Review Open Access
Copyright ©2012 Baishideng. All rights reserved.
World J Obstet Gynecol. Oct 10, 2012; 1(3): 20-28
Published online Oct 10, 2012. doi: 10.5317/wjog.v1.i3.20
Intrauterine growth restriction and genetic determinants - existing findings, problems, and further direction
Xiu-Quan Zhang, Department of Obstetrics and Gynecology, Department of Surgery, Division of Cardiothoracic Surgery, University of Utah School of Medicine, Salt Lake City, UT 84124, United States
Author contributions: Zhang XQ solely contributed to this paper.
Correspondence to: Xiu-Quan Zhang, MD, PhD, Department of Obstetrics and Gynecology, Department of Surgery, Division of Cardiothoracic Surgery, University of Utah School of Medicine, Salt Lake City, UT 84124, United States. xiuquan.zhang@hsc.utah.edu
Telephone: +1-801-5815311 Fax: +1-801-5853936
Received: April 18, 2012
Revised: July 6, 2012
Accepted: September 12, 2012
Published online: October 10, 2012

Abstract

Fetal growth is determined largely by the nutrient supply, placental transport function, and growth hormones. Recently, gene mutation and expression, especially of those genes associated with the proteins that are related to the fetal growth, have been reported to play an important role in the development of intrauterine growth restriction (IUGR). Fetal growth epigenetics, a new concept in fetal growth, has resulted from studies on fetal programing. This paper outlines the findings of our serial studies on IUGR, and summarizes data on IUGR animal models, placental function in transferring nutrients, cell proliferation dynamics in IUGR, and experimental treatment of IUGR. We review genetic approaches to IUGR, especially those relating to growth factor genes, angiotensinogen genes and other gene mutations. We also discuss the epigenetics of fetal growth and future study directions on fetal growth restriction. These should be valuable in elucidating the mechanisms employed by the fetus and in helping to develop interventional strategies that might prevent the development of IUGR.

Key Words: Intrauterine growth restriction, Gene, Epigenetics, Placenta



INTRODUCTION

Intrauterine growth restriction (IUGR) is an important obstetric complication affecting 5% of pregnancies[1]. This condition represents an in utero shift from the expected pattern of fetal growth potential into reduced birth weight. This leads to increased risk for intrauterine compromise, stillbirth, preterm birth and adverse perinatal and long-term outcomes[2-5]. It is generally accepted that IUGR is associated with a poor nutrient and oxygen supply, although the specific mechanisms involved in IUGR development are mostly unknown.

Fetal growth is determined largely by the nutrient supply, which is dependent on placental transport functions. Though placental transportation depends on the concentration gradient between maternal and fetal blood, it seems that placental blood flow and the activity of specific membrane transporters play more important role[6,7]. Regardless of what the exact mechanism for IUGR might be, growth factors and their interaction with their receptors are most likely to be involved[8]. In addition to nutrition, placental function, and growth hormones, gene mutation and expression have been reported to be associated with IUGR[9-13]. Those genes associated with the proteins that are related to the fetal growth such as growth factor genes are particularly implicated. In recent years, a new concept in fetal growth, fetal growth epigenetics, has resulted from studies on fetal programing[3,14-16].

The information in this paper outlines findings from our serial studies on IUGR and reviews the genetic approaches in recent years. This should be valuable in the elucidation of the mechanisms employed by the fetus and help in the development of interventional strategies that might prevent the development of IUGR.

EXISTING FINDINGS
IUGR animal models

Several methods have been used to establish the animal model of IUGR. These include uterine artery ligation, passive smoking, alcohol exposure, administration of L-arginine or actinomycin, and maternal malnutrition. To date, mouse, rat, guinea pig, rabbit, sheep, monkey, dog, and pig have been used to establish the IUGR model[17-19].

We have reported the establishment of an IUGR rabbit model using passive smoking and have studied the fetal and maternal plasma amino acid concentration, the blood flow of utero-placenta and changes in cell proliferation cycle in brain, liver and placenta[20,21]. In addition, the effect of treatment with histiocyte activators and vasodilators on the IUGR was also observed[7,18]. Recent studies on gene mutations have shown that specific genes might play an important role in the development of IUGR[9-16].

Active transfer of nutrients through the placenta

Although fetal plasma nutrient concentration is related to the maternal concentration, some amino acids, especially most essential amino acids, show higher concentrations in fetal plasma than in maternal plasma, proving that the absorption of amino acids from the maternal body to the fetus is an active process[22]. Maternal rabbit total plasma amino acid concentration at 28-d of pregnancy was less than that of non-pregnant rabbits. These changes were more apparent with essential amino acids. In the rabbit IUGR model, smoking could reduce the ability of the placenta to transfer nutrients from the maternal to the fetal side. It was found that the total amino acid concentration in fetus was lower in rabbits with IUGR induced by smoking than in controls. In fetal plasma, the levels of all essential amino acids, except threonine, and of nonessential amino acids including serine, glutamine, alanine, tyrosine, arginine and histidine showed the same patterns in smoking groups as in the control group[7,20].

Dynamics of cell proliferation in IUGR

Cell proliferation in the fetus is very active, especially in early embryo phase. Proliferation depends on the cell division, as measured by analyzing the cell cycle using flow cytometry. The more rapid the cell division, the shorter the G0 and G1 phases. In another words, lower proliferation will leave more cells in G0 or G1 phase, while active proliferation will result in more in S + G2 + M phases. By analyzing the cell cycle using flow cytometry, we found that the ratio of the cells in G0 + G1 phases to the total cells was increased and the ratio of the cells in S + G2 + M phases to the total cells was decreased in fetal brain, liver and placenta in IUGR rabbits induced by passive smoking. This change was particularly prominent in brain. These results suggest that the proliferation of the fetal cells and placenta were inhibited significantly, resulting in growth retardation of the fetus. This data demonstrated that the transformation from G1 to S phases was restricted in IUGR and that cell proliferation is prevented[21,23].

