenic risk scores (PRS) that take into consideration variation at multiple genetic loci and their associated weights and found subtle effects of positive selection in shaping height variation globally as well as within Europeans and other populations. Guo and colleagues [61] found evidence that selection altered the allele frequencies and LD patterns of height variants, and noted that PRS in Europeans were higher than in Africans, which were in turn higher than East Asians, matching expectations based on observed phenotypic differences in mean height across these geographic regions. Zoledziewska and colleagues [42], Turchin and colleagues [142] and Robinson and colleagues [143] found that within Europeans, approximately half of GWAS loci show some evidence of positive selection along a north-south cline, with taller populations in the north having an abundance of height increasing alleles and shorter populations in the south having an abundance of height decreasing alleles. Field and colleagues [144] using their singleton-density score (SDS) on the UK10K Project data, found that during the past 2000–3000 years, selection for increased height may have drove allele frequency shifts across most of the genome, favoring taller alleles in Britons. Finally, studies comparing ancient and modern genomes and using GWAS height data indicate that the patterns of height variant distribution in Europe could reflect contributions from at least three different founder human populations, each of these populations may have varied significantly in a number of anthropometric traits [128, 129].

Recently, the extent of polygenic selection for height has been called into serious question. At least three new studies [145147] have reexamined the north-south cline in height variation in Europe and positive selection signals for taller heights in Britons, each originally determined using summary statistics from GIANT, but now reassessed using genotype and phenotype data from the larger, more homogeneous UK Biobank dataset. These new studies found that the signals for polygenic adaptation to be markedly attenuated especially when based on a large number of variants falling below genome-wide significance which are extremely sensitive to biases due to uncorrected population structure in GWAS [145147]. They emphasize that while the height associations in the original studies are reproducible, the PRS calculated using this new dataset are not nearly as strong, putting into question the extent to which height loci in Europeans was shaped by positive selection rather than the effects of genetic drift or other population demographic processes.

GWAS height variants have also been studied for signals of stabilizing selection and negative selection. In both instances, height is modeled to reflect that it not only has many inputs (i.e., is highly polygenic) but is also deeply tied to other phenotypes (i.e., is highly pleiotropic). In the case of stabilizing selection, selection likely has not acted in a directional manner to shape only height or its architecture but rather on intermediate phenotypes that permit adaptive (and non-detrimental) changes in co-evolving traits. To this end, Sanjak and colleagues [3] in their analysis of height associations from the UK biobank found signals of weak stabilizing selection of GWAS loci influencing height, while Simons and colleagues [148] developed a model demonstrating the effects of stabilizing selection and pleiotropy on height. In the case of negative selection, selection against pleiotropically acting variants has likely had an impact on height variation. Zeng and colleagues [149] found in examining GWAS loci, that for height, lower MAF variants in general tended to have larger effect sizes than common variants suggesting that the former are tolerated in affecting height only because they occur at low frequencies. This was supported by findings on LD structure and allele frequency distributions by other studies [150, 151] that each found evidence of the effects of negative selection on height loci and other anthropometric traits. Finally, a recent whole genome sequencing study by Wainschtein and colleagues [59] found that variants (MAF<0.1) in low LD bins were enriched for non-synonymous and protein truncating variants, which also on average contributed much more to heritability estimates than synonymous or non-coding variants, and in turn may be indicative of negative selection.

Overall, future efforts need to improve methods to control for population demography, and better tease out the effects that each of these three different types of selection has had at GWAS height loci. Disentangling the influence of selection as compared to population demography is important especially because it helps focus causal variant discovery efforts and our understanding of why such variants rose to high levels in populations. To this end, it will still be important to study individual loci and their effects on height variation in individual populations in the past. Detailed understanding of the evolutionary history of the haplotypes harboring true causal height variants at individual loci will shed light on the extent to which height is pleiotropic versus modular, as well as whether signals of the polygenic basis of height are evenly distributed across the genome or focused in specific functional regions of importance (i.e., both insights impinge on our understanding of an omnigenic model [64]). Such an understanding may even help understand connections between height and other disease risks or traits. For example, selection on the short height “A” regulatory variant (rs4911178) in the GDF5-UQCC1 locus led to a substantial increase in frequency of its linked 130 kb haplotype, which consequently is also associated with an increased risk of knee and hip osteoarthritis [67]. On this haplotype there are likely separate genetic regulatory variants, which have not themselves been under selection (i.e., since osteoarthritis typically occurs well after reproductive age), but which confer osteoarthritis risk at joint specific sites [67]. Indeed, detailed studies at individual loci can reveal functional interconnections between variants that may be missed by broader, genome-wide scale investigations.

Finally, as humans have experienced recent environmental and cultural changes, such as the industrial revolution, and modern improvements in nutrition and infection disease prevention and treatment, cultural factors also have significantly shaped human height variation (reviewed in [62, 152]). There is a wealth of data on how infectious disease state, access to adequate nutrition, socioeconomic status, among other factors, influence human stature but they currently fall short of being directly connected to genetic and epigenetic mechanisms influencing underlying height biology. A greater understanding of how these factors epigenetically influence height loci, and how genetic variants at these loci mediate such effects is needed. Ultimately, for these environmental and cultural factors to have an evolutionary effect on height variation, it will be important to link them to specific functional loci and ultimately reveal how they impact fitness.

Fig. 2. Genomic approaches provide insight into biological mechanisms underlying height variation.

Fig. 2

GWAS and functional genomic approaches (e.g., exome sequencing, DNA methylation analysis, ATAC-seq and CRISPR-Cas9) have contributed to discovery of novel height related traits and shed light into complex biological mechanisms that drive variation in human height. See text for details.

Acknowledgments:

The authors would like to thank Drs. Graham Coop (University of California, Davis) and Peter Visscher (University of Queensland) for helpful discussions on the genetics and evolution of height, as well as members of the Capellini Laboratory and the Department of Human Evolutionary Biology at Harvard University for many insightful discussions. We would also like to acknowledgement the terrific community of scholars studying height for whom we were unable to cite their work; studies on the biology of height are vast and it was challenging to choose relevant articles from so much excellent work.

Funding: This research has been funded in part by NIH NIAMS R01 (1R01AR070139–01A1) to TDC.

Footnotes

Publisher's Disclaimer: This Author Accepted Manuscript is a PDF file of a an unedited peer-reviewed manuscript that has been accepted for publication but has not been copyedited or corrected. The official version of record that is published in the journal is kept up to date and so may therefore differ from this version.

Conflict of Interest: The authors declare that they have no conflict of interest.

Conflict of Interest

Pushpanathan Muthuirulan and Terence D. Capellini declare no conflict of interest.

Human and Animal Rights and Informed Consent

This article does not contain any studies with human or animal subjects performed by any of the authors.

References

Papers of particular interest, published recently, have been highlighted as:

• Of importance

••Of major importance

  • 1.Jelenkovic A, Hur YM, Sund R, Yokoyama Y, Siribaddana SH, Hotopf M et al. Genetic and environmental influences on adult human height across birth cohorts from 1886 to 1994. eLife. 2016;5. doi: 10.7554/eLife.20320. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 2.Perkins JM, Subramanian SV, Davey Smith G, Ozaltin E. Adult height, nutrition, and population health. Nutr Rev. 2016;74(3):149–65. doi: 10.1093/nutrit/nuv105. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • *3.Sanjak JS, Sidorenko J, Robinson MR, Thornton KR, Visscher PM. Evidence of directional and stabilizing selection in contemporary humans. Proc Natl Acad Sci U S A. 2018;115(1):151–6. doi: 10.1073/pnas.1707227114. [DOI] [PMC free article] [PubMed] [Google Scholar]; This paper describes a moving optimum model for the evolution of complex, quantitative traits in humans.
  • 4.Kronenberg HM. Developmental regulation of the growth plate. Nature. 2003;423(6937):332–6. doi: 10.1038/nature01657. [DOI] [PubMed] [Google Scholar]
  • 5.Romereim SM, Conoan NH, Chen B, Dudley AT. A dynamic cell adhesion surface regulates tissue architecture in growth plate cartilage. Development. 2014;141(10):2085–95. doi: 10.1242/dev.105452. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • **6.Tsang KY, Chan D, Cheah KS. Fate of growth plate hypertrophic chondrocytes: death or lineage extension? Dev Growth Differ. 2015;57(2):179–92. doi: 10.1111/dgd.12203. [DOI] [PubMed] [Google Scholar]; This paper describes the contribution of chondrocytes to the osteoblast lineage in bone development, disease, and repair.
  • **7.Aghajanian P, Mohan S. The art of building bone: emerging role of chondrocyte-to-osteoblast transdifferentiation in endochondral ossification. Bone Res. 2018;6:19. doi: 10.1038/s41413-018-0021-z. [DOI] [PMC free article] [PubMed] [Google Scholar]; This paper describes an emerging model for chondrocyte-to-osteoblast transdifferentiation and their implications for the treatment of skeletal diseases.
  • **8.Zhou X, von der Mark K, Henry S, Norton W, Adams H, de Crombrugghe B. Chondrocytes transdifferentiate into osteoblasts in endochondral bone during development, postnatal growth and fracture healing in mice. PLoS Genet. 2014;10(12):e1004820. doi: 10.1371/journal.pgen.1004820. [DOI] [PMC free article] [PubMed] [Google Scholar]; This paper describes the process of chondrocyte to osteoblast transdifferentiation important for endochondral bone formation and bone fracture healing.
  • 9.Pineault KM, Swinehart IT, Garthus KN, Ho E, Yao Q, Schipani E et al. Hox11 genes regulate postnatal longitudinal bone growth and growth plate proliferation. Biol Open. 2015;4(11):1538–48. doi: 10.1242/bio.012500. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 10.Kumar S Tall stature in children: differential diagnosis and management. International journal of pediatric endocrinology. 2013;2013(1):p53. [Google Scholar]
  • 11.Barstow C, Rerucha C. Evaluation of Short and Tall Stature in Children. Am Fam Physician. 2015;92(1):43–50. [PubMed] [Google Scholar]
  • 12.Axenovich TI, Zorkoltseva IV, Belonogova NM, Struchalin MV, Kirichenko AV, Kayser M et al. Linkage analysis of adult height in a large pedigree from a Dutch genetically isolated population. Hum Genet. 2009;126(3):457–71. doi: 10.1007/s00439-009-0686-x. [DOI] [PubMed] [Google Scholar]
  • *13.Nilsson O, Guo MH, Dunbar N, Popovic J, Flynn D, Jacobsen C et al. Short stature, accelerated bone maturation, and early growth cessation due to heterozygous aggrecan mutations. J Clin Endocrinol Metab. 2014;99(8):E1510–8. doi: 10.1210/jc.2014-1332. [DOI] [PMC free article] [PubMed] [Google Scholar]; This paper demonstrates that a heterozygous mutation in ACAN can cause short stature with advanced bone age.
  • 14.Hauer NN, Popp B, Schoeller E, Schuhmann S, Heath KE, Hisado-Oliva A et al. Clinical relevance of systematic phenotyping and exome sequencing in patients with short stature. Genet Med. 2018;20(6):630–8. doi: 10.1038/gim.2017.159. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 15.Lauing KL, Cortes M, Domowicz MS, Henry JG, Baria AT, Schwartz NB. Aggrecan is required for growth plate cytoarchitecture and differentiation. Dev Biol. 2014;396(2):224–36. doi: 10.1016/j.ydbio.2014.10.005. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • **16.Guo MH, Shen Y, Walvoord EC, Miller TC, Moon JE, Hirschhorn JN et al. Whole exome sequencing to identify genetic causes of short stature. Horm Res Paediatr. 2014;82(1):44–52. doi: 10.1159/000360857. [DOI] [PMC free article] [PubMed] [Google Scholar]; This paper demonstrates how whole exome sequencing can be utilized for clincially relevant diagnoses of patients with short stature.
  • 17.Chen J, Yang J, Zhao S, Ying H, Li G, Xu C. Identification of a novel mutation in the FGFR3 gene in a Chinese family with Hypochondroplasia. Gene. 2018;641:355–60. doi: 10.1016/j.gene.2017.10.062. [DOI] [PubMed] [Google Scholar]
  • 18.Wen X, Li X, Tang Y, Tang J, Zhou S, Xie Y et al. Chondrocyte FGFR3 Regulates Bone Mass by Inhibiting Osteogenesis. J Biol Chem. 2016;291(48):24912–21. doi: 10.1074/jbc.M116.730093. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 19.Zhu H, Yang Z, Sun J, Li W, Yang H, Wang L et al. Diagnostic Whole Exome Sequencing in Patients with Short Stature. bioRxiv. 2018. doi: 10.1101/414987. [DOI] [Google Scholar]
  • 20.Romero CJ, Mehta L, Rapaport R. Genetic Techniques in the Evaluation of Short Stature. Endocrinol Metab Clin North Am. 2016;45(2):345–58. doi: 10.1016/j.ecl.2016.02.006. [DOI] [PubMed] [Google Scholar]
  • 21.Dixon JR, Selvaraj S, Yue F, Kim A, Li Y, Shen Y et al. Topological domains in mammalian genomes identified by analysis of chromatin interactions. Nature. 2012;485(7398):376–80. doi: 10.1038/nature11082. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • *22.Lupianez DG, Kraft K, Heinrich V, Krawitz P, Brancati F, Klopocki E et al. Disruptions of topological chromatin domains cause pathogenic rewiring of gene-enhancer interactions. Cell. 2015;161(5):1012–25. doi: 10.1016/j.cell.2015.04.004. [DOI] [PMC free article] [PubMed] [Google Scholar]; This paper demostrates the functional importance of topologically associated domains (TADs) for orchestrating gene expression and also for predicting the pathogenecity of human structural variants in non-coding regions of the human genome.
  • 23.Spielmann M, Brancati F, Krawitz PM, Robinson PN, Ibrahim DM, Franke M et al. Homeotic arm-to-leg transformation associated with genomic rearrangements at the PITX1 locus. American journal of human genetics. 2012;91(4):629–35. doi: 10.1016/j.ajhg.2012.08.014. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 24.Seoighe DM, Gadancheva V, Regan R, McDaid J, Brenner C, Ennis S et al. A chromosomal 5q31.1 gain involving PITX1 causes Liebenberg syndrome. Am J Med Genet A. 2014;164A(11):2958–60. doi: 10.1002/ajmg.a.36712. [DOI] [PubMed] [Google Scholar]
  • 25.Alvarado DM, McCall K, Aferol H, Silva MJ, Garbow JR, Spees WM et al. Pitx1 haploinsufficiency causes clubfoot in humans and a clubfoot-like phenotype in mice. Hum Mol Genet. 2011;20(20):3943–52. doi: 10.1093/hmg/ddr313. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 26.Ouarezki Y, Cizmecioglu FM, Mansour C, Jones JH, Gault EJ, Mason A et al. Measured parental height in Turner syndrome-a valuable but underused diagnostic tool. Eur J Pediatr. 2018;177(2):171–9. doi: 10.1007/s00431-017-3045-2. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 27.Bonomi M, Rochira V, Pasquali D, Balercia G, Jannini EA, Ferlin A et al. Klinefelter syndrome (KS): genetics, clinical phenotype and hypogonadism. J Endocrinol Invest. 2017;40(2):123–34. doi: 10.1007/s40618-016-0541-6. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 28.Fukami M, Seki A, Ogata T. SHOX Haploinsufficiency as a Cause of Syndromic and Nonsyndromic Short Stature. Mol Syndromol. 2016;7(1):3–11. doi: 10.1159/000444596. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 29.Volejnikova J, Zapletalova J, Jarosova M, Urbankova H, Petr V, Klaskova E et al. Acute lymphoblastic leukemia in a child with Leri-Weill syndrome and complete SHOX gene deletion: A Case Report. Biomed Pap Med Fac Univ Palacky Olomouc Czech Repub. 2018;162(1):65–70. doi: 10.5507/bp.2018.002. [DOI] [PubMed] [Google Scholar]
