Hostname: page-component-848d4c4894-pftt2 Total loading time: 0 Render date: 2024-05-23T17:25:00.830Z Has data issue: false hasContentIssue false

Genetics of osteoporosis

Symposium on ‘Genetic polymorphisms and disease risk’

Published online by Cambridge University Press:  30 April 2007

Stuart H. Ralston
Affiliation:
Molecular Medicine Centre, Rheumatic Diseases Unit, Edinburgh University, Western General Hospital, Edinburgh EH4 2XU, UK
Rights & Permissions [Opens in a new window]

Abstract

Osteoporosis is a common disease with a strong genetic component characterised by reduced bone mass and an increased risk of fragility fractures. Twin and family studies have shown that genetic factors contribute to osteoporosis by influencing bone mineral density (BMD), and other phenotypes that are associated with fracture risk, although the heritability of fracture itself is modest. Linkage studies have identified several quantitative trait loci that regulate BMD but most causal genes remain to be identified. In contrast, linkage studies in monogenic bone diseases have been successful in gene identification, and polymorphisms in many of these genes have been found to contribute to the regulation of bone mass in the normal population. Population-based studies have identified polymorphisms in several candidate genes that have been associated with bone mass or osteoporotic fracture, although individually these polymorphisms only account for a small amount of the genetic contribution to BMD regulation. Environmental factors such as diet and physical activity are also important determinants of BMD, and in some cases specific nutrients have been found to interact with genetic polymorphisms to regulate BMD. From a clinical standpoint, advances in knowledge about the genetic basis of osteoporosis are likely to be important in increasing the understanding of the pathophysiology of the disease; providing new genetic markers with which to assess fracture risk and in identifying genes and pathways that form molecular targets for the design of the next generation of drug treatments.

Type
Research Article
Copyright
Copyright © The Author 2007

Abbreviations:
BMD

bone mineral density

LRP

lipoprotein receptor-related protein

QTL

quantitative trait loci

VDR

vitamin D receptor

Genetic factors play an important role in the pathogenesis of osteoporosis. Twin- and family-based studies have indicated that 60–85% of the variance in bone mineral density (BMD) is genetically determined (Krall & Dawson-Hughes, Reference Krall and Dawson-Hughes1993; Gueguen et al. Reference Gueguen, Jouanny, Guillemin, Kuntz, Pourel and Siest1995), and other risk factors for osteoporotic fractures, such as quantitative ultrasound properties of bone, femoral neck geometry and bone turnover markers range, have also been shown to have a strong heritable component (Arden et al. Reference Arden, Baker, Hogg, Baan and Spector1996; Garnero et al. Reference Garnero, Arden, Griffiths, Delmas and Spector1996). Family history of fracture has been shown in several studies to be a risk factor for fractures independently of BMD (Cummings et al. Reference Cummings, Nevitt, Browner, Stone, Fox, Ensrud, Cauley, Black and Vogt1995; Torgerson et al. Reference Torgerson, Campbell, Thomas and Reid1996), and in keeping with this finding several investigators (Deng et al. Reference Deng, Chen, Recker, Stegman, Li, Davies, Zhou, Deng, Heaney and Recker2000; Andrew et al. Reference Andrew, Antioniades, Scurrah, MacGregor and Spector2005) have reported that fracture may have a heritable component. In one study of post-menopausal women (Deng et al. Reference Deng, Chen, Recker, Stegman, Li, Davies, Zhou, Deng, Heaney and Recker2000) heritability of wrist fracture was estimated as about 25%, whereas another study of twins (Andrew et al. Reference Andrew, Antioniades, Scurrah, MacGregor and Spector2005) has suggested that the heritability of wrist fracture may be as much as 54%. Interestingly, the heritability of wrist fracture in both these studies was shown to be largely independent of BMD, suggesting that predisposition may have been mediated through genetic influences on other factors such as bone turnover, bone geometry or even perhaps the risk of falling. However, another study (Kannus et al. Reference Kannus, Palvanen, Kaprio, Parkkari and Koskenvuo1999) has failed to detect any evidence for heritability of fractures in elderly twins. These divergent results are probably explained by the fact that the heritability of fracture decreases with age, as environmental factors become more important. This relationship has been demonstrated in a large study of Swedish twins (Michaelsson et al. Reference Michaelsson, Melhus, Ferm, Ahlbom and Pedersen2005), which has shown that the heritability of hip fracture is about 68% in those aged <65 years but drops off rapidly with age to reach a value of almost zero by the eighth decade. This finding illustrates that identifying genes that are related to risk factors for osteoporosis such as BMD does not necessarily mean that these genes will influence the risk of fracture.

Human linkage studies

Several genome-wide linkage scans have been performed to try to identify loci that regulate BMD. The most important quantitative trait loci (QTL) for BMD identified by these studies are summarised in Table 1. Few of the genome-wide scans so far performed have identified QTL that meet the criteria for genome-wide significance, and only one gene that regulates susceptibility to osteoporosis has been identified by this approach; the BMP2 gene that encodes bone morphogenic protein 2, an important regulator of osteoblast differentiation (Styrkarsdottir et al. Reference Styrkarsdottir, Cazier, Kong, Rolfsson, Larsen and Bjarnadottir2003). Several important findings have emerged from these studies. First, it appears that the genes that regulate BMD probably do so in a site-specific and gender-specific manner (Peacock et al. Reference Peacock, Koller, Fishburn, Krishnan, Lai, Hui, Johnston, Foroud and Econs2005; Ralston et al. Reference Ralston, Galwey, Mackay, Albagha, Cardon and Compston2005); second, it appears that the genes that regulate peak bone mass differ from those that regulate BMD in older individuals (Kammerer et al. Reference Kammerer, Schneider, Cole, Hixson, Samollow and O'Connell2003; Karasik et al. Reference Karasik, Cupples, Hannan and Kiel2003; Ralston et al. Reference Ralston, Galwey, Mackay, Albagha, Cardon and Compston2005). The lack of replication of linkage peaks between studies mirrors experience in other complex diseases and might reflect the fact that genes that regulate BMD differ in different populations or that genes that predispose to osteoporosis have modest effects that are difficult to detect by conventional linkage analysis. Technical advances such as the development of densely-spaced panels of single-nucleotide polymorphisms for genome-wide scans are likely to improve the chances of detecting genes of modest effect size in the future (Sawcer et al. Reference Sawcer, Maranian, Singlehurst, Yeo, Compston and Daly2004). There is also a prospect that meta-analysis of genome-wide scans may reveal significant QTL that have not been detected by individual studies (Fisher et al. Reference Fisher, Lanchbury and Lewis2003).

Table 1. Quantitative trait loci for bone mineral density detected by genome-wide linkage scanFootnote *

cM, centimorgans; Fem, femoral.

* The loci shown are those identified by genome-wide scan for which the LOD score exceeded +2·2. The LOD score is the logarithm of the odds that the disease gene and the marker being studied are linked. For complex diseases linkage is considered significant when the LOD score exceeds +3·6, whereas linkage is considered suggestive when the LOD score exceeds +2·2.

A measure of the physical distance between the locus identified and the telomere (tip) of the chromosome.