Experimental treatment of IUGR

IUGR rabbits caused by passive smoking were treated with histiocyte activators (include ATP, Co-A, and Cyt-C) and vasodilators (InjectioSalviaeMoltiorrhizaeComposita). The results showed that the vasodilators could increase the uteroplacental blood flow by about 35% and induced some progress in fetal development, as indicated by a 9.8% increase in fetal body weight. The fetal plasma amino acid concentration was also higher than that of the control group. It seems that the histiocyte activators were more effective in promoting fetal development (24.6% increase of fetal body weight) than the vasodilators. Although the effect of the histiocyte activators was not as marked as that of vasodilators in increasing uteroplacental blood flow, the amino acid concentration of fetal plasma increased more after histiocyte activator treatment than vasodilator treatment. This suggests that vasodilators could be used to treat the IUGR caused by vasospasm. Histiocyte activators appear to be better than vasodilators in increasing the nutrient transportation of the placenta[24].

IUGR AND GENE APPROACHES

Although the human genome contains about 30 000 genes only a small number of them are turned on within a particular tissue type. Cell proliferation and differentiation results from the ability of tissues to express different genes from the same basic set of genetic information stored in DNA. This small subset of genes allows the cells to produce proteins unique to their functions. Although gene expression is controlled by epigenetic modifications, the gene sequence obviously plays the most important role in the proliferation and differentiation of cells.

IUGR and growth factor genes

Several studies have demonstrated altered concentrations of human placental growth hormones (hPGH), insulin-like growth factor-I (IGF-I) and IGF banding protein (IGFBPs) in the maternal circulation and the fetal compartment in pregnancies with IUGR[25,26]. Koutsaki, evaluating the expression status of the hPGH, the IGF-I, IGFBP-1 and IGFBP-3 genes in placentas from human IUGR pregnancies of no apparent etiology, found that hPGH, IGF-I, IGFBP-1 and IGFBP-3 expression was significantly lower than that in placentas showing normal fetal growth. However, these alterations are not known causative factors of IUGR or associated with other pathogenetic mechanisms[27].

Type I insulin-like growth factor receptor (IGF-IR) is widely expressed across many cell types in fetal and postnatal tissues. The activation of this receptor, after the binding of secreted IGF-I and IGF-IR promotes cell differentiation and proliferation. An association has been found between IGF-IR gene mutations and low birth weight[28]. IGF-IR gene anomalies such as heterozygous IGF-IR mutation or insufficiency of the IGF-IR gene were found in patients presenting with low birth weight and birth height[29,30]. This phenotype is associated with family history of low birth weight and a normal or increased IGF-I level and/or a normal or increased GH response to GH stimulation test[31,32]. However, such patients show less response to GH treatment than common small for gestational age short-stature patients[33].

Choi et al[34] reported a family with both a novel heterozygous mutation of the IGF-1R gene and a segmental deletion encompassing the entire IGF-1R, resulting in IGF-I resistance and leading to IUGR and postnatal growth failure. In vitro studies of fibroblasts from subjects carrying the gene clearly demonstrated reduced IGF-1R expression and subsequent IGF-I resistance, as assessed by IGF-1R phosphorylation and post-receptor signal transduction. This indicates that IGF-1R mutations should be considered in the differential diagnosis of familial IUGR patients with persistent short stature[34]. Umbers found that material inflammation of the placenta could disturb the IGF expression and cause IUGR[35].

Experiments in animals demonstrated the importance of IGF in the regulation of both intrauterine and postnatal growth. Baker reported that isolated inactivation of IGF-I resulted in restrictions in fetal development (40% delay compared with wild type mice) and Liu found that postnatal growth was further impaired, reaching only 30% of that in normal mice[36,37]. Knockout of both IGF-I and IGF-II or knockout of both IGF-II and IGF-I receptors resulted in severe growth retardation (Figure 1)[38]. These experiments clearly demonstrated that IGF-I is a major regulator of both pre- and postnatal growth. Several authors reporting IGF-IR gene mutations observed effects on birth weight, height, serum IGF-I and additional complications (Table 1).

Table 1 IGF-I receptor mutations.
Gene mutationBirth weight (SD)Birth height (SD)Complications
Arg108Gln/Lys115Asn-3.5-4.8Microcephaly, abnormal speech[29]
Arg59Ter-3.5-3.0Microcephaly, Delay in speech[29]
Arg709Gln-1.5-2.6Mental retardation[28]
Gly1050Lys-2.1-4.0Insulin resistance[32]
Arg281Gln-3.1-5.0Decreased cell proliferation[31]
Val599Glu-2.3-2.1Developmental delay[33]
Gly1125Ala-1.8-3.6Microcephaly, clinodactyly, delayed menarche, diabetes mellitus[30]
Figure 1
Figure 1 Effects of disruption of the insulin-like growth factor system on fetal growth in mice, expressed as a percentage of normal body weight[38]. IGF: Insulin-like growth factor.
IUGR and angiotensinogen gene

IUGR has been associated with insufficient placental circulation, which may result from failed maternal physiologic changes such as abnormal spiral artery remodeling and reduced maternal blood volume. Morgan reported that spiral artery remodeling might be related to the angiotensinogen gene[39]. We compared maternal blood DNA in 174 patients with IUGR, 62 patients with placental abruption, and 60 patients with both preeclampsia compared with a control group comprising 400 consecutive cases of women with term pregnancies and infants with birth weight between the 5th and 95th percentiles. We also examined 162 DNA samples from fetal blood with IUGR and 240 normal fetuses from control group for the Thr235 polymorphism[40] (Tables 2 and 3).