  • 30.Baron J, Savendahl L, De Luca F, Dauber A, Phillip M, Wit JM et al. Short and tall stature: a new paradigm emerges. Nat Rev Endocrinol. 2015;11(12):735–46. doi: 10.1038/nrendo.2015.165. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 31.Grunauer M, Jorge AAL. Genetic short stature. Growth Horm IGF Res. 2018;38:29–33. doi: 10.1016/j.ghir.2017.12.003. [DOI] [PubMed] [Google Scholar]
  • 32.Cho YS, Go MJ, Kim YJ, Heo JY, Oh JH, Ban HJ et al. A large-scale genome-wide association study of Asian populations uncovers genetic factors influencing eight quantitative traits. Nat Genet. 2009;41(5):527–34. doi: 10.1038/ng.357. [DOI] [PubMed] [Google Scholar]
  • 33.Lei SF, Yang TL, Tan LJ, Chen XD, Guo Y, Guo YF et al. Genome-wide association scan for stature in Chinese: evidence for ethnic specific loci. Hum Genet. 2009;125(1):1–9. doi: 10.1007/s00439-008-0590-9. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 34.Kim JJ, Lee HI, Park T, Kim K, Lee JE, Cho NH et al. Identification of 15 loci influencing height in a Korean population. J Hum Genet. 2010;55(1):27–31. doi: 10.1038/jhg.2009.116. [DOI] [PubMed] [Google Scholar]
  • 35.Okada Y, Kamatani Y, Takahashi A, Matsuda K, Hosono N, Ohmiya H et al. A genome-wide association study in 19 633 Japanese subjects identified LHX3-QSOX2 and IGF1 as adult height loci. Hum Mol Genet. 2010;19(11):2303–12. doi: 10.1093/hmg/ddq091. [DOI] [PubMed] [Google Scholar]
  • 36.Croteau-Chonka DC, Marvelle AF, Lange EM, Lee NR, Adair LS, Lange LA et al. Genome-wide association study of anthropometric traits and evidence of interactions with age and study year in Filipino women. Obesity (Silver Spring). 2011;19(5):1019–27. doi: 10.1038/oby.2010.256. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 37.He Q, Morris BJ, Grove JS, Petrovitch H, Ross W, Masaki KH et al. Shorter men live longer: association of height with longevity and FOXO3 genotype in American men of Japanese ancestry. PLoS One. 2014;9(5):e94385. doi: 10.1371/journal.pone.0094385. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 38.Kang SJ, Chiang CW, Palmer CD, Tayo BO, Lettre G, Butler JL et al. Genome-wide association of anthropometric traits in African- and African-derived populations. Hum Mol Genet. 2010;19(13):2725–38. doi: 10.1093/hmg/ddq154. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 39.Carty CL, Johnson NA, Hutter CM, Reiner AP, Peters U, Tang H et al. Genome-wide association study of body height in African Americans: the Women’s Health Initiative SNP Health Association Resource (SHARe). Hum Mol Genet. 2012;21(3):711–20. doi: 10.1093/hmg/ddr489. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 40.N’Diaye A, Chen GK, Palmer CD, Ge B, Tayo B, Mathias RA et al. Identification, replication, and fine-mapping of Loci associated with adult height in individuals of african ancestry. PLoS Genet. 2011;7(10):e1002298. doi: 10.1371/journal.pgen.1002298. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • *41.Fumagalli M, Moltke I, Grarup N, Racimo F, Bjerregaard P, Jorgensen ME et al. Greenlandic Inuit show genetic signatures of diet and climate adaptation. Science. 2015;349(6254):1343–7. doi: 10.1126/science.aab2319. [DOI] [PubMed] [Google Scholar]; This paper describes the genetic and physiological adapatations of the Inuit to a diet rich in omega-3 polyunsaturated fatty acids (PUFAs).
  • *42.Zoledziewska M, Sidore C, Chiang CWK, Sanna S, Mulas A, Steri M et al. Height-reducing variants and selection for short stature in Sardinia. Nat Genet. 2015;47(11):1352–6. doi: 10.1038/ng.3403. [DOI] [PMC free article] [PubMed] [Google Scholar]; This paper demonstrates how selection for decreased height was acting among the Neolithic ancestors of the Sardinians.
  • *43.Asgari S, Luo Y, Belbin GM, Bartell E, Roger R, Slowikowski K et al. A positively selected, common, missense variant in FBN1 confers a 2.2 centimeter reduction of height in the Peruvian population. bioRxiv. 2019:561241. [Google Scholar]; This paper describes the genetic basis of short stature in Peruvian population.
  • 44.Thomas DC, Haile RW, Duggan D. Recent developments in genomewide association scans: a workshop summary and review. American journal of human genetics. 2005;77(3):337–45. doi: 10.1086/432962. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 45.Morris AP, Cardon LR. Whole genome association. Handbook of statistical genetics. 2007:1238–63. doi: 10.1002/9780470061619.ch37. [DOI] [Google Scholar]
  • 46.Weedon MN, Lettre G, Freathy RM, Lindgren CM, Voight BF, Perry JR et al. A common variant of HMGA2 is associated with adult and childhood height in the general population. Nat Genet. 2007;39(10):1245–50. doi: 10.1038/ng2121. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 47.Sanna S, Jackson AU, Nagaraja R, Willer CJ, Chen WM, Bonnycastle LL et al. Common variants in the GDF5-UQCC region are associated with variation in human height. Nat Genet. 2008;40(2):198–203. doi: 10.1038/ng.74. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 48.Lettre G, Jackson AU, Gieger C, Schumacher FR, Berndt SI, Sanna S et al. Identification of ten loci associated with height highlights new biological pathways in human growth. Nat Genet. 2008;40(5):584–91. doi: 10.1038/ng.125. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 49.Lango Allen H, Estrada K, Lettre G, Berndt SI, Weedon MN, Rivadeneira F et al. Hundreds of variants clustered in genomic loci and biological pathways affect human height. Nature. 2010;467(7317):832–8. doi: 10.1038/nature09410. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 50.Wood AR, Esko T, Yang J, Vedantam S, Pers TH, Gustafsson S et al. Defining the role of common variation in the genomic and biological architecture of adult human height. Nat Genet. 2014;46(11):1173–86. doi: 10.1038/ng.3097. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • **51.Yengo L, Sidorenko J, Kemper KE, Zheng Z, Wood AR, Weedon MN et al. Meta-analysis of genome-wide association studies for height and body mass index in approximately 700000 individuals of European ancestry. Hum Mol Genet. 2018;27(20):3641–9. doi: 10.1093/hmg/ddy271. [DOI] [PMC free article] [PubMed] [Google Scholar]; This paper demonstrates how genome-wide association studies (GWAS) can be utilized to achieve deeper insights into complex trait biology such as height.
  • 52.Manolio TA, Collins FS, Cox NJ, Goldstein DB, Hindorff LA, Hunter DJ et al. Finding the missing heritability of complex diseases. Nature. 2009;461(7265):747–53. doi: 10.1038/nature08494. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 53.Yang J, Benyamin B, McEvoy BP, Gordon S, Henders AK, Nyholt DR et al. Common SNPs explain a large proportion of the heritability for human height. Nat Genet. 2010;42(7):565–9. doi: 10.1038/ng.608. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 54.Yang J, Bakshi A, Zhu Z, Hemani G, Vinkhuyzen AA, Lee SH et al. Genetic variance estimation with imputed variants finds negligible missing heritability for human height and body mass index. Nat Genet. 2015;47(10):1114–20. doi: 10.1038/ng.3390. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 55.Okbay A, Beauchamp JP, Fontana MA, Lee JJ, Pers TH, Rietveld CA et al. Genome-wide association study identifies 74 loci associated with educational attainment. Nature. 2016;533(7604):539–42. doi: 10.1038/nature17671. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 56.Shi H, Kichaev G, Pasaniuc B. Contrasting the Genetic Architecture of 30 Complex Traits from Summary Association Data. American journal of human genetics. 2016;99(1):139–53. doi: 10.1016/j.ajhg.2016.05.013. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 57.Zhang L, Shen YP, Hu WZ, Ran S, Lin Y, Lei SF et al. A new method for estimating effect size distribution and heritability from genome-wide association summary results. Hum Genet. 2016;135(2):171–84. doi: 10.1007/s00439-015-1621-y. [DOI] [PubMed] [Google Scholar]
  • 58.Speed D, Cai N, Consortium U, Johnson MR, Nejentsev S, Balding DJ. Reevaluation of SNP heritability in complex human traits. Nat Genet. 2017;49(7):986–92. doi: 10.1038/ng.3865. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • **59.Wainschtein P, Jain DP, Yengo L, Zheng Z, TOPMed Anthropometry Working Group, Trans-Omics for Precision Medicine Consortium et al. Recovery of trait heritability from whole genome sequence data. bioRxiv. 2019. doi: 10.1101/588020. [DOI] [Google Scholar]; This paper describes that a portion of the missing heritability of complex traits such as height may be accounted for by rare variants.