Interest has also focused on identifying QTL for the regulation of other phenotypes relevant to the pathogenesis of osteoporosis. For example, genome-wide linkage scans have identified several QTL that strongly influence hip geometry (Koller et al. Reference Koller, Liu, Econs, Hui, Morin and Joslyn2001), while others have been performed that have identified QTL that regulate the quantitative ultrasound properties of bone (Wilson et al. Reference Wilson, Reed, Andrew, Barber, Lindersson and Langdown2004).

Animal linkage studies

Linkage studies in mice (Klein et al. Reference Klein, Mitchell, Phillips, Belknap and Orwoll1998; Beamer et al. Reference Beamer, Shultz, Donahue, Churchill, Sen, Wergedal, Baylink and Rosen2001), rats (Koller et al. Reference Koller, Alam, Sun, Liu, Fishburn, Carr, Econs, Foroud and Turner2005) and primates (Mahaney et al. Reference Mahaney, Morin, Rodriguez, Newman and Rogers1997) have resulted in the identification of several QTL that regulate BMD. Linkage analysis has also been used to localise QTL for other osteoporosis-related phenotypes such as bone structure, bone shape and bone strength (Turner et al. Reference Turner, Sun, Schriefer, Pitner, Price, Bouxsein, Rosen, Donahue, Shultz and Beamer2003; Alam et al. Reference Alam, Sun, Liu, Koller, Fishburn, Carr, Econs, Foroud and Turner2005) and circulating levels of insulin-like growth factor-1 (Bouxsein et al. Reference Bouxsein, Rosen, Turner, Ackert, Shultz and Donahue2002). Loci for the regulation of BMD have now been identified on almost all mouse chromosomes, and several rat chromosomes with replication of some QTL across different strains, and replication of some human BMD QTL (Koller et al. Reference Koller, Alam, Sun, Liu, Fishburn, Carr, Econs, Foroud and Turner2005). These studies have also shown that the genes that regulate BMD in mice have effects that are site-specific and gender-specific (Beamer et al. Reference Beamer, Shultz, Donahue, Churchill, Sen, Wergedal, Baylink and Rosen2001; Orwoll et al. Reference Orwoll, Belknap and Klein2001). To date, only one gene that regulates BMD, the Alox15 gene, has been identified, by studies in mice (Klein et al. Reference Klein, Allard, Avnur, Nikolcheva, Rotstein, Carlos, Shea, Waters, Belknap, Peltz and Orwoll2004). In this study a QTL for the regulation of BMD was identified on mouse chromosome 11 by linkage in a cross of DBA/2 and C57BL/6 mice, and subsequent microarray analysis has shown that the parental DBA2 strain of mice (low BMD) has a 20-fold increase in expression of the Alox15 mRNA when compared with C57BL/6 (high BMD) mice. From this observation the authors had suspected that Alox15 might act as a negative regulator of bone mass and they have confirmed this hypothesis by finding that Alox15-knock-out mice have increased BMD and that inhibition of Alox15 protects against ovariectomy-induced bone loss. The mechanism by which Alox15 reduces BMD is unclear, but the gene encodes a lipoxygenase enzyme that converts arachidonic and linoleic acids into ligands for the transcription factor PPARγ, which is thought to regulate differentiation of mesenchymal cells into adipocytes and osteoblasts. Recent studies have shown that genetic variation in a human homologue of Alox15 accounts for some of the heritable component of spine BMD regulation in man (Ichikawa et al. Reference Ichikawa, Koller, Johnson, Lai, Xuei and Edenberg2006).

Candidate gene studies

Candidate gene association studies have identified several polymorphisms that are associated with BMD, bone loss or osteoporotic fractures. Some of the most important candidate genes that have been implicated in the pathogenesis of osteoporosis will be discussed.

Vitamin D receptor

The vitamin D receptor (VDR) was the first candidate gene to be studied in relation to BMD regulation and most attention has focused on polymorphisms situated on the 3′ flank of VDR recognised by the restriction enzymes BsmI, ApaI and TaqI. A meta-analysis of association studies that have genotyped the BsmI polymorphism have concluded that there is evidence of an association between spine BMD and the BsmI polymorphism, equivalent to approximately 0·15 Z-score units, between the BB genotype and the other genotypes (Thakkinstian et al. Reference Thakkinstian, D'Este, Eisman, Nguyen and Attia2004). Another polymorphism affecting exon 2 of VDR has been identified that creates an alternative translational start site, resulting in the production of two isoforms of the VDR protein that differ in length by three amino acids (Gross et al. Reference Gross, Eccleshall, Malloy, Villa, Marcus and Feldman1997). This polymorphism has been associated with BMD in some studies, but functional studies have yielded inconclusive results (Gross et al. Reference Gross, Krishnan, Malloy, Eccleshall, Zhao and Feldman1998). A polymorphism has been identified in the promoter of VDR at a binding site for the transcription factor Cdx-2 that has been associated with BMD in Japanese subjects and appears to be functional (Arai et al. Reference Arai, Miyamoto, Yoshida, Yamamoto, Taketani and Morita2001). This polymorphism has been associated with fracture in other populations, but not with BMD (Fang et al. Reference Fang, van Meurs, Bergink, Hofman, van Duijn, van Leeuwen, Pols and Uitterlinden2003). The most comprehensive study of VDR alleles in relation to osteoporosis has been in the Rotterdam study in which haplotype-tagging single-nucleotide polymorphisms of VDR were analysed in 6418 subjects (Fang et al. Reference Fang, van Meurs, D'Alesio, Jhamai, Zhao, Rivadeneira, Hofman, van Leeuwen, Jehan, Pols and Uitterlinden2005). Alleles were identified in the promoter region and 3′ untranslated region that were shown to be associated with an increased risk of fracture, and a subgroup of individuals who carried risk alleles at both sites were found to have an increased risk of fracture when compared with control subjects. Functional studies have also shown that the promoter haplotype that increases fracture risk is associated with reduced VDR expression in reporter assays whereas the 3′ UTR risk haplotype is associated with increased degradation of VDR mRNA. The data would be consistent with a model whereby the combination of risk haplotypes results in a lower VDR mRNA level as a result of decreased transcription and increased degradation. Interestingly, the risk alleles for fracture identified in this study were not associated with differences in BMD. In view of this finding, the mechanism by which these polymorphisms predispose to fracture is unclear. Moreover, if correction had been applied for all the combinations of VDR haplotypes tested in this study in relation to fracture, the association would not have been significant.