Table 2 Maternal AGT Thr235 genotypes[40].
GroupsNo.Genotype
P value
MM (%)MT (%)TT (%)
Control400170 (42.5)158 (39.5)72 (18.0)
IUGR17433 (19.0)66 (37.9)75 (43.1)< 0.001
Preeclampsia + IUGR6011 (18.3)24 (40.0)25 (41.7)< 0.001
Placental abruption629 (14.5)27 (43.5)26 (41.9)< 0.001
Figure 2
Figure 2 Real-time polymerase chain reaction for angiotensinogen M235T genotyping with a single labeled fluorescein probe[42]. A: 5’-fluorescein-labeled 23-mer (blocked at the 3’-end with phosphate) homologous to the wild-type sequence, covered the polymorphic site with the fluorescein opposite two genes; B: The probe was included in the PCR amplification mixture with primers. Melting curve data are presented as fluorescein vs temperature; C: Curves for homozygous wild type (peak at 69.4oC), homozygous mutant type (peak at 63.54oC), heterozygous (two peaks at 63.54oC and 69.44oC respectively) and no-template control are shown; D: Placental cross section HE staining from AGT MM, MT, and TT placentas. Quantitative analysis of the placenta indicated that the villus area (red) and capillary area (bright red) in smoking-induced intrauterine growth restriction were significantly decreased.
Table 3 AGT Thr235 alleles frequencies analysis[40].
GroupsAGT genotype
T Allele Freq.P value
Met235 allelesThr235 alleles
Maternal DNA
Control (400)4983020.378
IUGR (174)1402080.598< 0.001
Preeclampsia + IUGR (60)44760.633< 0.001
Placental abruption (62)45790.637< 0.001
Fetal DNA
Control (240)2981820.379
IUGR (160)1311890.591< 0.001

According to the AGT genotyping results using real time PCR, angiotensinogen genotypes were divided into three groups: MM (homozygous for angiotensinogen Met235 allele), TT (homozygous for angiotensinogen Thr235 allele), and MT (heterozygous). It has been demonstrated that maternal and fetal angiotensinogen Thr235 genotypes are associated with an increased risk of IUGR[40]. The angiotensinogen Thr235 allele may predispose women to deliver growth-restricted fetuses. Further studies have shown that mutation of angiotensinogen gene Thr235 is also a risk factor for preeclampsia and placental abruption[41] (Table 2). In addition, quantitative analysis of the placenta indicated that the villus area was decreased significantly in the IUGR group induced by smoking exposure. The capillary area in the villus was also significantly less in IUGR groups. There were also some important changes observed under the electronic microscope, such as the retardation and reduced number of microvilli, fatty degeneration and mitochondrial swelling of the syncytial cells[42]. Studies on human placenta showed that villus volume and capillary area in IUGR were significantly decreased (Figure 2 and Table 4).

Table 4 AGT genotype and placental findings[42].
AGT genotypeMMMTTT
No. of placentas81314
Clinical findings
Maternal age (yr)26.0 ± 4.726.6 ± 6.629.8 ± 6.5
Gestational age (wk)36.2 ± 4.536.6 ± 1.936.7 ± 3.1
Fetal birth wt. (g)2730 ± 9672642 ± 5412620 ± 535
Placental quantitative findings:
Number of villi (mm2)155.3 ± 14.1145.8 ± 28.0146.3 ± 27.7
Villous CS area (μm2/villous)4422.2 ± 550.04400.9 ± 813.54248.6 ± 1191.9
Villous volume/1 cm3 placenta (cm3)0.668 ± 0.0340.626 ± 0.022b0.587 ± 0.059bc
Capillary volume/1 cm3 placenta (cm3)0.131 ± 0.0290.107 ± 0.0340.070 ± 0.030b
Intervillous volume/1 cm3 placenta (cm3)0.332 ± 0.0340.374 ± 0.022b0.413 ± 0.059bc
Placental quantitative analysis:
Volume of trimmed placenta (cm3)324.8 ± 128.0355.2 ± 80.3374.1 ± 70.2
Villous total volume per placenta (cm3)215.8 ± 81.8222.1 ± 50.7219.9 ± 52.6
Intervillous space per placenta (cm3)108.9 ± 47.8133.1 ± 31.1154.2 ± 38.6a
Villous capillary volume per placenta (cm3)45.1 ± 27.441.4 ± 23.226.6 ± 14.4a
Percentage of villous capillary volume (%)19.81 ± 5.1217.41 ± 7.312.06 ± 5.45a
Villous surface area per placenta (m2)9.029 ± 3.2859.560 ± 2.03110.370 ± 2.725
Villous surface per 1 g plac. Villi (cm2)390.6 ± 35.0401.6 ± 40.5438.6 ± 73.8
Other gene mutations

It has been reported in many publications that certain gene mutations are associated with IUGR (Table 3). Pericentric inversion of chromosome 6 causes haplo-insufficiency of the CDK19 gene resulting in microcephaly, congenital bilateral falciform retinal folds, nystagmus, and mental retardation[9]. Chromosome 1p32-p31-deletion syndrome and haplo-insufficiency of the NFIA gene may present as ventriculomegaly, corpus callosum hypogenesis, abnormal external genitalia, and IUGR in the third trimester[10]. Smigiel reported on two brothers affected with restrictive dermopathy, who died in the neonatal period[43]. Molecular analyses were performed in the second child, for whom biological material was available, and on both parents. Compound heterozygous frameshifting mutations were identified in exon 1 (c.50delA) and exon 5 (c.584_585delAT) of the ZMPSTE24 gene. The autosomal recessive inheritance was confirmed by genomic analysis of the parents. Partial loss of Ascl2 function affects all three layers of the mature placenta and causes IUGR[44]. The expression of c-fos is critical in the oxidative stress pathway. In fetal alcohol syndrome in mouse, it was shown that alcohol administration during pregnancy results in differential gene expression in the stress signal pathway, particularly in c-fos. C-fos expression in the decidua increases from 6 to 24 h after alcohol injection, but does not change in the embryo, which may contribute to alcohol-induced damage in fetal alcohol syndrome (Table 5)[45].