  • 60.Fisher RA. XV.—The correlation between relatives on the supposition of Mendelian inheritance. Earth and Environmental Science Transactions of the Royal Society of Edinburgh. 1919;52(2):399–433. [Google Scholar]
  • 61.Guo J, Wu Y, Zhu Z, Zheng Z, Trzaskowski M, Zeng J et al. Global genetic differentiation of complex traits shaped by natural selection in humans. Nat Commun. 2018;9(1):1865. doi: 10.1038/s41467-018-04191-y. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 62.Stulp G, Barrett L. Evolutionary perspectives on human height variation. Biol Rev Camb Philos Soc. 2016;91(1):206–34. doi: 10.1111/brv.12165. [DOI] [PubMed] [Google Scholar]
  • 63.Pickrell JK, Berisa T, Liu JZ, Segurel L, Tung JY, Hinds DA. Detection and interpretation of shared genetic influences on 42 human traits. Nat Genet. 2016;48(7):709–17. doi: 10.1038/ng.3570. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • **64.Boyle EA, Li YI, Pritchard JK. An Expanded View of Complex Traits: From Polygenic to Omnigenic. Cell. 2017;169(7):1177–86. doi: 10.1016/j.cell.2017.05.038. [DOI] [PMC free article] [PubMed] [Google Scholar]; This paper describes how complex traits are driven by enormously large numbers of variants of small effects present across the genome.
  • 65.Lettre G Genetic regulation of adult stature. Curr Opin Pediatr. 2009;21(4):515–22. doi: 10.1097/MOP.0b013e32832c6dce. [DOI] [PubMed] [Google Scholar]
  • 66.Cousminer DL, Berry DJ, Timpson NJ, Ang W, Thiering E, Byrne EM et al. Genome-wide association and longitudinal analyses reveal genetic loci linking pubertal height growth, pubertal timing and childhood adiposity. Hum Mol Genet. 2013;22(13):2735–47. doi: 10.1093/hmg/ddt104. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • *67.Capellini TD, Chen H, Cao J, Doxey AC, Kiapour AM, Schoor M et al. Ancient selection for derived alleles at a GDF5 enhancer influencing human growth and osteoarthritis risk. Nat Genet. 2017;49(8):1202–10. doi: 10.1038/ng.3911. [DOI] [PMC free article] [PubMed] [Google Scholar]; This paper discovers an ancient regulatory variant that decreases height in modern human populations.
  • 68.Thomas JT, Lin K, Nandedkar M, Camargo M, Cervenka J, Luyten FP. A human chondrodysplasia due to a mutation in a TGF-beta superfamily member. Nat Genet. 1996;12(3):315–7. doi: 10.1038/ng0396-315. [DOI] [PubMed] [Google Scholar]
  • 69.Faiyaz-Ul-Haque M, Ahmad W, Zaidi SH, Haque S, Teebi AS, Ahmad M et al. Mutation in the cartilage-derived morphogenetic protein-1 (CDMP1) gene in a kindred affected with fibular hypoplasia and complex brachydactyly (DuPan syndrome). Clin Genet. 2002;61(6):454–8. [DOI] [PubMed] [Google Scholar]
  • 70.Umair M, Rafique A, Ullah A, Ahmad F, Ali RH, Nasir A et al. Novel homozygous sequence variants in the GDF5 gene underlie acromesomelic dysplasia type-grebe in consanguineous families. Congenit Anom (Kyoto). 2017;57(2):45–51. doi: 10.1111/cga.12187. [DOI] [PubMed] [Google Scholar]
  • 71.Storm EE, Huynh TV, Copeland NG, Jenkins NA, Kingsley DM, Lee SJ. Limb alterations in brachypodism mice due to mutations in a new member of the TGF beta-superfamily. Nature. 1994;368(6472):639–43. doi: 10.1038/368639a0. [DOI] [PubMed] [Google Scholar]
  • 72.van der Valk RJ, Kreiner-Moller E, Kooijman MN, Guxens M, Stergiakouli E, Saaf A et al. A novel common variant in DCST2 is associated with length in early life and height in adulthood. Hum Mol Genet. 2015;24(4):1155–68. doi: 10.1093/hmg/ddu510. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 73.Lin YJ, Liao WL, Wang CH, Tsai LP, Tang CH, Chen CH et al. Association of human height-related genetic variants with familial short stature in Han Chinese in Taiwan. Sci Rep. 2017;7(1):6372. doi: 10.1038/s41598-017-06766-z. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • **74.Marouli E, Graff M, Medina-Gomez C, Lo KS, Wood AR, Kjaer TR et al. Rare and low-frequency coding variants alter human adult height. Nature. 2017;542(7640):186–90. doi: 10.1038/nature21039. [DOI] [PMC free article] [PubMed] [Google Scholar]; This paper describes how rare and low-frequency coding variants influence the architecture of a complex human traits and also reports new genes and biological pathways implicated in human growth.
  • 75.Chang AC, Hook J, Lemckert FA, McDonald MM, Nguyen MA, Hardeman EC et al. The murine stanniocalcin 2 gene is a negative regulator of postnatal growth. Endocrinology. 2008;149(5):2403–10. doi: 10.1210/en.2007-1219. [DOI] [PubMed] [Google Scholar]
  • 76.de Vasconcellos JF, Byrnes C, Lee YT, Allwardt JM, Kaushal M, Rabel A et al. Tough decoy targeting of predominant let-7 miRNA species in adult human hematopoietic cells. J Transl Med. 2017;15(1):169. doi: 10.1186/s12967-017-1273-x. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 77.Luo W, Liu L, Yang L, Dong Y, Liu T, Wei X et al. The vitamin D receptor regulates miR-140–5p and targets the MAPK pathway in bone development. Metabolism. 2018;85:139–50. doi: 10.1016/j.metabol.2018.03.018. [DOI] [PubMed] [Google Scholar]
  • 78.Niu T, Liu N, Zhao M, Xie G, Zhang L, Li J et al. Identification of a novel FGFRL1 MicroRNA target site polymorphism for bone mineral density in meta-analyses of genome-wide association studies. Hum Mol Genet. 2015;24(16):4710–27. doi: 10.1093/hmg/ddv144. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 79.Okada Y, Muramatsu T, Suita N, Kanai M, Kawakami E, Iotchkova V et al. Significant impact of miRNA-target gene networks on genetics of human complex traits. Sci Rep. 2016;6:22223. doi: 10.1038/srep22223. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 80.Choufani S, Shuman C, Weksberg R. Molecular findings in Beckwith-Wiedemann syndrome. Am J Med Genet C Semin Med Genet. 2013;163C(2):131–40. doi: 10.1002/ajmg.c.31363. [DOI] [PubMed] [Google Scholar]
  • 81.Benonisdottir S, Oddsson A, Helgason A, Kristjansson RP, Sveinbjornsson G, Oskarsdottir A et al. Epigenetic and genetic components of height regulation. Nat Commun. 2016;7:13490. doi: 10.1038/ncomms13490. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • *82.Chan Y, Salem RM, Hsu YH, McMahon G, Pers TH, Vedantam S et al. Genome-wide Analysis of Body Proportion Classifies Height-Associated Variants by Mechanism of Action and Implicates Genes Important for Skeletal Development. American journal of human genetics. 2015;96(5):695–708. doi: 10.1016/j.ajhg.2015.02.018. [DOI] [PMC free article] [PubMed] [Google Scholar]; This paper describes how sitting height ratio (SHR) can be utilized to dissect the genetic basis of skeletal growth of the trunk and limbs.
  • 83.de Rooij KE, van der Velde M, de Wilt E, Deckers MM, Bezemer M, Waarsing JH et al. Identification of receptor-type protein tyrosine phosphatase mu as a new marker for osteocytes. Histochem Cell Biol. 2015;144(1):1–11. doi: 10.1007/s00418-015-1319-1. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 84.Lui JC, Baron J. Effects of glucocorticoids on the growth plate. Endocr Dev. 2011;20:187–93. doi: 10.1159/000321244. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 85.Ballif BC, Theisen A, Rosenfeld JA, Traylor RN, Gastier-Foster J, Thrush DL et al. Identification of a recurrent microdeletion at 17q23.1q23.2 flanked by segmental duplications associated with heart defects and limb abnormalities. American journal of human genetics. 2010;86(3):454–61. doi: 10.1016/j.ajhg.2010.01.038. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 86.Radio FC, Bernardini L, Loddo S, Bottillo I, Novelli A, Mingarelli R et al. TBX2 gene duplication associated with complex heart defect and skeletal malformations. Am J Med Genet A. 2010;152A(8):2061–6. doi: 10.1002/ajmg.a.33506. [DOI] [PubMed] [Google Scholar]