Collagen type I αI

COLIA1, the gene encoding the αI chain of type I collagen is an important functional candidate for the pathogenesis of osteoporosis, as type I collagen is the major protein of bone. Extensive studies have been conducted on a polymorphism that lies within intron 1 of the COLIA1 gene at a Sp1 binding site (Grant et al. Reference Grant, Reid, Blake, Herd, Fogelman and Ralston1996). The thymidine-containing allele of this polymorphism has been associated with reduced bone density (Grant et al. Reference Grant, Reid, Blake, Herd, Fogelman and Ralston1996; Uitterlinden et al. Reference Uitterlinden, Burger, Huang, Yue, McGuigan, Grant, Hofman, van Leeuwen, Pols and Ralston1998) and other osteoporosis-related phenotypes such as post-menopausal bone loss (Harris et al. Reference Harris, Patel, Cole and Dawson-Hughes2000; MacDonald et al. Reference MacDonald, McGuigan, New, Campbell, Golden, Ralston and Reid2001), bone geometry (Qureshi et al. Reference Qureshi, McGuigan, Seymour, Hutchison, Reid and Ralston2001), bone quality (Mann et al. Reference Mann, Hobson, Li, Stewart, Grant, Robins, Aspden and Ralston2001) and bone mineralization (Stewart et al. Reference Stewart, Roschger, Misof, Mann, Fratzl, Klaushofer, Aspden and Ralston2005). Functional analysis has shown that the osteoporosis-associated T allele (‘s’) of the Sp1 polymorphism is associated with increased DNA–protein binding, increased transcription from the T allele and abnormally increased production of the collagen type I α1 mRNA and protein (Mann et al. Reference Mann, Hobson, Li, Stewart, Grant, Robins, Aspden and Ralston2001). It is thought that the resulting imbalance between the αI and α2 chains of collagen type I may contribute to impairment of bone strength and reduced bone mass in carriers of the T allele by subtly affecting bone mineralization (Stewart et al. Reference Stewart, Roschger, Misof, Mann, Fratzl, Klaushofer, Aspden and Ralston2005). Meta-analyses of published studies (Efstathiadou et al. Reference Efstathiadou, Tsatsoulis and Ioannidis2001; Mann et al. Reference Mann, Hobson, Li, Stewart, Grant, Robins, Aspden and Ralston2001; Mann & Ralston, Reference Mann and Ralston2003) have concluded that carriage of the T allele is associated with reduced BMD at the lumbar spine and femoral neck and with vertebral fractures. More recently, two polymorphisms (–1997G/T and –1663delT) have been identified in the COLIA1 promoter region and have been associated with BMD (Garcia-Giralt et al. Reference Garcia-Giralt, Nogues, Enjuanes, Puig, Mellibovsky, Bay-Jensen, Carreras, Balcells, Diez-Perez and Grinberg2002). These polymorphisms are in linkage disequilibrium with the Sp1 polymorphism, and functional studies have shown that the polymorphisms influence COLIA1 transcription in promoter–reporter assays (Garcia-Giralt et al. Reference Garcia-Giralt, Enjuanes, Bustamante, Mellibovsky, Nogues, Carreras, Diez-Perez, Grinberg and Balcells2005). The –1997G/T promoter polymorphism has been studied in relation to BMD in other populations and in family-based studies (Liu et al. Reference Liu, Lu, Long, Xu, Shen, Recker and Deng2004; Yamada et al. Reference Yamada, Ando, Niino and Shimokata2005; Zhang et al. Reference Zhang, Lei, Mo, Wang, Li and Deng2005) with mixed results, although most of these studies have been of limited sample size. Current evidence indicates that the promoter polymorphisms and the Sp1 polymorphism interact to regulate BMD in women (Stewart et al. Reference Stewart, Jin, McGuigan, Albagha, Garcia-Giralt, Bassiti, Grinberg, Balcells, Reid and Ralston2006), indicating that the previously-reported associations between the Sp1 polymorphism and osteoporosis-related phenotypes may in fact be driven by an extended haplotype involving the Sp1 and promoter polymorphisms.

Oestrogen receptor α

The oestrogen receptor α, encoded by the ESR1 gene, is another important functional candidate for the regulation of bone mass. A large number of investigators have looked for evidence of an association between ESR1 alleles and BMD, mostly focusing on two polymorphisms within intron 1, recognised by the XbaI and PvuII restriction enzymes. There is some evidence to suggest that these polymorphisms regulate ESR1 transcription and that they may therefore be functionally important (Herrington et al. Reference Herrington, Howard, Brosnihan, McDonnell, Li, Hawkins, Reboussin, Xu, Zheng, Meyers and Bleecker2002). A meta-analysis of published studies performed up until 2001 (Ioannidis et al. Reference Ioannidis, Stavrou, Trikalinos, Zois, Brandi, Gennari, Albagha, Ralston and Tsatsoulis2002) has shown an association between the XbaI polymorphism, BMD and fractures, with higher BMD values and reduced fracture risk in ‘XX’ homozygotes. Recently, a large prospective analysis in the Genetic Markers for Osteoporosis Project has confirmed that XX homozygotes have a reduced risk of fracture (Ioannidis et al. Reference Ioannidis, Ralston, Bennett, Brandi, Grinberg and Karassa2004), but no association with BMD was observed, indicating that ESR1 might influence fracture risk by mechanisms that are independent of BMD. One possible mechanism might be through effects on bone quality or bone turnover, since ESR1 alleles have recently been associated with ultrasound properties of bone and rates of post-menopausal bone loss (Albagha et al. Reference Albagha, Pettersson, Stewart, McGuigan, MacDonald, Reid and Ralston2005)

Transforming growth factor β1

Several polymorphisms of the TGFB1 gene that encodes the growth factor transforming growth factor β-1 have been identified and some of them have been associated with BMD and/or osteoporotic fracture in various studies (Langdahl et al. Reference Langdahl, Knudsen, Jensen, Gregersen and Eriksen1997; Yamada et al. Reference Yamada, Miyauchi, Takagi, Tanaka, Mizuno and Harada2001). The best functional candidate is a C/T polymorphism that causes a proline to leucine amino acid substitution at position 10 in the transforming growth factor β-1 signal peptide that has been associated with circulating transforming growth factor β-1 levels. However, this polymorphism is in strong linkage disequilibrium within other polymorphisms in the TGFB1 promoter and effects on transcription are also possible (Shah et al. Reference Shah, Rahaman, Hurley and Posch2005). Although many association studies have been performed, most are of limited sample size and further large-scale studies will be required to confirm or refute the status of TGFB1 as a true susceptibility gene for osteoporosis.

Lipoprotein receptor-related protein 5

Inactivating mutations of the lipoprotein receptor-related protein (LRP) 5 gene are the cause of the rare recessive disorder osteoporosis pseudoglioma syndrome (Gong et al. Reference Gong, Slee, Fukai, Rawadi, Roman-Roman and Reginato2001), whereas activating mutations in the same gene cause autosomal dominant inheritance of high bone mass (Little et al. Reference Little, Carulli, Del Mastro, Dupuis, Osborne and Folz2002). The involvement of LRP5 in these rare monogenic bone disorders led several investigators to evaluate the role of LRP5 as a candidate gene for BMD regulation in the normal population. Six studies have now been published showing evidence of an allelic association between polymorphisms in LRP5 and BMD (Ferrari et al. Reference Ferrari, Deutsch, Choudhury, Chevalley, Bonjour, Dermitzakis, Rizzoli and Antonarakis2004; Koay et al. Reference Koay, Woon, Zhang, Miles, Duncan and Ralston2004; Koh et al. Reference Koh, Jung, Hong, Park, Chang, Shin, Kim and Kim2004; Mizuguchi et al. Reference Mizuguchi, Furuta, Watanabe, Tsukamoto, Tomita and Tsujihata2004; Urano et al. Reference Urano, Shiraki, Ezura, Fujita, Sekine, Hoshino, Hosoi, Orimo, Emi, Ouchi and Inoue2004; van Meurs et al. Reference van Meurs, Rivadeneira, Jhamai, Hugens, Hofman, van Leeuwen, Pols and Uitterlinden2006). Many variants have been studied, but the most likely functional candidate is an alanine to valine amino acid substitution at position 1330 (A1330V). The mechanism by which this variant affects LRP5 signalling has not been investigated, but evidence of an interaction between the LRP5 A1330V variant and a coding polymorphism of LRP6 (1062V) has been gained in the Rotterdam study (van Meurs et al. Reference van Meurs, Rivadeneira, Jhamai, Hugens, Hofman, van Leeuwen, Pols and Uitterlinden2006), in which polymorphisms of both genes were found to interact to affect fracture susceptibility. One consistent feature to emerge from these studies is that the association between LRP5 alleles and BMD is stronger in males (Ferrari et al. Reference Ferrari, Deutsch, Choudhury, Chevalley, Bonjour, Dermitzakis, Rizzoli and Antonarakis2004; Koay et al. Reference Koay, Woon, Zhang, Miles, Duncan and Ralston2004), which suggests that LRP5 may regulate bone mass in a gender-specific manner.