Table 5 Gene and fetal growth.
GeneGene mutation/expressionPhneotypic effects/complications
CDK19Chromosome breakpoints in 6p12.1 and 6q21Microcephaly, congenital bilateral falciform retinal folds, nystagmus, and mental retardation[9]
NFIAChromosome 1p32-p31 deletion syndromeVentriculomegaly, corpus callosum hypogenesis, abnormal external genitalia, and intrauterine growth restriction in the third trimester[44]
IGF1RNovel c.420del mutation in exon 2 of the IGF1R geneReduced IGF1R expression and represents haploinsufficiency of the IGF1R gene. IUGR and neonatal growth retardation[10]
hPGH, IGF-I, IGFBP-1Decreased expressionDecreased expression is associated with IUGR[27]
c-fosDecreased expressionFetal alcohol syndrome[45]
11b-HSD2Glucocorticoid metabolismUnder expression cause IUGR, small placenta[55]
GSTP1Glutathione transferase enzymes pathwayFetal growth and neonatal growth[56]
ZMPSTE24Fetal growthIUGR, dermopathy, neonatal death[43]
Ascl2Placenta developmentThree layers malformation, IUGR[44]
TFRCTransferrin receptor functionIUGR[13]
DIO3Type 3 deiodinaseHighly expressed in placenta and fetus. IUGR and hypothyroidism[47]
DLK1Growth promoterExpressed in placental villi. Methylation defects associated with IUGR[16]
HYMAINon-coding RNATransient neonatal diabetes and IUGR[57]
IGF2Growth FactorPlacental and fetal growth restriction[58]
KCNQ1OT1Non-coding RNAControl placental Kcnq1 domain. Involved in Beckwith-Wiedemann syndrome[59]
MAGEL2/NDNL1Similarity to NDNNeonatal growth retardation, alter metabolism[60]
MESTNeuronal differentiationFetal growth restriction, smaller placentas[52]
PEG10Retrotransposon-derivedgeneSevere growth retardation, absence of spongiotrophoblast layer, embryonic lethality[61]
PEG3Inhibits WNT-signallingPlacental and fetal growth restriction and abnormal maternal behavior[62]
PLAGL1 Zac1Tumor suppressorSkeletal defects, neonatal lethality, IUGR, and disrupted transactivation of Igf2[63]
SFRP2WNT signalingReduction in vitro of extra villous trophoblast invasion[64]
HBII-85/PWScrC/D Box small RNAImplicated in Prader Willi, Postnatal growth retardation[65]
Epigenetics of fetal growth

Mechanisms leading to the attenuation of fetal birth weight and adverse pregnancy outcomes are complex. Many studies have begun to focus, not only on the contribution of maternal and fetal genes to phenotypic outcome, but also on epigenetic changes associated with fetal growth[46].

Imprinted genes have a central role in the development and function of the placenta and have been implicated in a variety of disorders affecting fetal growth[14-16,47,48]. Gene inactivation studies in mice and chromosomal rearrangements in humans have demonstrated that many of these imprinted genes play key roles in placental development and function as well as in fetal growth. Those studies have also demonstrated that imprinted genes act in a complex manner at many levels to affect the energy balance between the mother and fetus. More recent studies also support a role of imprinted genes in a compensatory response to reduced fetal growth in humans.

During early embryonic development, the first stage of embryo differentiation establishes two cell lineages. These are the inner cell mass that forms all the tissues of the adult, and the trophectoderm that eventually produces placental structures. Generally speaking, the inner cell mass becomes gene hypermethylated and the trophectoderm becomes gene hypomethylated. These methylation patterns may be preserved throughout the whole pregnancy period[49,50]. Mouse and human models suggest the epigenetic regulation of fetal growth may also play a significant role through the placental imprinted genes.

Lambertini et al[48] investigated the differential methylation status of the imprinted genome to gain insights into the importance of the epigenetic regulation of these genes in fetal development by comparing IUGR with normal placentas. They found that differential methylation showed a highly significant correlation with gene length. The data also suggests that differential methylation changes in growth-restricted placentas occur throughout the genomic regions encompassing genes actively expressed in the placenta. Kumar studied 11 IGF related genes and found upregulated ZNF127 gene expression and down regulated PHLDA2 gene expression. This change confirmed an increased placental expression of growth-promoting imprinted genes and decreased expression of growth-suppressive imprinted genes with advancing gestational age. These changes in placental gene expression could potentially explain accelerated fetal growth seen in the third trimester[51]. McMinn presented an excellent example of the complexity of the imprint gene in placenta[52]. He showed that a small group of imprinted genes (PHLDA2, MEST, MEG3, GATM, GNAS and PLAGL1) and an additional 400 genes were affected in IUGR placentas.

Many studies have reported changes in imprinted gene expression and methylation levels in response to IUGR. These include reports of abnormal methylation in multiple imprinted loci[16], IUGR placental methylation decrease in the IGF2/H19 locus[53], and differential expression in multiple imprinted genes in IUGR placentas[54].

PROBLEMS AND FUTURE DIRECTION
IUGR animal models and clinical IUGR

IUGR animal models have been established by several methods. Most of the methods affect nutrients, placental function, or maternal fetal circulation although some may affect gene expression. These IUGR models have provided valuable simulations allowing study human IUGR, especially the IUGR resulting from nutritional factors, placental function, or toxic materials. However, no animal IUGR model induced by gene mutation or altered gene expression, simulating human IUGR related to specific genes has been reported. To date, the development of gene engineering and transgenic techniques have made it possible to produce specific gene mutation/expression altered IUGR animal models. This kind of animal model will provide a direct and definite method to study the specific genes involved in IUGR.

Single gene study and many genes related to IUGR

A concern of the present studies is that only limited numbers of genes have been studied among the thousands genes that are expressed in placenta, while many genes are in fact associated with fetal growth. Evaluation of their important roles in fetal growth still requires the accumulation of multiple collections of data on the mutation, phenotype, epigenetics, and metabolomics associated with fetal growth characteristics. Specifically, the pathways involved in the mutation or epigenetics resulting in IUGR phenotype are mostly unknown. Basic study of the pathways involving these genes may help us to understand why and how the phenomenon occurs at the molecular level.

Role of genes in IUGR

Although it seems simple to define the diagnosis of IUGR as an estimated fetal weight falling below the 10th percentile of normal body weight, that definition also captures births that are part of the normal variation in a population. Clinical diagnostic criteria are based on the fetal growth curve and umbilical blood velocity abnormalities. However, IUGR is still an extremely complex phenotype to dissect because of the many factors involved, maternal, fetal, placental, and environmental. Almost all recent gene expression studies are based on small number of samples. It is unrealistic to expect to find one or a few genes responsible for causing IUGR using this approach. Most gene expression studies have found genes that are dysregulated in IUGR. In addition, most of these genes, such as IGF-I, IGFBP1, corticotropin-releasing hormone, are reported to be related to the regulation of cell division and proliferation. We are still unable to to determine whether this is a compensatory response to restricted fetal restricted growth or a the factors which induces fetal growth restriction. The use of transgenic engineering technique in animals to study some specific gene may provide an ideal model to study the phenotype related to these genes in human clinical subjects.