  • 87.Visel A, Thaller C, Eichele G. GenePaint.org: an atlas of gene expression patterns in the mouse embryo. Nucleic acids research. 2004;32(Database issue):D552–6. doi: 10.1093/nar/gkh029. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 88.Yakar S, Rosen CJ, Beamer WG, Ackert-Bicknell CL, Wu Y, Liu JL et al. Circulating levels of IGF-1 directly regulate bone growth and density. J Clin Invest. 2002;110(6):771–81. doi: 10.1172/JCI15463. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 89.Hardy K, Hardy PJ. 1(st) trimester miscarriage: four decades of study. Transl Pediatr. 2015;4(2):189–200. doi: 10.3978/j.issn.2224-4336.2015.03.05. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 90.Sudmant PH, Rausch T, Gardner EJ, Handsaker RE, Abyzov A, Huddleston J et al. An integrated map of structural variation in 2,504 human genomes. Nature. 2015;526(7571):75–81. doi: 10.1038/nature15394. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 91.Li Y, Laue K, Temtamy S, Aglan M, Kotan LD, Yigit G et al. Temtamy preaxial brachydactyly syndrome is caused by loss-of-function mutations in chondroitin synthase 1, a potential target of BMP signaling. American journal of human genetics. 2010;87(6):757–67. doi: 10.1016/j.ajhg.2010.10.003. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 92.Dauber A, Yu Y, Turchin MC, Chiang CW, Meng YA, Demerath EW et al. Genome-wide association of copy-number variation reveals an association between short stature and the presence of low-frequency genomic deletions. American journal of human genetics. 2011;89(6):751–9. doi: 10.1016/j.ajhg.2011.10.014. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 93.Kim YK, Moon S, Hwang MY, Kim DJ, Oh JH, Kim YJ et al. Gene-based copy number variation study reveals a microdeletion at 12q24 that influences height in the Korean population. Genomics. 2013;101(2):134–8. doi: 10.1016/j.ygeno.2012.11.002. [DOI] [PubMed] [Google Scholar]
  • 94.Consortium EP. An integrated encyclopedia of DNA elements in the human genome. Nature. 2012;489(7414):57–74. doi: 10.1038/nature11247. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 95.Roadmap Epigenomics C, Kundaje A, Meuleman W, Ernst J, Bilenky M, Yen A et al. Integrative analysis of 111 reference human epigenomes. Nature. 2015;518(7539):317–30. doi: 10.1038/nature14248. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 96.Lizio M, Harshbarger J, Shimoji H, Severin J, Kasukawa T, Sahin S et al. Gateways to the FANTOM5 promoter level mammalian expression atlas. Genome Biol. 2015;16:22. doi: 10.1186/s13059-014-0560-6. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 97.Lui JC, Nilsson O, Chan Y, Palmer CD, Andrade AC, Hirschhorn JN et al. Synthesizing genome-wide association studies and expression microarray reveals novel genes that act in the human growth plate to modulate height. Hum Mol Genet. 2012;21(23):5193–201. doi: 10.1093/hmg/dds347. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • *98.Guo M, Liu Z, Willen J, Shaw CP, Richard D, Jagoda E et al. Epigenetic profiling of growth plate chondrocytes sheds insight into regulatory genetic variation influencing height. eLife. 2017;6. doi: 10.7554/eLife.29329. [DOI] [PMC free article] [PubMed] [Google Scholar]; This papers demonstrates how integrating biologically relevant epigenetic information from growth plates with genetic association results can help to identify biological mechanisms important for human growth.
  • 99.Simeone P, Alberti S. Epigenetic heredity of human height. Physiol Rep. 2014;2(6). doi: 10.14814/phy2.12047. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 100.McRae AF, Marioni RE, Shah S, Yang J, Powell JE, Harris SE et al. Identification of 55,000 Replicated DNA Methylation QTL. Sci Rep. 2018;8(1):17605. doi: 10.1038/s41598-018-35871-w. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 101.Muurinen M, Hannula-Jouppi K, Reinius LE, Soderhall C, Merid SK, Bergstrom A et al. Hypomethylation of HOXA4 promoter is common in Silver-Russell syndrome and growth restriction and associates with stature in healthy children. Sci Rep. 2017;7(1):15693. doi: 10.1038/s41598-017-16070-5. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 102.Hanna P, Grybek V, Perez de Nanclares G, Tran LC, de Sanctis L, Elli F et al. Genetic and Epigenetic Defects at the GNAS Locus Lead to Distinct Patterns of Skeletal Growth but Similar Early-Onset Obesity. J Bone Miner Res. 2018;33(8):1480–8. doi: 10.1002/jbmr.3450. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 103.Ouni M, Belot MP, Castell AL, Fradin D, Bougneres P. The P2 promoter of the IGF1 gene is a major epigenetic locus for GH responsiveness. Pharmacogenomics J. 2016;16(1):102–6. doi: 10.1038/tpj.2015.26. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 104.Trynka G, Westra HJ, Slowikowski K, Hu X, Xu H, Stranger BE et al. Disentangling the Effects of Colocalizing Genomic Annotations to Functionally Prioritize Non-coding Variants within Complex-Trait Loci. American journal of human genetics. 2015;97(1):139–52. doi: 10.1016/j.ajhg.2015.05.016. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 105.Buenrostro JD, Giresi PG, Zaba LC, Chang HY, Greenleaf WJ. Transposition of native chromatin for fast and sensitive epigenomic profiling of open chromatin, DNA-binding proteins and nucleosome position. Nat Methods. 2013;10(12):1213–8. doi: 10.1038/nmeth.2688. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 106.Buenrostro JD, Wu B, Chang HY, Greenleaf WJ. ATAC-seq: A Method for Assaying Chromatin Accessibility Genome-Wide. Curr Protoc Mol Biol. 2015;109:21 9 1–9. doi: 10.1002/0471142727.mb2129s109. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 107.Wilson DG, Phamluong K, Lin WY, Barck K, Carano RA, Diehl L et al. Chondroitin sulfate synthase 1 (Chsy1) is required for bone development and digit patterning. Dev Biol. 2012;363(2):413–25. doi: 10.1016/j.ydbio.2012.01.005. [DOI] [PubMed] [Google Scholar]
  • 108.Temtamy SA, Meguid NA, Ismail SI, Ramzy MI. A new multiple congenital anomaly, mental retardation syndrome with preaxial brachydactyly, hyperphalangism, deafness and orodental anomalies. Clin Dysmorphol. 1998;7(4):249–55. [DOI] [PubMed] [Google Scholar]
  • 109.Tian J, Ling L, Shboul M, Lee H, O’Connor B, Merriman B et al. Loss of CHSY1, a secreted FRINGE enzyme, causes syndromic brachydactyly in humans via increased NOTCH signaling. American journal of human genetics. 2010;87(6):768–78. doi: 10.1016/j.ajhg.2010.11.005. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 110.Groucutt HS, Petraglia MD, Bailey G, Scerri EM, Parton A, Clark-Balzan L et al. Rethinking the dispersal of Homo sapiens out of Africa. Evol Anthropol. 2015;24(4):149–64. doi: 10.1002/evan.21455. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 111.Campbell MC, Tishkoff SA. African genetic diversity: implications for human demographic history, modern human origins, and complex disease mapping. Annu Rev Genomics Hum Genet. 2008;9:403–33. doi: 10.1146/annurev.genom.9.081307.164258. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 112.Choudhury A, Aron S, Sengupta D, Hazelhurst S, Ramsay M. African genetic diversity provides novel insights into evolutionary history and local adaptations. Hum Mol Genet. 2018;27(R2):R209–R18. doi: 10.1093/hmg/ddy161. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 113.Lopez S, van Dorp L, Hellenthal G. Human Dispersal Out of Africa: A Lasting Debate. Evol Bioinform Online. 2015;11(Suppl 2):57–68. doi: 10.4137/EBO.S33489. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 114.Henn BM, Cavalli-Sforza LL, Feldman MW. The great human expansion. Proc Natl Acad Sci U S A. 2012;109(44):17758–64. doi: 10.1073/pnas.1212380109. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 115.Nielsen R, Akey JM, Jakobsson M, Pritchard JK, Tishkoff S, Willerslev E. Tracing the peopling of the world through genomics. Nature. 2017;541(7637):302–10. doi: 10.1038/nature21347. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 116.Rogers AR, Bohlender RJ, Huff CD. Early history of Neanderthals and Denisovans. Proc Natl Acad Sci U S A. 2017;114(37):9859–63. doi: 10.1073/pnas.1706426114. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 117.Dannemann M, Andres AM, Kelso J. Introgression of Neandertal- and Denisovan-like Haplotypes Contributes to Adaptive Variation in Human Toll-like Receptors. American journal of human genetics. 2016;98(1):22–33. doi: 10.1016/j.ajhg.2015.11.015. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 118.Sankararaman S, Mallick S, Dannemann M, Prufer K, Kelso J, Paabo S et al. The genomic landscape of Neanderthal ancestry in present-day humans. Nature. 2014;507(7492):354–7. doi: 10.1038/nature12961. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 119.Vernot B, Akey JM. Resurrecting surviving Neandertal lineages from modern human genomes. Science. 2014;343(6174):1017–21. doi: 10.1126/science.1245938. [DOI] [PubMed] [Google Scholar]