Sclerostin

Mutations affecting the SOST gene, which encodes sclerostin, are the cause of the sclerosing bone dysplasias Van Buchem disease and sclerosteosis (Balemans et al. Reference Balemans, Ebeling, Patel, Van Hul, Olson and Dioszegi2001, Reference Balemans, Foernzler, Parsons, Ebeling, Thompson, Reid, Lindpaintner, Ralston and Van Hul2002b; Brunkow et al. Reference Brunkow, Gardner, Van Ness, Paeper, Kovacevich and Proll2001). Polymorphisms of SOST have been evaluated in relation to BMD in two studies. In one study (Balemans et al. Reference Balemans, Foernzler, Parsons, Ebeling, Thompson, Reid, Lindpaintner, Ralston and Van Hul2002a) no association between SOST polymorphism and BMD was found in peri-menopausal women using a case–control design, whereas in another study of older women (Uitterlinden et al. Reference Uitterlinden, Arp, Paeper, Charmley, Proll and Rivadeneira2004) evidence of an association with BMD was observed in men and women, with effects that increase with age. These data suggest that SOST polymorphisms may regulate BMD, especially in older individuals.

TCIRG1

The TCIRG1 gene encodes the APT6i subunit of the osteoclast-specific proton pump (Frattini et al. Reference Frattini, Orchard, Sobacchi, Giliani, Abinun and Mattsson2000). Inactivating mutations in TCIRGI are responsible for a subset of patients with recessive osteopetrosis. Recent work indicates that polymorphisms of TCIRG1 might contribute to regulation of BMD in the normal population; a study by Frattini (Sobacchi et al. Reference Sobacchi, Vezzoni, Reid, McGuigan, Frattini, Mirolo, Albhaga, Musio, Villa and Ralston2004) has shown evidence of an association between a polymorphism affecting an activator protein 1-binding site in the TCIRG1 promoter and BMD in peri-menopausal women. However, functional studies need to be performed to identify the mechanisms that underlie this association and to replicate the finding in other populations.

CLCN7

The CLCN7 gene encodes a chloride channel that is highly expressed in osteoclasts and essential for acidification of the resorption lacuna. Homozygous inactivation mutations in CLCN7 cause a severe form of recessive osteopetrosis whereas heterozygous missense mutations of CLCN7 cause autosomal dominant osteopetrosis (Balemans et al. Reference Balemans, Van Wesenbeeck and Van Hul2005). Prompted by this observation Pettersson et al. (Reference Pettersson, Albagha, Mirolo, Taranta, Frattini and McGuigan2005) have looked for evidence of an association between polymorphisms of CLCN7 and BMD in normal individuals and have found that a common polymorphism in exon 15 of CLCN7 that results in a methionine to valine amino acid change is associated with BMD in normal women. Further studies will be required to determine whether this polymorphism is functionally important and to replicate the observation in other populations.

Gene–environment interactions

Several studies have been performed in which interactions have been sought between environmental factors and polymorphic variants in candidate genes. For the most part, these studies have been underpowered and conflicting results have been obtained. The most widely studied gene–environment interaction is between VDR alleles and Ca or vitamin D intake. An early study (Krall et al. Reference Krall, Parry, Lichter and Dawson-Hughes1995) has shown faster rates of post-menopausal bone loss in women with the BB genotype of VDR than in women with other genotype groups, but have found that Ca supplements (500 mg daily) abolish this difference. Another study (Ferrari et al. Reference Ferrari, Rizzoli, Chevally, Slosman, Eisman and Bonjour1995) has shown that change in lumbar spine bone density is associated with Ca intake in elderly subjects who have the VDR ‘Bb’ genotype group, but not in other genotypes. The largest study of VDR alleles in relation to bone mass, bone loss and Ca intake is that of MacDonald et al. (Reference MacDonald, McGuigan, Stewart, Black, Fraser, Ralston and Reid2006) who found little evidence of an interaction between Ca intake, VDR alleles and bone mass or bone loss. A functional polymorphism at position 677 (C677T) in the methylene tetrahydrofolate reductase gene has been associated with BMD and fracture in various studies (Miyao et al. Reference Miyao, Morita, Hosoi, Kurihara, Inoue, Hoshino, Shiraki, Yazaki and Ouchi2000; Jorgensen et al. Reference Jorgensen, Madsen, Madsen, Saleh, Abrahamsen, Fenger and Lauritzen2002; Abrahamsen et al. Reference Abrahamsen, Madsen, Tofteng, Stilgren, Bladbjerg, Kristensen, Brixen and Mosekilde2003; Bathum et al. Reference Bathum, von Bornemann, Christiansen, Madsen, Skytthe and Christensen2004), but the results have been inconsistent, since in some studies (Abrahamsen et al. Reference Abrahamsen, Madsen, Tofteng, Stilgren, Bladbjerg, Kristensen, Brixen and Mosekilde2003; Bathum et al. Reference Bathum, von Bornemann, Christiansen, Madsen, Skytthe and Christensen2004) the T allele has been associated with osteoporosis and in others (Jorgensen et al. Reference Jorgensen, Madsen, Madsen, Saleh, Abrahamsen, Fenger and Lauritzen2002) the C allele has been associated with osteoporosis. Recent work has suggested that folate status (McLean et al. Reference McLean, Karasik, Selhub, Tucker, Ordovas, Russo, Cupples, Jacques and Kiel2004) or riboflavin status (MacDonald et al. Reference MacDonald, McGuigan, Fraser, New, Ralston and Reid2004) might influence the association between methylene tetrahydrofolate reductase alleles and BMD.

Conclusions

Many advances have been made in understanding the role of genetic factors in osteoporosis over the past 10 years, but a great deal of additional research is required to identify the genes that regulate BMD and other phenotypes relevant to the pathogenesis of osteoporotic fractures. Until recently, most of the studies in the area of osteoporosis genetics have been underpowered, leading to results that have seldom been replicated (Ioannidis, Reference Ioannidis2003). It has now become clear that large-scale studies need to be assembled to evaluate the true role of genetic polymorphisms in osteoporosis and other complex diseases (Ioannidis et al. Reference Ioannidis, Gwinn, Little, Higgins, Bernstein and Boffetta2006). It is likely that large-scale studies, when combined with technological advances such as genome-wide association, will assist in identifying and validating the role of candidate gene polymorphisms in the regulation of BMD and other markers of osteoporosis susceptibility.

