Skip to main content Accessibility help
×
Hostname: page-component-8448b6f56d-c4f8m Total loading time: 0 Render date: 2024-04-19T15:31:14.286Z Has data issue: false hasContentIssue false

10 - Trophoblast and pre-eclampsia

from General discussion II

Published online by Cambridge University Press:  07 August 2009

C. W. G. Redman
Affiliation:
Nuffield Department of Obstetrics and Gynaecology, University of Oxford, UK
I. L. Sargent
Affiliation:
Nuffield Department of Obstetrics and Gynaecology, University of Oxford, UK
E .A. Linton
Affiliation:
Nuffield Department of Obstetrics and Gynaecology, University of Oxford, UK
Ashley Moffett
Affiliation:
University of Cambridge
Charlie Loke
Affiliation:
University of Cambridge
Anne McLaren
Affiliation:
Cancer Research, UK
Get access

Summary

Introduction

In this chapter, the importance of trophoblast in generating the maternal signs and symptoms that comprise the syndrome of pre-eclampsia will be discussed, but not the processes of placentation in early pregnancy that may be important in the preclinical phases of the disorder. pre-eclampsia is a potentially dangerous and highly variable complication of the second half of pregnancy, labour or the early puerperium. It has been known for nearly 100 years that it originates in the placenta (Holland 1909). The presence of a placenta is both necessary and sufficient to cause the disorder (Redman 1991). A fetus is not required as pre-eclampsia can occur with hydatidiform mole (Chun et al. 1964). A uterus is probably not required because pre-eclampsia may develop with abdominal pregnancy (Piering et al. 1993). Central to management is delivery, which removes the causative organ, namely the placenta.

The maternal illness of pre-eclampsia was originally thought to be caused primarily by generalised maternal endothelial activation and dysfunction (Roberts et al. 1989). Later this concept was broadened by incorporating endothelial dysfunction as one of several components of a maternal systemic inflammatory response in pre-eclampsia (Redman et al. 1999).

A key feature is that systemic inflammation is not only characteristic of the pre-eclamptic woman but an intrinsic part of every normal pregnancy, which becomes most evident in the third trimester (Redman et al. 1999). The inflammatory response of pre-eclampsia does not differ in type, but only in degree, being more intense.

Type
Chapter
Information
Publisher: Cambridge University Press
Print publication year: 2006