Gene promoting and IUGR

Some specific imprinted genes are related to IUGR. These genes can be separated into two categories: genes participating in reducing fetal growth, and those which increase fetal growth as a compensatory response when it is sensed that the fetus is at risk. The evidence supporting mutations in imprinted genes that negatively influence fetal growth is summarized in the table and genes which the positively influence fetal growth are also indicated. We may expect that reducing expression of the negative genes and increase the expression of the positive genes would result in fetal growth. However, there is still a long way to go from basic animal model studies to reach clinical applications using gene-regulating techniques.

Footnotes

Peer reviewer: Chie-Pein Chen, MD, PhD, Division of High Risk Pregnancy, Mackay Memorial Hospital, 92, Sec. 2 Zhong-Shan North Road, Taipei 104, Taiwan

S- Editor Wang JL L- Editor Hughes D E- Editor Zheng XM

References
1.  Romo A, Carceller R, Tobajas J. Intrauterine growth retardation (IUGR): epidemiology and etiology. Pediatr Endocrinol Rev. 2009;6 Suppl 3:332-336.  [PubMed]  [DOI]  [Cited in This Article: ]
2.  Pallotto EK, Kilbride HW. Perinatal outcome and later implications of intrauterine growth restriction. Clin Obstet Gynecol. 2006;49:257-269.  [PubMed]  [DOI]  [Cited in This Article: ]
3.  Jones JE, Jurgens JA, Evans SA, Ennis RC, Villar VA, Jose PA. Mechanisms of fetal programming in hypertension. Int J Pediatr. 2012;2012:584831.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 13]  [Cited by in F6Publishing: 13]  [Article Influence: 1.1]  [Reference Citation Analysis (0)]
4.  Uma R, Forsyth JS, Struthers AD, Fraser CG, Godfrey V, Murphy DJ. Polymorphisms of the angiotensin converting enzyme gene in relation to intrauterine growth restriction. Acta Obstet Gynecol Scand. 2010;89:1197-1201.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 2]  [Cited by in F6Publishing: 2]  [Article Influence: 0.1]  [Reference Citation Analysis (0)]
5.  Kensara OA, Wootton SA, Phillips DI, Patel M, Jackson AA, Elia M. Fetal programming of body composition: relation between birth weight and body composition measured with dual-energy X-ray absorptiometry and anthropometric methods in older Englishmen. Am J Clin Nutr. 2005;82:980-987.  [PubMed]  [DOI]  [Cited in This Article: ]
6.  van Meer H, van Straten EM, Baller JF, van Dijk TH, Plösch T, Kuipers F, Verkade HJ. The effects of intrauterine malnutrition on maternal-fetal cholesterol transport and fetal lipid synthesis in mice. Pediatr Res. 2010;68:10-15.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 9]  [Cited by in F6Publishing: 9]  [Article Influence: 0.6]  [Reference Citation Analysis (0)]
7.  Battaglia FC. Placental transport: a function of permeability and perfusion. Am J Clin Nutr. 2007;85:591S-597S.  [PubMed]  [DOI]  [Cited in This Article: ]
8.  Randhawa R, Cohen P. The role of the insulin-like growth factor system in prenatal growth. Mol Genet Metab. 2005;86:84-90.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 165]  [Cited by in F6Publishing: 155]  [Article Influence: 8.2]  [Reference Citation Analysis (0)]
9.  Mukhopadhyay A, Kramer JM, Merkx G, Lugtenberg D, Smeets DF, Oortveld MA, Blokland EA, Agrawal J, Schenck A, van Bokhoven H. CDK19 is disrupted in a female patient with bilateral congenital retinal folds, microcephaly and mild mental retardation. Hum Genet. 2010;128:281-291.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 43]  [Cited by in F6Publishing: 42]  [Article Influence: 3.0]  [Reference Citation Analysis (0)]
10.  Chen CP, Su YN, Chen YY, Chern SR, Liu YP, Wu PC, Lee CC, Chen YT, Wang W. Chromosome 1p32-p31 deletion syndrome: prenatal diagnosis by array comparative genomic hybridization using uncultured amniocytes and association with NFIA haploinsufficiency, ventriculomegaly, corpus callosum hypogenesis, abnormal external genitalia, and intrauterine growth restriction. Taiwan J Obstet Gynecol. 2011;50:345-352.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 21]  [Cited by in F6Publishing: 25]  [Article Influence: 2.1]  [Reference Citation Analysis (0)]
11.  Börzsönyi B, Demendi C, Rigó J Jr, Szentpéteri I, Rab A, Joó JG. The regulation of apoptosis in intrauterine growth restriction: a study of Bcl-2 and Bax gene expression in human placenta. J Matern Fetal Neonatal Med. 2012;Epub ahead of print.  [PubMed]  [DOI]  [Cited in This Article: ]
12.  Chelbi ST, Wilson ML, Veillard AC, Ingles SA, Zhang J, Mondon F, Gascoin-Lachambre G, Doridot L, Mignot TM, Rebourcet R. Genetic and epigenetic mechanisms collaborate to control SERPINA3 expression and its association with placental diseases. Hum Mol Genet. 2012;21:1968-1978.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 58]  [Cited by in F6Publishing: 71]  [Article Influence: 5.9]  [Reference Citation Analysis (0)]
13.  Mandò C, Tabano S, Colapietro P, Pileri P, Colleoni F, Avagliano L, Doi P, Bulfamante G, Miozzo M, Cetin I. Transferrin receptor gene and protein expression and localization in human IUGR and normal term placentas. Placenta. 2011;32:44-50.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 132]  [Cited by in F6Publishing: 142]  [Article Influence: 10.1]  [Reference Citation Analysis (0)]
14.  He H, Liyanarachchi S, Akagi K, Nagy R, Li J, Dietrich RC, Li W, Sebastian N, Wen B, Xin B. Mutations in U4atac snRNA, a component of the minor spliceosome, in the developmental disorder MOPD I. Science. 2011;332:238-240.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 188]  [Cited by in F6Publishing: 190]  [Article Influence: 14.6]  [Reference Citation Analysis (0)]