References

Abrahamsen, B, Madsen, JS, Tofteng, CL, Stilgren, L, Bladbjerg, EM, Kristensen, SR, Brixen, K & Mosekilde, L (2003) A common methylenetetrahydrofolate reductase (C677T) polymorphism is associated with low bone mineral density and increased fracture incidence after menopause: longitudinal data from the Danish osteoporosis prevention study. Journal of Bone and Mineral Research 18, 723729.CrossRefGoogle ScholarPubMed
Alam, I, Sun, Q, Liu, L, Koller, DL, Fishburn, T, Carr, LG, Econs, MJ, Foroud, T & Turner, CH (2005) Whole-genome scan for linkage to bone strength and structure in inbred Fischer 344 and Lewis rats. Journal of Bone and Mineral Research 20, 15891596.CrossRefGoogle ScholarPubMed
Albagha, OM, Pettersson, U, Stewart, A, McGuigan, FE, MacDonald, HM, Reid, DM & Ralston, SH (2005) Association of oestrogen receptor alpha gene polymorphisms with postmenopausal bone loss, bone mass, and quantitative ultrasound properties of bone. Journal of Medical Genetics 42, 240246.CrossRefGoogle ScholarPubMed
Andrew, T, Antioniades, L, Scurrah, KJ, MacGregor, AJ & Spector, TD (2005) Risk of wrist fracture in women is heritable and is influenced by genes that are largely independent of those influencing BMD. Journal of Bone and Mineral Research 20, 6774.CrossRefGoogle ScholarPubMed
Arai, H, Miyamoto, KI, Yoshida, M, Yamamoto, H, Taketani, Y, Morita, K et al. (2001) The polymorphism in the caudal-related homeodomain protein Cdx-2 binding element in the human vitamin D receptor gene. Journal of Bone and Mineral Research 16, 12561264.CrossRefGoogle ScholarPubMed
Arden, NK, Baker, J, Hogg, C, Baan, K & Spector, TD (1996) The heritability of bone mineral density, ultrasound of the calcaneus and hip axis length: a study of postmenopausal twins. Journal of Bone and Mineral Research 11, 530534.CrossRefGoogle ScholarPubMed
Balemans, W, Ebeling, M, Patel, N, Van Hul, E, Olson, P, Dioszegi, M et al. (2001) Increased bone density in sclerosteosis is due to the deficiency of a novel secreted protein (SOST). Human Molecular Genetics 10, 537543.CrossRefGoogle Scholar
Balemans, W, Foernzler, D, Parsons, C, Ebeling, M, Thompson, A, Reid, DM, Lindpaintner, K, Ralston, SH & Van Hul, W (2002 a) Lack of association between the SOST gene and bone mineral density in perimenopausal women: analysis of five polymorphisms. Bone 31, 515519.CrossRefGoogle ScholarPubMed
Balemans, W, Patel, N, Ebeling, M, Van Hul, E, Wuyts, W, Lacza, C et al. (2002 b) Identification of a 52 kb deletion downstream of the SOST gene in patients with van Buchem disease. Journal of Medical Genetics 39, 9197.CrossRefGoogle ScholarPubMed
Balemans, W, Van Wesenbeeck, L & Van Hul, W (2005) A clinical and molecular overview of the human osteopetroses. Calcified Tissue International 77, 263274.CrossRefGoogle ScholarPubMed
Bathum, L, von Bornemann, HJ, Christiansen, L, Madsen, JS, Skytthe, A & Christensen, K (2004) Evidence for an association of methylene tetrahydrofolate reductase polymorphism C677T and an increased risk of fractures: results from a population-based Danish twin study. Osteoporosis International 15, 659664.CrossRefGoogle Scholar
Beamer, WG, Shultz, KL, Donahue, LR, Churchill, GA, Sen, S, Wergedal, JR, Baylink, DJ & Rosen, CJ (2001) Quantitative trait loci for femoral and lumbar vertebral bone mineral density in C57BL/6J and C3H/HeJ inbred strains of mice. Journal of Bone and Mineral Research 16, 11951206.CrossRefGoogle ScholarPubMed
Bouxsein, ML, Rosen, CJ, Turner, CH, Ackert, CL, Shultz, KL, Donahue, LR et al. (2002) Generation of a new congenic mouse strain to test the relationships among serum insulin-like growth factor I, bone mineral density, and skeletal morphology in vivo. Journal of Bone and Mineral Research 17, 570579.CrossRefGoogle ScholarPubMed
Brunkow, M, Gardner, J, Van Ness, J, Paeper, B, Kovacevich, B, Proll, S et al. (2001) Bone dysplasia sclerosteosis results from loss of the sost gene product, a novel cystine knot-containing protein. American Journal of Human Genetics 68, 577589.CrossRefGoogle ScholarPubMed
Cummings, SR, Nevitt, MC, Browner, WS, Stone, K, Fox, KM, Ensrud, KE, Cauley, J, Black, D & Vogt, TM (1995) Risk factors for hip fracture in white women. Study of Osteoporotic Fractures Research Group. New England Journal of Medicine 332, 767773.CrossRefGoogle ScholarPubMed
Deng, HW, Chen, WM, Recker, S, Stegman, MR, Li, JL, Davies, KM, Zhou, Y, Deng, H, Heaney, R & Recker, RR (2000) Genetic determination of Colles' fracture and differential bone mass in women with and without Colles' fracture. Journal of Bone and Mineral Research 15, 12431252.CrossRefGoogle ScholarPubMed
Deng, HW, Xu, FH, Huang, QY, Shen, H, Deng, H, Conway, T et al. (2002) A whole-genome linkage scan suggests several genomic regions potentially containing quantitative trait loci for osteoporosis. Journal of Clinical Endocrinology and Metabolism 87, 51515159.CrossRefGoogle ScholarPubMed
Devoto, M, Shimoya, K, Caminis, J, Ott, J, Tenenhouse, A, Whyte, MP et al. (1998) First-stage autosomal genome screen in extended pedigrees suggests genes predisposing to low bone mineral density on chromosomes 1p, 2p and 4q. European Journal of Human Genetics 6, 151157.CrossRefGoogle ScholarPubMed
Efstathiadou, Z, Tsatsoulis, A & Ioannidis, JP (2001) Association of collagen Ialpha 1 Sp1 polymorphism with the risk of prevalent fractures: a meta-analysis. Journal of Bone and Mineral Research 16, 15861592.CrossRefGoogle ScholarPubMed
Fang, Y, van Meurs, JB, Bergink, AP, Hofman, A, van Duijn, CM, van Leeuwen, JP, Pols, HA & Uitterlinden, AG (2003) Cdx-2 polymorphism in the promoter region of the human vitamin D receptor gene determines susceptibility to fracture in the elderly. Journal of Bone and Mineral Research 18, 16321641.CrossRefGoogle ScholarPubMed
Fang, Y, van Meurs, JB, D'Alesio, A, Jhamai, M, Zhao, H, Rivadeneira, F, Hofman, A, van Leeuwen, JP, Jehan, F, Pols, HA & Uitterlinden, AG (2005) Promoter and 3′-untranslated-region haplotypes in the vitamin D receptor gene predispose to osteoporotic fracture: the Rotterdam Study. American Journal of Human Genetics 77, 807823.CrossRefGoogle ScholarPubMed