  • 120.Jagoda E, Lawson DJ, Wall JD, Lambert D, Muller C, Westaway M et al. Disentangling Immediate Adaptive Introgression from Selection on Standing Introgressed Variation in Humans. Mol Biol Evol. 2017. doi: 10.1093/molbev/msx314. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 121.Dannemann M, Kelso J. The Contribution of Neanderthals to Phenotypic Variation in Modern Humans. American journal of human genetics. 2017;101(4):578–89. doi: 10.1016/j.ajhg.2017.09.010. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 122.Rinker DC, Simonti C, McArthur E, Shaw D, Hodges E, Capra JA. Neanderthal introgression reintroduced functional alleles lost in the human out of Africa bottleneck. bioRxiv. 2019. doi: 10.1101/533257. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 123.Stulp G, Simons MJ, Grasman S, Pollet TV. Assortative mating for human height: A meta-analysis. Am J Hum Biol. 2017;29(1). doi: 10.1002/ajhb.22917. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 124.Wu DD, Li GM, Jin W, Li Y, Zhang YP. Positive selection on the osteoarthritis-risk and decreased-height associated variants at the GDF5 gene in East Asians. PLoS One. 2012;7(8):e42553. doi: 10.1371/journal.pone.0042553. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 125.Voight BF, Kudaravalli S, Wen X, Pritchard JK. A map of recent positive selection in the human genome. PLoS Biol. 2006;4(3):e72. doi: 10.1371/journal.pbio.0040072. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 126.Grossman SR, Andersen KG, Shlyakhter I, Tabrizi S, Winnicki S, Yen A et al. Identifying recent adaptations in large-scale genomic data. Cell. 2013;152(4):703–13. doi: 10.1016/j.cell.2013.01.035. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 127.Allen JA. The influence of physical conditions in the genesis of species. Radical Review. 1877;1:108–40. [Google Scholar]
  • 128.Mathieson I, Lazaridis I, Rohland N, Mallick S, Patterson N, Roodenberg SA et al. Genome-wide patterns of selection in 230 ancient Eurasians. Nature. 2015;528(7583):499–503. doi: 10.1038/nature16152. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 129.Mathieson I, Alpaslan-Roodenberg S, Posth C, Szecsenyi-Nagy A, Rohland N, Mallick S et al. The genomic history of southeastern Europe. Nature. 2018;555(7695):197–203. doi: 10.1038/nature25778. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 130.Liu YF, Zan LS, Li K, Zhao SP, Xin YP, Lin Q et al. A novel polymorphism of GDF5 gene and its association with body measurement traits in Bos taurus and Bos indicus breeds. Mol Biol Rep. 2010;37(1):429–34. doi: 10.1007/s11033-009-9604-5. [DOI] [PubMed] [Google Scholar]
  • 131.Randhawa IA, Khatkar MS, Thomson PC, Raadsma HW. Composite selection signals can localize the trait specific genomic regions in multi-breed populations of cattle and sheep. BMC Genet. 2014;15:34. doi: 10.1186/1471-2156-15-34. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 132.Liu Y, Zan L, Zhao S, Huang H, Li Y, Tang Z et al. Molecular cloning, expression and characterization of bovine UQCC and its association with body measurement traits. Mol Cells. 2010;30(5):393–401. doi: 10.1007/s10059-010-0129-5. [DOI] [PubMed] [Google Scholar]
  • 133.Turnbull CM. Wayward servants: the two worlds of the African Pygmies. The Natural History Press. 1965(14):391. [Google Scholar]
  • 134.Perry GH, Dominy NJ. Evolution of the human pygmy phenotype. Trends Ecol Evol. 2009;24(4):218–25. doi: 10.1016/j.tree.2008.11.008. [DOI] [PubMed] [Google Scholar]
  • 135.Jarvis JP, Scheinfeldt LB, Soi S, Lambert C, Omberg L, Ferwerda B et al. Patterns of ancestry, signatures of natural selection, and genetic association with stature in Western African pygmies. PLoS Genet. 2012;8(4):e1002641. doi: 10.1371/journal.pgen.1002641. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 136.Becker NS, Verdu P, Georges M, Duquesnoy P, Froment A, Amselem S et al. The role of GHR and IGF1 genes in the genetic determination of African pygmies’ short stature. Eur J Hum Genet. 2013;21(6):653–8. doi: 10.1038/ejhg.2012.223. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 137.Hsieh P, Veeramah KR, Lachance J, Tishkoff SA, Wall JD, Hammer MF et al. Whole-genome sequence analyses of Western Central African Pygmy hunter-gatherers reveal a complex demographic history and identify candidate genes under positive natural selection. Genome Res. 2016;26(3):279–90. doi: 10.1101/gr.192971.115. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 138.Pemberton TJ, Verdu P, Becker NS, Willer CJ, Hewlett BS, Le Bomin S et al. A genome scan for genes underlying adult body size differences between Central African hunter-gatherers and farmers. Hum Genet. 2018;137(6–7):487–509. doi: 10.1007/s00439-018-1902-3. [DOI] [PubMed] [Google Scholar]
  • 139.Migliano AB, Romero IG, Metspalu M, Leavesley M, Pagani L, Antao T et al. Evolution of the pygmy phenotype: evidence of positive selection fro genome-wide scans in African, Asian, and Melanesian pygmies. Hum Biol. 2013;85(1–3):251–84. doi: 10.3378/027.085.0313. [DOI] [PubMed] [Google Scholar]
  • *140.Bergey CM, Lopez M, Harrison GF, Patin E, Cohen JA, Quintana-Murci L et al. Polygenic adaptation and convergent evolution on growth and cardiac genetic pathways in African and Asian rainforest hunter-gatherers. Proc Natl Acad Sci U S A. 2018;115(48):E11256–E63. doi: 10.1073/pnas.1812135115. [DOI] [PMC free article] [PubMed] [Google Scholar]; This paper describes the pattern of positive selection on sets of genes associated with growth and cardiac development in hunter-gatherers population.
  • 141.Berg JJ, Coop G. A population genetic signal of polygenic adaptation. PLoS Genet. 2014;10(8):e1004412. doi: 10.1371/journal.pgen.1004412. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 142.Turchin MC, Chiang CW, Palmer CD, Sankararaman S, Reich D, Genetic Investigation of ATC et al. Evidence of widespread selection on standing variation in Europe at height-associated SNPs. Nat Genet. 2012;44(9):1015–9. doi: 10.1038/ng.2368. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 143.Robinson MR, Hemani G, Medina-Gomez C, Mezzavilla M, Esko T, Shakhbazov K et al. Population genetic differentiation of height and body mass index across Europe. Nat Genet. 2015;47(11):1357–62. doi: 10.1038/ng.3401. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 144.Field Y, Boyle EA, Telis N, Gao Z, Gaulton KJ, Golan D et al. Detection of human adaptation during the past 2000 years. Science. 2016;354(6313):760–4. doi: 10.1126/science.aag0776. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • **145.Sohail M, Maier RM, Ganna A, Bloemendal A, Martin AR, Turchin MC et al. Signals of polygenic adaptation on height have been overestimated due to uncorrected population structure in genome-wide association studies. bioRxiv. 2018. doi: 10.1101/355057. [DOI] [PMC free article] [PubMed] [Google Scholar]; This paper emphasizes that typical constructions of polygenic scores are sensitive to population stratification and that population-level differences should be interpreted with caution.
  • **146.Berg JJ, Harpak A, Sinnott-Armstrong N, Joergensen AM, Mostafavi H, Field Y et al. Reduced signal for polygenic adaptation of height in UK Biobank. eLife. 2019;8. doi: 10.7554/eLife.39725. [DOI] [PMC free article] [PubMed] [Google Scholar]; This paper describes that the detection of polygenic adaptation can be highly prone to the cumulative bias due to uncorrected population structure in GWAS and thus the analysis should be undertaken with great care.
  • 147.Uricchio LH, Kitano HC, Gusev A, Zaitlen NA. An evolutionary compass for elucidating selection mechanisms shaping complex traits. bioRxiv. 2017. doi: 10.1101/173815. [DOI] [Google Scholar]
  • *148.Simons YB, Bullaughey K, Hudson RR, Sella G. A population genetic interpretation of GWAS findings for human quantitative traits. PLoS Biol. 2018;16(3):e2002985. doi: 10.1371/journal.pbio.2002985. [DOI] [PMC free article] [PubMed] [Google Scholar]; This paper describes a model that can help explain the GWAS results for human quantitative traits such as height and body mass index.
  • 149.Zeng J, de Vlaming R, Wu Y, Robinson MR, Lloyd-Jones LR, Yengo L et al. Signatures of negative selection in the genetic architecture of human complex traits. Nat Genet. 2018;50(5):746–53. doi: 10.1038/s41588-018-0101-4. [DOI] [PubMed] [Google Scholar]