15.  Edery P, Marcaillou C, Sahbatou M, Labalme A, Chastang J, Touraine R, Tubacher E, Senni F, Bober MB, Nampoothiri S. Association of TALS developmental disorder with defect in minor splicing component U4atac snRNA. Science. 2011;332:240-243.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 168]  [Cited by in F6Publishing: 169]  [Article Influence: 13.0]  [Reference Citation Analysis (0)]
16.  Turner CL, Mackay DM, Callaway JL, Docherty LE, Poole RL, Bullman H, Lever M, Castle BM, Kivuva EC, Turnpenny PD. Methylation analysis of 79 patients with growth restriction reveals novel patterns of methylation change at imprinted loci. Eur J Hum Genet. 2010;18:648-655.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 73]  [Cited by in F6Publishing: 79]  [Article Influence: 5.6]  [Reference Citation Analysis (0)]
17.  Vogt E, Ng AK, Rote NS. A model for the antiphospholipid antibody syndrome: monoclonal antiphosphatidylserine antibody induces intrauterine growth restriction in mice. Am J Obstet Gynecol. 1996;174:700-707.  [PubMed]  [DOI]  [Cited in This Article: ]
18.  Zhang XQ, Yan JH, Hong SY, Lin QD, Zhao AM, Huang LJ. [Establishment of intrauterine growth restriction using passive smoking]. Shanghai Dier Yike Daxue Xuebao. 1993;13:314-317.  [PubMed]  [DOI]  [Cited in This Article: ]
19.  Hayashi TT, Dorko ME. A rat model for the study of intrauterine growth retardation. Am J Obstet Gynecol. 1988;158:1203-1207.  [PubMed]  [DOI]  [Cited in This Article: ]
20.  Zhang XQ. [Assay of plasma amino acid concentration in maternal and growth-retarded fetuses caused by passive smoking in rabbits]. Zhonghua Fuchanke Zazhi. 1993;28:582-585, 633.  [PubMed]  [DOI]  [Cited in This Article: ]
21.  Zhang XQ. [The change in dynamics of cell proliferation cycle by passive smoking in fetal rabbit]. Zhonghua Fuchanke Zazhi. 1993;28:720-722, 759.  [PubMed]  [DOI]  [Cited in This Article: ]
22.  Zhang XQ, Yan JH. [Plasma amino acids concentration in maternal and fetal rabbits before and in late pregnancy]. Zhongshan Yike Daxue Xuebao. 1995;16:32-35.  [PubMed]  [DOI]  [Cited in This Article: ]
23.  Greenwood PL, Slepetis RM, Hermanson JW, Bell AW. Intrauterine growth retardation is associated with reduced cell cycle activity, but not myofibre number, in ovine fetal muscle. Reprod Fertil Dev. 1999;11:281-291.  [PubMed]  [DOI]  [Cited in This Article: ]
24.  Zhang X, Yan J, Hong S. [An experimental study on the treatment of fetal rabbits with intrauterine growth retardation]. Zhonghua Fuchanke Zazhi. 1996;31:97-99.  [PubMed]  [DOI]  [Cited in This Article: ]
25.  Street ME, Seghini P, Fieni S, Ziveri MA, Volta C, Martorana D, Viani I, Gramellini D, Bernasconi S. Changes in interleukin-6 and IGF system and their relationships in placenta and cord blood in newborns with fetal growth restriction compared with controls. Eur J Endocrinol. 2006;155:567-574.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 82]  [Cited by in F6Publishing: 86]  [Article Influence: 4.8]  [Reference Citation Analysis (0)]
26.  Verkauskiene R, Beltrand J, Claris O, Chevenne D, Deghmoun S, Dorgeret S, Alison M, Gaucherand P, Sibony O, Lévy-Marchal C. Impact of fetal growth restriction on body composition and hormonal status at birth in infants of small and appropriate weight for gestational age. Eur J Endocrinol. 2007;157:605-612.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 60]  [Cited by in F6Publishing: 63]  [Article Influence: 3.7]  [Reference Citation Analysis (0)]
27.  Koutsaki M, Sifakis S, Zaravinos A, Koutroulakis D, Koukoura O, Spandidos DA. Decreased placental expression of hPGH, IGF-I and IGFBP-1 in pregnancies complicated by fetal growth restriction. Growth Horm IGF Res. 2011;21:31-36.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 49]  [Cited by in F6Publishing: 48]  [Article Influence: 3.7]  [Reference Citation Analysis (0)]
28.  Kawashima Y, Takahashi S, Kanzaki S. Familial short stature with IGF-I receptor gene anomaly. Endocr J. 2012;59:179-185.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 12]  [Cited by in F6Publishing: 14]  [Article Influence: 1.2]  [Reference Citation Analysis (0)]
29.  Abuzzahab MJ, Schneider A, Goddard A, Grigorescu F, Lautier C, Keller E, Kiess W, Klammt J, Kratzsch J, Osgood D. IGF-I receptor mutations resulting in intrauterine and postnatal growth retardation. N Engl J Med. 2003;349:2211-2222.  [PubMed]  [DOI]  [Cited in This Article: ]
30.  Kruis T, Klammt J, Galli-Tsinopoulou A, Wallborn T, Schlicke M, Müller E, Kratzsch J, Körner A, Odeh R, Kiess W. Heterozygous mutation within a kinase-conserved motif of the insulin-like growth factor I receptor causes intrauterine and postnatal growth retardation. J Clin Endocrinol Metab. 2010;95:1137-1142.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 59]  [Cited by in F6Publishing: 63]  [Article Influence: 4.5]  [Reference Citation Analysis (0)]
31.  Inagaki K, Tiulpakov A, Rubtsov P, Sverdlova P, Peterkova V, Yakar S, Terekhov S, LeRoith D. A familial insulin-like growth factor-I receptor mutant leads to short stature: clinical and biochemical characterization. J Clin Endocrinol Metab. 2007;92:1542-1548.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 64]  [Cited by in F6Publishing: 62]  [Article Influence: 3.6]  [Reference Citation Analysis (0)]
32.  Walenkamp MJ, van der Kamp HJ, Pereira AM, Kant SG, van Duyvenvoorde HA, Kruithof MF, Breuning MH, Romijn JA, Karperien M, Wit JM. A variable degree of intrauterine and postnatal growth retardation in a family with a missense mutation in the insulin-like growth factor I receptor. J Clin Endocrinol Metab. 2006;91:3062-3070.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 92]  [Cited by in F6Publishing: 102]  [Article Influence: 5.7]  [Reference Citation Analysis (0)]