Ferrari, S, Rizzoli, R, Chevally, T, Slosman, D, Eisman, JA & Bonjour, J-P (1995) Vitamin D receptor gene polymorphisms and change in lumbar spine bone mineral density. Lancet 345, 423424.CrossRefGoogle ScholarPubMed
Ferrari, SL, Deutsch, S, Choudhury, U, Chevalley, T, Bonjour, JP, Dermitzakis, ET, Rizzoli, R & Antonarakis, SE (2004) Polymorphisms in the low-density lipoprotein receptor-related protein 5 (LRP5) gene are associated with variation in vertebral bone mass, vertebral bone size, and stature in whites. American Journal of Human Genetics 74, 866875.CrossRefGoogle ScholarPubMed
Fisher, SA, Lanchbury, JS & Lewis, CM (2003) Meta-analysis of four rheumatoid arthritis genome-wide linkage studies: confirmation of a susceptibility locus on chromosome 16. Arthritis and Rheumatism 48, 12001206.CrossRefGoogle ScholarPubMed
Frattini, PJ, Orchard, C, Sobacchi, S, Giliani, M, Abinun, JP, Mattsson, DJ et al. (2000) Defects in TCIRG1 subunit of the vacuolar proton pump are responsible for a subset of human autosomal recessive osteopetrosis. Nature Genetics 25, 343346.CrossRefGoogle ScholarPubMed
Garcia-Giralt, N, Enjuanes, A, Bustamante, M, Mellibovsky, L, Nogues, X, Carreras, R, Diez-Perez, A, Grinberg, D & Balcells, S (2005) In vitro functional assay of alleles and haplotypes of two COL1A1-promoter SNPs. Bone 36, 902908.CrossRefGoogle ScholarPubMed
Garcia-Giralt, N, Nogues, X, Enjuanes, A, Puig, J, Mellibovsky, L, Bay-Jensen, A, Carreras, R, Balcells, S, Diez-Perez, A & Grinberg, D (2002) Two new single nucleotide polymorphisms in the COLIA1 upstream regulatory region and their relationship with bone mineral density. Journal of Bone and Mineral Research 17, 384393.CrossRefGoogle Scholar
Garnero, P, Arden, NK, Griffiths, G, Delmas, PD & Spector, TD (1996) Genetic influence on bone turnover in postmenopausal twins. Journal of Clinical Endocrinology and Metabolism 81, 140146.Google Scholar
Gong, Y, Slee, RB, Fukai, N, Rawadi, G, Roman-Roman, S, Reginato, AM et al. (2001) LDL receptor-related protein 5 (LRP5) affects bone accrual and eye development. Cell 107, 513523.CrossRefGoogle ScholarPubMed
Grant, SFA, Reid, DM, Blake, G, Herd, R, Fogelman, I & Ralston, SH (1996) Reduced bone density and osteoporosis associated with a polymorphic Sp1 site in the collagen type I alpha 1 gene. Nature Genetics 14, 203205.CrossRefGoogle ScholarPubMed
Gross, C, Eccleshall, TR, Malloy, PJ, Villa, ML, Marcus, R & Feldman, D (1997) The presence of a polymorphism at the translation initiation site of the vitamin D receptor gene is associated with low bone mineral density in postmenopausal Mexican-American women. Journal of Bone and Mineral Research 12, 18501856.Google Scholar
Gross, C, Krishnan, AV, Malloy, PJ, Eccleshall, TR, Zhao, XY & Feldman, D (1998) The vitamin D receptor gene start codon polymorphism: a functional analysis of FokI variants. Journal of Bone and Mineral Research 13, 16911699.CrossRefGoogle ScholarPubMed
Gueguen, R, Jouanny, P, Guillemin, F, Kuntz, C, Pourel, J & Siest, G (1995) Segregation analysis and variance components analysis of bone mineral density in healthy families. Journal of Bone and Mineral Research 12, 20172022.CrossRefGoogle Scholar
Harris, SS, Patel, MS, Cole, DE & Dawson-Hughes, B (2000) Associations of the collagen type I alpha1 Sp1 polymorphism with five-year rates of bone loss in older adults. Calcified Tissue International 66, 268271.CrossRefGoogle Scholar
Herrington, DM, Howard, TD, Brosnihan, KB, McDonnell, DP, Li, X, Hawkins, GA, Reboussin, DM, Xu, J, Zheng, SL, Meyers, DA & Bleecker, ER (2002) Common estrogen receptor polymorphism augments effects of hormone replacement therapy on E-selectin but not C-reactive protein. Circulation 105, 18791882.CrossRefGoogle Scholar
Ichikawa, S, Koller, DL, Johnson, ML, Lai, D, Xuei, X, Edenberg, HJ et al. (2006) Human ALOX12, but not ALOX15, is associated with BMD in white men and women. Journal of Bone and Mineral Research 21, 556564.CrossRefGoogle Scholar
Ioannidis, JP (2003) Genetic associations: false or true? Trends in Molecular Medicine 9, 135138.CrossRefGoogle ScholarPubMed
Ioannidis, JP, Gwinn, M, Little, J, Higgins, JP, Bernstein, JL, Boffetta, P et al. (2006) A road map for efficient and reliable human genome epidemiology. Nature Genetics 38, 35.CrossRefGoogle ScholarPubMed
Ioannidis, JP, Ralston, SH, Bennett, ST, Brandi, ML, Grinberg, D, Karassa, FB et al. (2004) Differential genetic effects of ESR1 gene polymorphisms on osteoporosis outcomes. Journal of the American Medical Association 292, 21052114.CrossRefGoogle ScholarPubMed
Ioannidis, JP, Stavrou, I, Trikalinos, TA, Zois, C, Brandi, ML, Gennari, L, Albagha, O, Ralston, SH & Tsatsoulis, A (2002) Association of polymorphisms of the estrogen receptor alpha gene with bone mineral density and fracture risk in women: a meta-analysis. Journal of Bone and Mineral Research 17, 20482060.CrossRefGoogle ScholarPubMed
Jorgensen, HL, Madsen, JS, Madsen, B, Saleh, MM, Abrahamsen, B, Fenger, M & Lauritzen, JB (2002) Association of a common allelic polymorphism (C677T) in the methylene tetrahydrofolate reductase gene with a reduced risk of osteoporotic fractures. A case control study in Danish postmenopausal women. Calcified Tissue International 71, 386392.Google ScholarPubMed
Kammerer, CM, Schneider, JL, Cole, SA, Hixson, JE, Samollow, PB, O'Connell, JR et al. (2003) Quantitative trait loci on chromosomes 2p, 4p, and 13q influence bone mineral density of the forearm and hip in Mexican Americans. Journal of Bone and Mineral Research 18, 22452252.CrossRefGoogle ScholarPubMed
Kannus, P, Palvanen, M, Kaprio, J, Parkkari, J & Koskenvuo, M (1999) Genetic factors and osteoporotic fractures in elderly people: prospective 25 year follow up of a nationwide cohort of elderly Finnish twins. British Medical Journal 319, 13341337.CrossRefGoogle ScholarPubMed
Karasik, D, Cupples, LA, Hannan, MT & Kiel, DP (2003) Age, gender, and body mass effects on quantitative trait loci for bone mineral density: the Framingham Study. Bone 33, 308316.CrossRefGoogle ScholarPubMed