  • 150.Gazal S, Finucane HK, Furlotte NA, Loh PR, Palamara PF, Liu X et al. Linkage disequilibrium-dependent architecture of human complex traits shows action of negative selection. Nat Genet. 2017;49(10):1421–7. doi: 10.1038/ng.3954. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 151.Schoech AP, Jordan DM, Loh PR, Gazal S, O’Connor LJ, Balick DJ et al. Quantification of frequency-dependent genetic architectures in 25 UK Biobank traits reveals action of negative selection. Nat Commun. 2019;10(1):790. doi: 10.1038/s41467-019-08424-6. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 152.Collaboration NCDRF. A century of trends in adult human height. eLife. 2016;5. doi: 10.7554/eLife.13410. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 153.Edery P, Marcaillou C, Sahbatou M, Labalme A, Chastang J, Touraine R et al. Association of TALS developmental disorder with defect in minor splicing component U4atac snRNA. Science. 2011;332(6026):240–3. doi: 10.1126/science.1202205. [DOI] [PubMed] [Google Scholar]
  • 154.He H, Liyanarachchi S, Akagi K, Nagy R, Li J, Dietrich RC et al. Mutations in U4atac snRNA, a component of the minor spliceosome, in the developmental disorder MOPD I. Science. 2011;332(6026):238–40. doi: 10.1126/science.1200587. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 155.Rauch A, Thiel CT, Schindler D, Wick U, Crow YJ, Ekici AB et al. Mutations in the pericentrin (PCNT) gene cause primordial dwarfism. Science. 2008;319(5864):816–9. doi: 10.1126/science.1151174. [DOI] [PubMed] [Google Scholar]
  • 156.Jackson GC, Marcus-Soekarman D, Stolte-Dijkstra I, Verrips A, Taylor JA, Briggs MD. Type IX collagen gene mutations can result in multiple epiphyseal dysplasia that is associated with osteochondritis dissecans and a mild myopathy. Am J Med Genet A. 2010;152A(4):863–9. doi: 10.1002/ajmg.a.33240. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 157.Seo SG, Song HR, Kim HW, Yoo WJ, Shim JS, Chung CY et al. Comparison of orthopaedic manifestations of multiple epiphyseal dysplasia caused by MATN3 versus COMP mutations: a case control study. BMC Musculoskelet Disord. 2014;15:84. doi: 10.1186/1471-2474-15-84. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 158.Rossi A, Superti-Furga A. Mutations in the diastrophic dysplasia sulfate transporter (DTDST) gene (SLC26A2): 22 novel mutations, mutation review, associated skeletal phenotypes, and diagnostic relevance. Hum Mutat. 2001;17(3):159–71. doi: 10.1002/humu.1. [DOI] [PubMed] [Google Scholar]
  • 159.Kerns SL, Guevara-Aguirre J, Andrew S, Geng J, Guevara C, Guevara-Aguirre M et al. A novel variant in CDKN1C is associated with intrauterine growth restriction, short stature, and early-adulthood-onset diabetes. J Clin Endocrinol Metab. 2014;99(10):E2117–22. doi: 10.1210/jc.2014-1949. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 160.Doneray H, Aldahmesh M, Yilmaz G, Cinici E, Orbak Z. Infantile Nephropathic Cystinosis: A Novel CTNS Mutation. Eurasian J Med. 2017;49(2):148–51. doi: 10.5152/eurasianjmed.2017.17039. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 161.Seven M, Koparir E, Gezdirici A, Aydin H, Skladny H, Fenercioglu E et al. A novel frameshift mutation and infrequent clinical findings in two cases with Dyggve-Melchior-Clausen syndrome. Clin Dysmorphol. 2014;23(1):1–7. doi: 10.1097/MCD.0000000000000020. [DOI] [PubMed] [Google Scholar]
  • 162.Griffith E, Walker S, Martin CA, Vagnarelli P, Stiff T, Vernay B et al. Mutations in pericentrin cause Seckel syndrome with defective ATR-dependent DNA damage signaling. Nat Genet. 2008;40(2):232–6. doi: 10.1038/ng.2007.80. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 163.Al-Dosari MS, Shaheen R, Colak D, Alkuraya FS. Novel CENPJ mutation causes Seckel syndrome. J Med Genet. 2010;47(6):411–4. doi: 10.1136/jmg.2009.076646. [DOI] [PubMed] [Google Scholar]
  • 164.O’Driscoll M, Ruiz-Perez VL, Woods CG, Jeggo PA, Goodship JA. A splicing mutation affecting expression of ataxia-telangiectasia and Rad3-related protein (ATR) results in Seckel syndrome. Nat Genet. 2003;33(4):497–501. doi: 10.1038/ng1129. [DOI] [PubMed] [Google Scholar]
  • 165.Ogi T, Walker S, Stiff T, Hobson E, Limsirichaikul S, Carpenter G et al. Identification of the first ATRIP-deficient patient and novel mutations in ATR define a clinical spectrum for ATR-ATRIP Seckel Syndrome. PLoS Genet. 2012;8(11):e1002945. doi: 10.1371/journal.pgen.1002945. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 166.Kalay E, Yigit G, Aslan Y, Brown KE, Pohl E, Bicknell LS et al. CEP152 is a genome maintenance protein disrupted in Seckel syndrome. Nat Genet. 2011;43(1):23–6. doi: 10.1038/ng.725. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 167.Qvist P, Huertas P, Jimeno S, Nyegaard M, Hassan MJ, Jackson SP et al. CtIP Mutations Cause Seckel and Jawad Syndromes. PLoS Genet. 2011;7(10):e1002310. doi: 10.1371/journal.pgen.1002310. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 168.Durocher F, Faure R, Labrie Y, Pelletier L, Bouchard I, Laframboise R. A novel mutation in the EIF2AK3 gene with variable expressivity in two patients with Wolcott-Rallison syndrome. Clin Genet. 2006;70(1):34–8. doi: 10.1111/j.1399-0004.2006.00632.x. [DOI] [PubMed] [Google Scholar]
  • 169.Giri N, Batista DL, Alter BP, Stratakis CA. Endocrine abnormalities in patients with Fanconi anemia. J Clin Endocrinol Metab. 2007;92(7):2624–31. doi: 10.1210/jc.2007-0135. [DOI] [PubMed] [Google Scholar]
  • 170.Dimishkovska M, Kotori VM, Gucev Z, Kocheva S, Polenakovic M, Plaseska-Karanfilska D. Novel Founder Mutation in FANCA Gene (c.3446_3449dupCCCT) Among Romani Patients from the Balkan Region. Balkan Med J. 2018;35(1):108–11. doi: 10.4274/balkanmedj.2017.0618. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 171.Kant SG, Cervenkova I, Balek L, Trantirek L, Santen GW, de Vries MC et al. A novel variant of FGFR3 causes proportionate short stature. Eur J Endocrinol. 2015;172(6):763–70. doi: 10.1530/EJE-14-0945. [DOI] [PubMed] [Google Scholar]
  • 172.Quintos JB, Guo MH, Dauber A. Idiopathic short stature due to novel heterozygous mutation of the aggrecan gene. J Pediatr Endocrinol Metab. 2015;28(7–8):927–32. doi: 10.1515/jpem-2014-0450. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 173.van der Steen M, Pfundt R, Maas S, Bakker-van Waarde WM, Odink RJ, Hokken-Koelega ACS. ACAN Gene Mutations in Short Children Born SGA and Response to Growth Hormone Treatment. J Clin Endocrinol Metab. 2017;102(5):1458–67. doi: 10.1210/jc.2016-2941. [DOI] [PubMed] [Google Scholar]
  • 174.Dateki S ACAN mutations as a cause of familial short stature. Clin Pediatr Endocrinol. 2017;26(3):119–25. doi: 10.1297/cpe.26.119. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 175.Wang X, Xiao F, Yang Q, Liang B, Tang Z, Jiang L et al. A novel mutation in GDF5 causes autosomal dominant symphalangism in two Chinese families. Am J Med Genet A. 2006;140A(17):1846–53. doi: 10.1002/ajmg.a.31372. [DOI] [PubMed] [Google Scholar]
  • 176.Yang W, Cao L, Liu W, Jiang L, Sun M, Zhang D et al. Novel point mutations in GDF5 associated with two distinct limb malformations in Chinese: brachydactyly type C and proximal symphalangism. J Hum Genet. 2008;53(4):368–74. doi: 10.1007/s10038-008-0253-7. [DOI] [PubMed] [Google Scholar]
  • 177.Bicknell LS, Bongers EM, Leitch A, Brown S, Schoots J, Harley ME et al. Mutations in the pre-replication complex cause Meier-Gorlin syndrome. Nat Genet. 2011;43(4):356–9. doi: 10.1038/ng.775. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 178.Guernsey DL, Matsuoka M, Jiang H, Evans S, Macgillivray C, Nightingale M et al. Mutations in origin recognition complex gene ORC4 cause Meier-Gorlin syndrome. Nat Genet. 2011;43(4):360–4. doi: 10.1038/ng.777. [DOI] [PubMed] [Google Scholar]
  • 179.Silver HK, Kiyasu W, George J, Deamer WC. Syndrome of congenital hemihypertrophy, shortness of stature, and elevated urinary gonadotropins. Pediatrics. 1953;12(4):368–76. [PubMed] [Google Scholar]
  • 180.Liu D, Wang Y, Yang XA, Liu D. De Novo Mutation of Paternal IGF2 Gene Causing Silver-Russell Syndrome in a Sporadic Patient. Front Genet. 2017;8:105. doi: 10.3389/fgene.2017.00105. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 181.Laron Z Insulin-like growth factor-I treatment of children with Laron syndrome (primary growth hormone insensitivity). Pediatr Endocrinol Rev. 2008;5(3):766–71. [PubMed] [Google Scholar]