33.  Wallborn T, Wüller S, Klammt J, Kruis T, Kratzsch J, Schmidt G, Schlicke M, Müller E, van de Leur HS, Kiess W. A heterozygous mutation of the insulin-like growth factor-I receptor causes retention of the nascent protein in the endoplasmic reticulum and results in intrauterine and postnatal growth retardation. J Clin Endocrinol Metab. 2010;95:2316-2324.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 52]  [Cited by in F6Publishing: 57]  [Article Influence: 4.1]  [Reference Citation Analysis (0)]
34.  Choi JH, Kang M, Kim GH, Hong M, Jin HY, Lee BH, Park JY, Lee SM, Seo EJ, Yoo HW. Clinical and functional characteristics of a novel heterozygous mutation of the IGF1R gene and IGF1R haploinsufficiency due to terminal 15q26.2-& gt; qter deletion in patients with intrauterine growth retardation and postnatal catch-up growth failure. J Clin Endocrinol Metab. 2011;96:E130-E134.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 42]  [Cited by in F6Publishing: 45]  [Article Influence: 3.5]  [Reference Citation Analysis (0)]
35.  Umbers AJ, Boeuf P, Clapham C, Stanisic DI, Baiwog F, Mueller I, Siba P, King CL, Beeson JG, Glazier J. Placental malaria-associated inflammation disturbs the insulin-like growth factor axis of fetal growth regulation. J Infect Dis. 2011;203:561-569.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 66]  [Cited by in F6Publishing: 66]  [Article Influence: 5.1]  [Reference Citation Analysis (0)]
36.  Baker J, Liu JP, Robertson EJ, Efstratiadis A. Role of insulin-like growth factors in embryonic and postnatal growth. Cell. 1993;75:73-82.  [PubMed]  [DOI]  [Cited in This Article: ]
37.  Liu JP, Baker J, Perkins AS, Robertson EJ, Efstratiadis A. Mice carrying null mutations of the genes encoding insulin-like growth factor I (Igf-1) and type 1 IGF receptor (Igf1r). Cell. 1993;75:59-72.  [PubMed]  [DOI]  [Cited in This Article: ]
38.  Netchine I, Azzi S, Le Bouc Y, Savage MO. IGF1 molecular anomalies demonstrate its critical role in fetal, postnatal growth and brain development. Best Pract Res Clin Endocrinol Metab. 2011;25:181-190.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 64]  [Cited by in F6Publishing: 64]  [Article Influence: 4.9]  [Reference Citation Analysis (0)]
39.  Morgan T, Craven C, Lalouel JM, Ward K. Angiotensinogen Thr235 variant is associated with abnormal physiologic change of the uterine spiral arteries in first-trimester decidua. Am J Obstet Gynecol. 1999;180:95-102.  [PubMed]  [DOI]  [Cited in This Article: ]
40.  Zhang XQ, Varner M, Dizon-Townson D, Song F, Ward K. A molecular variant of angiotensinogen is associated with idiopathic intrauterine growth restriction. Obstet Gynecol. 2003;101:237-242.  [PubMed]  [DOI]  [Cited in This Article: ]
41.  Zhang XQ, Craven C, Nelson L, Varner MW, Ward KJ. Placental abruption is more frequent in women with the angiotensinogen Thr235 mutation. Placenta. 2007;28:616-619.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 11]  [Cited by in F6Publishing: 10]  [Article Influence: 0.6]  [Reference Citation Analysis (0)]
42.  Zhang X, Yan J, Hong S. [The quantitative and morphological analysis of placenta and brain in rabbits with intrauterine fetal growth retardation]. Zhonghua Fuchanke Zazhi. 1995;30:724-727.  [PubMed]  [DOI]  [Cited in This Article: ]
43.  Smigiel R, Jakubiak A, Esteves-Vieira V, Szela K, Halon A, Jurek T, Lévy N, De Sandre-Giovannoli A. Novel frameshifting mutations of the ZMPSTE24 gene in two siblings affected with restrictive dermopathy and review of the mutations described in the literature. Am J Med Genet A. 2010;152A:447-452.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 36]  [Cited by in F6Publishing: 37]  [Article Influence: 2.6]  [Reference Citation Analysis (0)]
44.  Oh-McGinnis R, Bogutz AB, Lefebvre L. Partial loss of Ascl2 function affects all three layers of the mature placenta and causes intrauterine growth restriction. Dev Biol. 2011;351:277-286.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 50]  [Cited by in F6Publishing: 43]  [Article Influence: 3.3]  [Reference Citation Analysis (0)]
45.  Poggi SH, Goodwin KM, Hill JM, Brenneman DE, Tendi E, Schninelli S, Spong CY. Differential expression of c-fos in a mouse model of fetal alcohol syndrome. Am J Obstet Gynecol. 2003;189:786-789.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 9]  [Cited by in F6Publishing: 10]  [Article Influence: 0.5]  [Reference Citation Analysis (0)]
46.  Suter M, Abramovici A, Aagaard-Tillery K. Genetic and epigenetic influences associated with intrauterine growth restriction due to in utero tobacco exposure. Pediatr Endocrinol Rev. 2010;8:94-102.  [PubMed]  [DOI]  [Cited in This Article: ]
47.  Hernandez A, Martinez ME, Fiering S, Galton VA, St Germain D. Type 3 deiodinase is critical for the maturation and function of the thyroid axis. J Clin Invest. 2006;116:476-484.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 243]  [Cited by in F6Publishing: 231]  [Article Influence: 12.8]  [Reference Citation Analysis (0)]
48.  Lambertini L, Lee TL, Chan WY, Lee MJ, Diplas A, Wetmur J, Chen J. Differential methylation of imprinted genes in growth-restricted placentas. Reprod Sci. 2011;18:1111-1117.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 32]  [Cited by in F6Publishing: 34]  [Article Influence: 2.6]  [Reference Citation Analysis (0)]
49.  Santos F, Hendrich B, Reik W, Dean W. Dynamic reprogramming of DNA methylation in the early mouse embryo. Dev Biol. 2002;241:172-182.  [PubMed]  [DOI]  [Cited in This Article: ]