Karasik, D, Myers, RH, Cupples, LA, Hannan, MT, Gagnon, DR, Herbert, A & Kiel, DP (2002) Genome screen for quantitative trait loci contributing to normal variation in bone mineral density: the Framingham Study. Journal of Bone and Mineral Research 17, 17181727.CrossRefGoogle ScholarPubMed
Klein, RF, Allard, J, Avnur, Z, Nikolcheva, T, Rotstein, D, Carlos, AS, Shea, M, Waters, RV, Belknap, JK, Peltz, G & Orwoll, ES (2004) Regulation of bone mass in mice by the lipoxygenase gene Alox15. Science 303, 229232.CrossRefGoogle ScholarPubMed
Klein, RF, Mitchell, SR, Phillips, TJ, Belknap, JK & Orwoll, ES (1998) Genetic analysis of bone mass in mice. Journal of Bone and Mineral Research 13, 16481656.CrossRefGoogle ScholarPubMed
Koay, MA, Woon, PY, Zhang, Y, Miles, LJ, Duncan, EL, Ralston, SH et al. (2004) Influence of LRP5 polymorphisms on normal variation in BMD. Journal of Bone and Mineral Research 19, 16191627.CrossRefGoogle ScholarPubMed
Koh, JM, Jung, MH, Hong, JS, Park, HJ, Chang, JS, Shin, HD, Kim, SY & Kim, GS (2004) Association between bone mineral density and LDL receptor-related protein 5 gene polymorphisms in young Korean men. Journal of Korean Medical Science 19, 407412.CrossRefGoogle ScholarPubMed
Koller, DL, Alam, I, Sun, Q, Liu, L, Fishburn, T, Carr, LG, Econs, MJ, Foroud, T & Turner, CH (2005) Genome screen for bone mineral density phenotypes in Fisher 344 and Lewis rat strains. Mammalian Genome 16, 578586.CrossRefGoogle ScholarPubMed
Koller, DL, Econs, MJ, Morin, PA, Christian, JC, Hui, SL, Parry, P et al. (2000) Genome Screen for QTLs Contributing to Normal Variation in Bone Mineral Density and Osteoporosis. Journal of Clinical Endocrinology and Metabolism 85, 31163120.Google ScholarPubMed
Koller, DL, Liu, G, Econs, MJ, Hui, SL, Morin, PA, Joslyn, G et al. (2001) Genome screen for quantitative trait loci underlying normal variation in femoral structure. Journal of Bone and Mineral Research 16, 985991.CrossRefGoogle ScholarPubMed
Krall, EA & Dawson-Hughes, B (1993) Heritable and life-style determinants of bone mineral density. Journal of Bone and Mineral Research 8, 19.CrossRefGoogle ScholarPubMed
Krall, EA, Parry, P, Lichter, JB & Dawson-Hughes, B (1995) Vitamin D receptor alleles and rates of bone loss: influence of years since menopause and calcium intake. Journal of Bone and Mineral Research 10, 978984.CrossRefGoogle ScholarPubMed
Langdahl, BL, Knudsen, JY, Jensen, HK, Gregersen, N & Eriksen, EF (1997) A sequence variation: 713–8 elC in the transforming growth factor-beta 1 gene has higher prevalence in osteoporotic women than in normal women and is associated with very low bone mass in osteoporotic women and increased bone turnover in both osteoporotic and normal women. Bone 20, 289294.CrossRefGoogle Scholar
Little, RD, Carulli, JP, Del Mastro, RG, Dupuis, J, Osborne, M, Folz, C et al. (2002) A mutation in the LDL receptor-related protein 5 gene results in the autosomal dominant high-bone-mass trait. American Journal of Human Genetics 70, 1119.CrossRefGoogle ScholarPubMed
Liu, PY, Lu, Y, Long, JR, Xu, FH, Shen, H, Recker, RR & Deng, HW (2004) Common variants at the PCOL2 and Sp1 binding sites of the COL1A1 gene and their interactive effect influence bone mineral density in Caucasians. Journal of Medical Genetics 41, 752757.CrossRefGoogle ScholarPubMed
MacDonald, HM, McGuigan, FE, Fraser, WD, New, SA, Ralston, SH & Reid, DM (2004) Methylenetetrahydrofolate reductase polymorphism interacts with riboflavin intake to influence bone mineral density. Bone 35, 957964.CrossRefGoogle ScholarPubMed
MacDonald, HM, McGuigan, FE, Stewart, A, Black, AJ, Fraser, WD, Ralston, S & Reid, DM (2006) Large-scale population-based study shows no evidence of association between common polymorphism of the VDR gene and BMD in British women. Journal of Bone and Mineral Research 21, 151162.CrossRefGoogle ScholarPubMed
MacDonald, HM, McGuigan, FEA, New, SA, Campbell, MK, Golden, MH, Ralston, SH & Reid, DM (2001) COL1A1 Sp1 polymorphism predicts perimenopausal and early postmenopausal spinal bone loss. Journal of Bone and Mineral Research 16, 16341641.CrossRefGoogle ScholarPubMed
McLean, RR, Karasik, D, Selhub, J, Tucker, KL, Ordovas, JM, Russo, GT, Cupples, LA, Jacques, PF & Kiel, DP (2004) Association of a common polymorphism in the methylenetetrahydrofolate reductase (MTHFR) gene with bone phenotypes depends on plasma folate status. Journal of Bone and Mineral Research 19, 410418.CrossRefGoogle ScholarPubMed
Mahaney, MC, Morin, P, Rodriguez, LA, Newman, DE & Rogers, J (1997) A quantitative trait locus on chromosome 11 may influence bone mineral density at several sites: quantitative analysis in pedigreed baboons. Journal of Bone and Mineral Research 12, Suppl. 1, s118Abstr.Google Scholar
Mann, V, Hobson, EE, Li, B, Stewart, TL, Grant, SF, Robins, SP, Aspden, RM & Ralston, SH (2001) A COL1A1 Sp1 binding site polymorphism predisposes to osteoporotic fracture by affecting bone density and quality. Journal of Clinical Investigation 107, 899907.CrossRefGoogle ScholarPubMed
Mann, V & Ralston, SH (2003) Meta-analysis of COL1A1 Sp1 polymorphism in relation to bone mineral density and osteoporotic fracture. Bone 32, 711717.CrossRefGoogle ScholarPubMed
Michaelsson, K, Melhus, H, Ferm, H, Ahlbom, A & Pedersen, NL (2005) Genetic liability to fractures in the elderly. Archives of Internal Medicine 165, 18251830.CrossRefGoogle ScholarPubMed
Miyao, M, Morita, H, Hosoi, T, Kurihara, H, Inoue, S, Hoshino, S, Shiraki, M, Yazaki, Y & Ouchi, Y (2000) Association of methylenetetrahydrofolate reductase (MTHFR) polymorphism with bone mineral density in postmenopausal Japanese women. Calcified Tissue International 66, 190194.CrossRefGoogle ScholarPubMed
Mizuguchi, T, Furuta, I, Watanabe, Y, Tsukamoto, K, Tomita, H, Tsujihata, M et al. (2004) LRP5, low-density-lipoprotein-receptor-related protein 5, is a determinant for bone mineral density. Journal of Human Genetics 49, 8086.CrossRefGoogle ScholarPubMed
Orwoll, ES, Belknap, JK & Klein, RF (2001) Gender specificity in the genetic determinants of peak bone mass. Journal of Bone and Mineral Research 16, 19621971.CrossRefGoogle ScholarPubMed