  • 182.Racacho L, Byrnes AM, MacDonald H, Dranse HJ, Nikkel SM, Allanson J et al. Two novel disease-causing variants in BMPR1B are associated with brachydactyly type A1. Eur J Hum Genet. 2015;23(12):1640–5. doi: 10.1038/ejhg.2015.38. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 183.Hellemans J, Mortier GR. IHH, Acrocapitofemoral Dysplasia, and Brachydactyly A1. Epstein’s Inborn Errors of Development: The Molecular Basis of Clinical Disorders of Morphogenesis. 2016;293:1–9. doi: 10.1093/med/9780199934522.003.0030. [DOI] [Google Scholar]
  • 184.Bacino CA. ROR2-Related Robinow Syndrome. In: Adam MP, Ardinger HH, Pagon RA, Wallace SE, Bean LJH, Stephens K et al. , editors. GeneReviews((R)) Seattle (WA)1993. [PubMed] [Google Scholar]
  • 185.White JJ, Mazzeu JF, Hoischen A, Bayram Y, Withers M, Gezdirici A et al. DVL3 Alleles Resulting in a −1 Frameshift of the Last Exon Mediate Autosomal-Dominant Robinow Syndrome. American journal of human genetics. 2016;98(3):553–61. doi: 10.1016/j.ajhg.2016.01.005. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 186.Person AD, Beiraghi S, Sieben CM, Hermanson S, Neumann AN, Robu ME et al. WNT5A mutations in patients with autosomal dominant Robinow syndrome. Dev Dyn. 2010;239(1):327–37. doi: 10.1002/dvdy.22156. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 187.Bunn KJ, Daniel P, Rosken HS, O’Neill AC, Cameron-Christie SR, Morgan T et al. Mutations in DVL1 cause an osteosclerotic form of Robinow syndrome. American journal of human genetics. 2015;96(4):623–30. doi: 10.1016/j.ajhg.2015.02.010. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 188.Rappold GA, Fukami M, Niesler B, Schiller S, Zumkeller W, Bettendorf M et al. Deletions of the homeobox gene SHOX (short stature homeobox) are an important cause of growth failure in children with short stature. J Clin Endocrinol Metab. 2002;87(3):1402–6. doi: 10.1210/jcem.87.3.8328. [DOI] [PubMed] [Google Scholar]
  • 189.Tartaglia M, Kalidas K, Shaw A, Song X, Musat DL, van der Burgt I et al. PTPN11 mutations in Noonan syndrome: molecular spectrum, genotype-phenotype correlation, and phenotypic heterogeneity. American journal of human genetics. 2002;70(6):1555–63. doi: 10.1086/340847. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 190.Tartaglia M, Pennacchio LA, Zhao C, Yadav KK, Fodale V, Sarkozy A et al. Gain-of-function SOS1 mutations cause a distinctive form of Noonan syndrome. Nat Genet. 2007;39(1):75–9. doi: 10.1038/ng1939. [DOI] [PubMed] [Google Scholar]
  • 191.Razzaque MA, Nishizawa T, Komoike Y, Yagi H, Furutani M, Amo R et al. Germline gain-of-function mutations in RAF1 cause Noonan syndrome. Nat Genet. 2007;39(8):1013–7. doi: 10.1038/ng2078. [DOI] [PubMed] [Google Scholar]
  • 192.Kouz K, Lissewski C, Spranger S, Mitter D, Riess A, Lopez-Gonzalez V et al. Genotype and phenotype in patients with Noonan syndrome and a RIT1 mutation. Genet Med. 2016;18(12):1226–34. doi: 10.1038/gim.2016.32. [DOI] [PubMed] [Google Scholar]
  • 193.Aoki Y, Niihori T, Kawame H, Kurosawa K, Ohashi H, Tanaka Y et al. Germline mutations in HRAS protooncogene cause Costello syndrome. Nat Genet. 2005;37(10):1038–40. doi: 10.1038/ng1641. [DOI] [PubMed] [Google Scholar]
  • 194.Lv Y, Zhu L, Zheng J, Wu D, Shao J. Growth Concerns in Coffin-Lowry Syndrome: A Case Report and Literature Review. Front Pediatr. 2018;6:430. doi: 10.3389/fped.2018.00430. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 195.Wormser O, Gradstein L, Yogev Y, Perez Y, Kadir R, Goliand I et al. SCAPER localizes to primary cilia and its mutation affects cilia length, causing Bardet-Biedl syndrome. Eur J Hum Genet. 2019. doi: 10.1038/s41431-019-0347-z. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 196.Canda MT, Doganay Caglayan L, Demir AB, Demir N. Prenatal detection of Peters plus-like syndrome. Turk J Obstet Gynecol. 2018;15(4):273–6. doi: 10.4274/tjod.45649. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 197.Jacob M, Menon S, Botti C, Marshall I. Heterozygous NPR2 Mutation in Two Family Members with Short Stature and Skeletal Dysplasia. Case Rep Endocrinol. 2018;2018:7658496. doi: 10.1155/2018/7658496. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 198.Huber C, Dias-Santagata D, Glaser A, O’Sullivan J, Brauner R, Wu K et al. Identification of mutations in CUL7 in 3-M syndrome. Nat Genet. 2005;37(10):1119–24. doi: 10.1038/ng1628. [DOI] [PubMed] [Google Scholar]
  • 199.Huber C, Fradin M, Edouard T, Le Merrer M, Alanay Y, Da Silva DB et al. OBSL1 mutations in 3-M syndrome are associated with a modulation of IGFBP2 and IGFBP5 expression levels. Hum Mutat. 2010;31(1):20–6. doi: 10.1002/humu.21150. [DOI] [PubMed] [Google Scholar]
  • 200.Hanson D, Murray PG, O’Sullivan J, Urquhart J, Daly S, Bhaskar SS et al. Exome sequencing identifies CCDC8 mutations in 3-M syndrome, suggesting that CCDC8 contributes in a pathway with CUL7 and OBSL1 to control human growth. American journal of human genetics. 2011;89(1):148–53. doi: 10.1016/j.ajhg.2011.05.028. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 201.Akiyama H, Lefebvre V. Unraveling the transcriptional regulatory machinery in chondrogenesis. J Bone Miner Metab. 2011;29(4):390–5. doi: 10.1007/s00774-011-0273-9. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 202.Renes JS, Knijnenburg J, Chitoe-Ramawadhdoebe S, Gille JJP, de Bruin C, Barge-Schaapveld D. Possible hints and pitfalls in diagnosing Peutz-Jeghers syndrome. J Pediatr Endocrinol Metab. 2018;31(12):1381–6. doi: 10.1515/jpem-2018-0265. [DOI] [PubMed] [Google Scholar]
  • 203.Lin Y, van Duyvenvoorde HA, Liu H, Yang C, Warsito D, Yin C et al. Characterization of an activating R1353H insulin-like growth factor 1 receptor variant in a male with extreme tall height. Eur J Endocrinol. 2018;179(2):85–95. doi: 10.1530/EJE-18-0176. [DOI] [PubMed] [Google Scholar]
  • 204.Suri T, Dixit A. The phenotype of EZH2 haploinsufficiency-1.2-Mb deletion at 7q36.1 in a child with tall stature and intellectual disability. Am J Med Genet A. 2017;173(10):2731–5. doi: 10.1002/ajmg.a.38356. [DOI] [PubMed] [Google Scholar]
  • 205.Nam HK, Nam MH, Ha KS, Rhie YJ, Lee KH. A Novel Fibrillin-1 Gene Mutation Leading to Marfan Syndrome in a Korean Girl. Ann Clin Lab Sci. 2017;47(2):221–5. [PubMed] [Google Scholar]
  • 206.Xin B, Cruz Marino T, Szekely J, Leblanc J, Cechner K, Sency V et al. Novel DNMT3A germline mutations are associated with inherited Tatton-Brown-Rahman syndrome. Clin Genet. 2017;91(4):623–8. doi: 10.1111/cge.12878. [DOI] [PubMed] [Google Scholar]
  • 207.Dong HY, Zeng H, Hu YQ, Xie L, Wang J, Wang XY et al. 19p13.2 Microdeletion including NFIX associated with overgrowth and intellectual disability suggestive of Malan syndrome. Mol Cytogenet. 2016;9:71. doi: 10.1186/s13039-016-0282-4. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 208.Tatton-Brown K, Rahman N. Sotos syndrome. Eur J Hum Genet. 2007;15(3):264–71. doi: 10.1038/sj.ejhg.5201686. [DOI] [PubMed] [Google Scholar]
  • 209.Toydemir RM, Brassington AE, Bayrak-Toydemir P, Krakowiak PA, Jorde LB, Whitby FG et al. A novel mutation in FGFR3 causes camptodactyly, tall stature, and hearing loss (CATSHL) syndrome. American journal of human genetics. 2006;79(5):935–41. doi: 10.1086/508433. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 210.Gagliardi L, Scott HS, Feng J, Torpy DJ. A case of Aromatase deficiency due to a novel CYP19A1 mutation. BMC Endocr Disord. 2014;14:16. doi: 10.1186/1472-6823-14-16. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 211.Doulla M, McIntyre AD, Hegele RA, Gallego PH. A novel MC4R mutation associated with childhood-onset obesity: A case report. Paediatr Child Health. 2014;19(10):515–8. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 212.Shimizu R, Mitsui N, Mori Y, Cho S, Yamamori S, Osawa M et al. Cryptic 17q22 deletion in a boy with a t(10;17)(p15.3;q22) translocation, multiple synostosis syndrome 1, and hypogonadotropic hypogonadism. Am J Med Genet A. 2008;146A(11):1458–61. doi: 10.1002/ajmg.a.32319. [DOI] [PubMed] [Google Scholar]
  • 213.Abali ZY, Yesil G, Kirkgoz T, Kaygusuz SB, Eltan M, Turan S et al. Evaluation of growth and puberty in a child with a novel TBX19 gene mutation and review of the literature. Hormones (Athens). 2019. doi: 10.1007/s42000-019-00096-7. [DOI] [PubMed] [Google Scholar]

RESOURCES