50.  Dean W, Ferguson-Smith A. Genomic imprinting: mother maintains methylation marks. Curr Biol. 2001;11:R527-R530.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 22]  [Cited by in F6Publishing: 23]  [Article Influence: 1.0]  [Reference Citation Analysis (0)]
51.  Kumar N, Leverence J, Bick D, Sampath V. Ontogeny of growth-regulating genes in the placenta. Placenta. 2012;33:94-99.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 25]  [Cited by in F6Publishing: 25]  [Article Influence: 1.9]  [Reference Citation Analysis (0)]
52.  McMinn J, Wei M, Schupf N, Cusmai J, Johnson EB, Smith AC, Weksberg R, Thaker HM, Tycko B. Unbalanced placental expression of imprinted genes in human intrauterine growth restriction. Placenta. 2006;27:540-549.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 229]  [Cited by in F6Publishing: 211]  [Article Influence: 11.1]  [Reference Citation Analysis (0)]
53.  Tabano S, Colapietro P, Cetin I, Grati FR, Zanutto S, Mandò C, Antonazzo P, Pileri P, Rossella F, Larizza L. Epigenetic modulation of the IGF2/H19 imprinted domain in human embryonic and extra-embryonic compartments and its possible role in fetal growth restriction. Epigenetics. 2010;5:313-324.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 86]  [Cited by in F6Publishing: 93]  [Article Influence: 6.6]  [Reference Citation Analysis (0)]
54.  Diplas AI, Lambertini L, Lee MJ, Sperling R, Lee YL, Wetmur J, Chen J. Differential expression of imprinted genes in normal and IUGR human placentas. Epigenetics. 2009;4:235-240.  [PubMed]  [DOI]  [Cited in This Article: ]
55.  Börzsönyi B, Demendi C, Pajor A, Rigó J, Marosi K, Agota A, Nagy ZB, Joó JG. Gene expression patterns of the 11β-hydroxysteroid dehydrogenase 2 enzyme in human placenta from intrauterine growth restriction: the role of impaired feto-maternal glucocorticoid metabolism. Eur J Obstet Gynecol Reprod Biol. 2012;161:12-17.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 24]  [Cited by in F6Publishing: 25]  [Article Influence: 2.1]  [Reference Citation Analysis (0)]
56.  Slama R, Gräbsch C, Lepeule J, Siroux V, Cyrys J, Sausenthaler S, Herbarth O, Bauer M, Borte M, Wichmann HE. Maternal fine particulate matter exposure, polymorphism in xenobiotic-metabolizing genes and offspring birth weight. Reprod Toxicol. 2010;30:600-612.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 14]  [Cited by in F6Publishing: 11]  [Article Influence: 0.8]  [Reference Citation Analysis (0)]
57.  Ma D, Shield JP, Dean W, Leclerc I, Knauf C, Burcelin R R, Rutter GA, Kelsey G. Impaired glucose homeostasis in transgenic mice expressing the human transient neonatal diabetes mellitus locus, TNDM. J Clin Invest. 2004;114:339-348.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 110]  [Cited by in F6Publishing: 112]  [Article Influence: 5.6]  [Reference Citation Analysis (0)]
58.  Dilworth MR, Kusinski LC, Cowley E, Ward BS, Husain SM, Constância M, Sibley CP, Glazier JD. Placental-specific Igf2 knockout mice exhibit hypocalcemia and adaptive changes in placental calcium transport. Proc Natl Acad Sci USA. 2010;107:3894-3899.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 73]  [Cited by in F6Publishing: 74]  [Article Influence: 5.3]  [Reference Citation Analysis (0)]
59.  Mohammad F, Mondal T, Guseva N, Pandey GK, Kanduri C. Kcnq1ot1 noncoding RNA mediates transcriptional gene silencing by interacting with Dnmt1. Development. 2010;137:2493-2499.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 213]  [Cited by in F6Publishing: 207]  [Article Influence: 14.8]  [Reference Citation Analysis (0)]
60.  Bischof JM, Stewart CL, Wevrick R. Inactivation of the mouse Magel2 gene results in growth abnormalities similar to Prader-Willi syndrome. Hum Mol Genet. 2007;16:2713-2719.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 142]  [Cited by in F6Publishing: 144]  [Article Influence: 8.5]  [Reference Citation Analysis (0)]
61.  Ono R, Nakamura K, Inoue K, Naruse M, Usami T, Wakisaka-Saito N, Hino T, Suzuki-Migishima R, Ogonuki N, Miki H. Deletion of Peg10, an imprinted gene acquired from a retrotransposon, causes early embryonic lethality. Nat Genet. 2006;38:101-106.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 299]  [Cited by in F6Publishing: 301]  [Article Influence: 15.8]  [Reference Citation Analysis (0)]
62.  Murphy SK, Wylie AA, Jirtle RL. Imprinting of PEG3, the human homologue of a mouse gene involved in nurturing behavior. Genomics. 2001;71:110-117.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 42]  [Cited by in F6Publishing: 44]  [Article Influence: 1.9]  [Reference Citation Analysis (0)]
63.  Varrault A, Gueydan C, Delalbre A, Bellmann A, Houssami S, Aknin C, Severac D, Chotard L, Kahli M, Le Digarcher A. Zac1 regulates an imprinted gene network critically involved in the control of embryonic growth. Dev Cell. 2006;11:711-722.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 356]  [Cited by in F6Publishing: 311]  [Article Influence: 18.3]  [Reference Citation Analysis (0)]
64.  Guilleret I, Osterheld MC, Braunschweig R, Gastineau V, Taillens S, Benhattar J. Imprinting of tumor-suppressor genes in human placenta. Epigenetics. 2009;4:62-68.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 22]  [Cited by in F6Publishing: 24]  [Article Influence: 1.6]  [Reference Citation Analysis (0)]
65.  Skryabin BV, Gubar LV, Seeger B, Pfeiffer J, Handel S, Robeck T, Karpova E, Rozhdestvensky TS, Brosius J. Deletion of the MBII-85 snoRNA gene cluster in mice results in postnatal growth retardation. PLoS Genet. 2007;3:e235.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 141]  [Cited by in F6Publishing: 147]  [Article Influence: 9.2]  [Reference Citation Analysis (0)]