Peacock, M, Koller, DL, Fishburn, T, Krishnan, S, Lai, D, Hui, S, Johnston, CC, Foroud, T & Econs, MJ (2005) Sex-specific and non-sex-specific quantitative trait loci contribute to normal variation in bone mineral density in men. Journal of Clinical Endocrinology and Metabolism 90, 30603066.CrossRefGoogle ScholarPubMed
Pettersson, U, Albagha, OM, Mirolo, M, Taranta, A, Frattini, A, McGuigan, FE et al. (2005) Polymorphisms of the CLCN7 gene are associated with BMD in women. Journal of Bone and Mineral Research 20, 19601967.CrossRefGoogle ScholarPubMed
Qureshi, AM, McGuigan, FEA, Seymour, DG, Hutchison, JD, Reid, DM & Ralston, SH (2001) Association between COLIA1 Sp1 alleles and femoral neck geometry. Calcified Tissue International 69, 6772.CrossRefGoogle ScholarPubMed
Ralston, SH, Galwey, N, Mackay, I, Albagha, OM, Cardon, L, Compston, JE et al. (2005) Loci for regulation of bone mineral density in men and women identified by genome wide linkage scan: the FAMOS study. Human Molecular Genetics 14, 943951.CrossRefGoogle ScholarPubMed
Sawcer, SJ, Maranian, M, Singlehurst, S, Yeo, T, Compston, A, Daly, MJ et al. (2004) Enhancing linkage analysis of complex disorders: an evaluation of high-density genotyping. Human Molecular Genetics 13, 19431949.Google ScholarPubMed
Shah, R, Rahaman, B, Hurley, CK & Posch, PE (2005) Allelic diversity in the TGFB1 regulatory region: characterization of novel functional single nucleotide polymorphisms. Human Genetics 119, 114.Google ScholarPubMed
Sobacchi, C, Vezzoni, P, Reid, DM, McGuigan, FE, Frattini, A, Mirolo, M, Albhaga, OM, Musio, A, Villa, A & Ralston, SH (2004) Association between a polymorphism affecting an AP1 binding site in the promoter of the TCIRG1 gene and bone mass in women. Calcified Tissue International 74, 3541.CrossRefGoogle ScholarPubMed
Stewart, TL, Jin, H, McGuigan, FE, Albagha, OM, Garcia-Giralt, N, Bassiti, A, Grinberg, D, Balcells, S, Reid, DM & Ralston, SH (2006) Haplotypes defined by promoter and intron 1 polymorphisms of the COLIA1 gene regulate bone mineral density in women. Journal of Clinical Endocrinology and Metabolism 91, 41124117.CrossRefGoogle ScholarPubMed
Stewart, TL, Roschger, P, Misof, BM, Mann, V, Fratzl, P, Klaushofer, K, Aspden, RM & Ralston, SH (2005) Association of COLIA1 Sp1 alleles with defective bone nodule formation in vitro and abnormal bone mineralisation in vivo. Calcified Tissue International 77, 113118.CrossRefGoogle ScholarPubMed
Styrkarsdottir, U, Cazier, J-B, Kong, A, Rolfsson, O, Larsen, H, Bjarnadottir, E et al. (2003) Linkage of osteoporosis to chromosome 20p12 and association to BMP2. PLoS Biology 1, E69.CrossRefGoogle ScholarPubMed
Thakkinstian, A, D'Este, C, Eisman, J, Nguyen, T & Attia, J (2004) Meta-analysis of molecular association studies: vitamin D receptor gene polymorphisms and BMD as a case study. Journal of Bone and Mineral Research 19, 419428.CrossRefGoogle ScholarPubMed
Torgerson, DJ, Campbell, MK, Thomas, RE & Reid, DM (1996) Prediction of perimenopausal fractures by bone mineral density and other risk factors. Journal of Bone and Mineral Research 11, 293297.CrossRefGoogle ScholarPubMed
Turner, CH, Sun, Q, Schriefer, J, Pitner, N, Price, R, Bouxsein, ML, Rosen, CJ, Donahue, LR, Shultz, KL & Beamer, WG (2003) Congenic mice reveal sex-specific genetic regulation of femoral structure and strength. Calcified Tissue International 73, 297303.CrossRefGoogle ScholarPubMed
Uitterlinden, AG, Arp, PP, Paeper, BW, Charmley, P, Proll, S, Rivadeneira, F et al. (2004) Polymorphisms in the sclerosteosis/van Buchem disease gene (SOST) region are associated with bone-mineral density in elderly whites. American Journal of Human Genetics 75, 10321045.CrossRefGoogle ScholarPubMed
Uitterlinden, AG, Burger, H, Huang, Q, Yue, F, McGuigan, FEA, Grant, SFA, Hofman, A, van Leeuwen, JPTM, Pols, HAP & Ralston, SH (1998) Relation of alleles of the collagen type I a 1 gene to bone density and risk of osteoporotic fractures in postmenopausal women. New England Journal of Medicine 338, 10161022.CrossRefGoogle Scholar
Urano, T, Shiraki, M, Ezura, Y, Fujita, M, Sekine, E, Hoshino, S, Hosoi, T, Orimo, H, Emi, M, Ouchi, Y & Inoue, S (2004) Association of a single-nucleotide polymorphism in low-density lipoprotein receptor-related protein 5 gene with bone mineral density. Journal of Bone and Mineral Metabolism 22, 341345.CrossRefGoogle ScholarPubMed
van Meurs, JB, Rivadeneira, F, Jhamai, M, Hugens, W, Hofman, A, van Leeuwen, JP, Pols, HA & Uitterlinden, AG (2006) Common genetic variation of the low-density lipoprotein receptor-related protein 5 and 6 genes determines fracture risk in elderly white men. Journal of Bone and Mineral Research 21, 141150.CrossRefGoogle ScholarPubMed
Wilson, SG, Reed, PW, Andrew, T, Barber, MJ, Lindersson, M, Langdown, M et al. (2004) A genome-screen of a large twin cohort reveals linkage for quantitative ultrasound of the calcaneus to 2q33–37 and 4q12–21. Journal of Bone and Mineral Research 19, 270277.CrossRefGoogle ScholarPubMed
Wilson, SG, Reed, PW, Bansal, A, Chiano, M, Lindersson, M, Langdown, M et al. (2003) Comparison of genome screens for two independent cohorts provides replication of suggestive linkage of bone mineral density to 3p21 and 1p36. American Journal of Human Genetics 72, 144155.CrossRefGoogle ScholarPubMed
Yamada, Y, Ando, F, Niino, N & Shimokata, H (2005) Association of a -1997G→T polymorphism of the collagen Ialpha1 gene with bone mineral density in postmenopausal Japanese women. Human Biology 77, 2736.CrossRefGoogle Scholar
Yamada, Y, Miyauchi, A, Takagi, Y, Tanaka, M, Mizuno, M & Harada, A (2001) Association of the C-509T polymorphism, alone of in combination with the T869→C polymorphism, of the transforming growth factor-beta1 gene with bone mineral density and genetic susceptibility to osteoporosis in Japanese women. Journal of Molecular Medicine 79, 149156.CrossRefGoogle Scholar
Zhang, YY, Lei, SF, Mo, XY, Wang, YB, Li, MX & Deng, HW (2005) The -1997 G/T polymorphism in the COLIA1 upstream regulatory region is associated with hip bone mineral density (BMD) in Chinese nuclear families. Calcified Tissue International 76, 107112.CrossRefGoogle ScholarPubMed
Figure 0

Table 1. Quantitative trait loci for bone mineral density detected by genome-wide linkage scan*