Access options

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

References

Abe, M., Shintani, Y., Eto, Y.et al. (2002). Potent induction of activin A secretion from monocytes and bone marrow stromal fibroblasts by cognate interaction with activated T cells. J. Leukocyte Biol., 72, 347–52.Google ScholarPubMed
Ahmed, A., Dunk, C., Ahmad, S. & Khaliq, A. (2000a). Regulation of placental vascular endothelial growth factor (VEGF) and placenta growth factor (PIGF) and soluble Flt-1 by oxygen – a review. Placenta, 21 (Suppl A), S16–24.CrossRefGoogle Scholar
Ahmed, I., Perkins, A. V., Glynn, B. P.et al. (2000b). Processing of procorticotrophin releasing hormone (proCRH): molecular forms of CRH in placentae from normal and pre-eclamptic pregnancies. J. Clin. Endocrinol. Metab., 85, 755–64.Google Scholar
Allaire, A. D., Ballenger, K. A., Wells, S. R., McMahon, M. J. & Lessey, B. A. (2000). Placental apoptosis in preeclampsia. Obstet. Gynecol., 96, 271–6.Google ScholarPubMed
Alsat, E., Wyplosz, P., Malassine, A.et al. (1996). Hypoxia impairs cell fusion and differentiation process in human cytotrophoblast, in vitro. J. Cell. Physiol., 168, 346–53.3.0.CO;2-1>CrossRefGoogle ScholarPubMed
Arnalich, F., Lopez, J., Codoceo, R.et al. (1999). Relationship of plasma leptin to plasma cytokines and human survival in sepsis and septic shock. J. Infect. Dis., 180, 908–11.CrossRefGoogle ScholarPubMed
Aupeix, K., Hugel, B., Martin, T.et al. (1997). The significance of shed membrane particles during programmed cell death in vitro, and in vivo, in HIV-1 infection. J. Clin. Invest., 99, 1546–54.CrossRefGoogle ScholarPubMed
Austgulen, R., Isaksen, C. V., Chedwick, L.et al. (2004). Preeclampsia: associated with increased syncytial apoptosis when the infant is small-for-gestational-age. J. Reprod. Immunol., 61, 39–50.CrossRefGoogle ScholarPubMed
Barleon, B., Reusch, P., Totzke, F.et al. (2001). Soluble VEGFR-1 secreted by endothelial cells and monocytes is present in human serum and plasma from healthy donors. Angiogenesis, 4, 143–54.CrossRefGoogle ScholarPubMed
Barr, F. G., Biegel, J. A., Sellinger, B., Womer, R. B. & Emanuel, B. S. (1991). Molecular and cytogenetic analysis of chromosomal arms 2q and 13q in alveolar rhabdomyosarcoma. Genes Chromosomes Cancer, 3, 153–61.CrossRefGoogle ScholarPubMed
Belgore, F. M., Lip, G. Y. & Blann, A. D. (2000). Vascular endothelial growth factor and its receptor, Flt-1, in smokers and non-smokers. Br. J. Biomed. Sci., 57, 207–13.Google ScholarPubMed
Belgore, F. M., Blann, A. D., Li, S. H., Beevers, D. G. & Lip, G. Y. (2001a). Plasma levels of vascular endothelial growth factor and its soluble receptor (sFlt-1) in essential hypertension. Am. J. Cardiol., 87, 805–7.CrossRefGoogle Scholar
Belgore, F. M., Blann, A. D. & Lip, G. Y. (2001b). Measurement of free and complexed soluble vascular endothelial growth factor receptor, Flt-1, in fluid samples: development and application of two new immunoassays. Clin. Sci. (Lond.), 100, 567–75.CrossRefGoogle Scholar
Blond, J. L., Lavillette, D., Cheynet, V.et al. (2000). An envelope glycoprotein of the human endogenous retrovirus HERV-W is expressed in the human placenta and fuses cells expressing the type D mammalian retrovirus receptor. J. Virol., 74, 3321–9.CrossRefGoogle ScholarPubMed
Blumenstein, M., Mitchell, M. D., Groome, N. P. & Keelan, J. A. (2002). Hypoxia inhibits activin A production by term villous trophoblast in vitro. Placenta, 23, 735–41.CrossRefGoogle ScholarPubMed
Bowen, R. S., Moodley, J., Dutton, M. F. & Theron, A. J. (2001). Oxidative stress in preeclampsia. Acta Obstet. Gynecol. Scand., 80, 719–25.CrossRefGoogle Scholar
Boyd, P. A. & Scott, A. (1985). Quantitative structural studies on human placentae associated with preeclampsia, essential hypertension and intrauterine growth retardation. Br. J. Obstet. Gynaecol., 92, 714–21.CrossRefGoogle Scholar
Boyd, P. A., Lindenbaum, R. H. & Redman, C. W. G. (1987). Preeclampsia and trisomy 13: a possible association. Lancet, ii, 425–7.CrossRefGoogle Scholar
Brosens, I. & Renaer, M. (1972). On the pathogenesis of placental infarcts in preeclampsia. J. Obstet. Gynaecol. Br. Commonw., 79, 794–9.CrossRefGoogle Scholar
Brosens, I. A., Robertson, W. B. & Dixon, H. G. (1972). The role of the spiral arteries in the pathogenesis of preeclampsia. Obstet. Gynecol. Annu., 1, 177–91.Google ScholarPubMed
Brosens, I., Dixon, H. G. & Robertson, W. B. (1977). Fetal growth retardation and the arteries of the placental bed. Br. J. Obstet. Gynaecol., 84, 656–63.CrossRefGoogle ScholarPubMed
Campbell, E. A., Linton, E. A., Wolfe, C. D.et al. (1987). Plasma corticotropin-releasing hormone concentrations during pregnancy and parturitionJ. Clin. Endocrinol. Metab., 64, 1054–9.CrossRefGoogle ScholarPubMed
Cester, N., Staffolani, R., Rabini, R. A.et al. (1994). Pregnancy induced hypertension: a role for peroxidation in microvillus plasma membranes. Mol. Cell. Biochem., 131, 151–5.CrossRefGoogle ScholarPubMed
Chun, D., Braga, C., Chow, C. & Lok, L. (1964). Clinical observations on some aspects of hydatidiform moles. J. Obstet. Gynaecol. Br. Commonw., 71, 180–4.CrossRefGoogle ScholarPubMed
Clark, D. E., Smith, S. K., He, Y.et al. (1998). A vascular endothelial growth factor antagonist is produced by the human placenta and released into the maternal circulation. Biol. Reprod., 59, 1540–8.CrossRefGoogle ScholarPubMed
Coppack, S. W. (2001). Pro-inflammatory cytokines and adipose tissue. Proc. Nutr. Soc., 60, 349–56.CrossRefGoogle ScholarPubMed
Waard, V., van-den Berg, B., Veken, J.et al. (1999). Serial analysis of gene expression to assess the endothelial cell response to an atherogenic stimulus. Gene, 226, 1–8.CrossRefGoogle Scholar
Wolf, F., Robertson, W. B. & Brosens, I. (1975). The ultrastructure of acute atherosis in hypertensive pregnancy. Am. J. Obstet. Gynecol., 123, 164–74.CrossRefGoogle ScholarPubMed
Wolf, F., Brosens, I. & Renaer, M. (1980). Fetal growth retardation and the maternal arterial supply of the human placenta in the absence of sustained hypertension. Br. J. Obstet. Gynaecol., 87, 678–85.CrossRefGoogle ScholarPubMed
Wolf, F., Brosens, I. & Robertson, W. B. (1982). Ultrastructure of uteroplacental arteries. Contrib. Gynecol. Obstet., 9, 86–99.Google ScholarPubMed
Douglas, K. A. & Redman, C. W. (1994). Eclampsia in the United Kingdom. BMJ, 309, 1395–1400.CrossRefGoogle ScholarPubMed
Eramaa, M., Hurme, M., Stenman, U. H. & Ritvos, O. (1992). Activin A/erythroid differentiation factor is induced during human monocyte activation. J. Exp. Med., 176, 1449–52.CrossRefGoogle ScholarPubMed
Esterman, A., Greco, M. A., Mitani, Y.et al. (1997). The effect of hypoxia on human trophoblast in culture: morphology, glucose transport and metabolism. Placenta, 18, 129–36.CrossRefGoogle ScholarPubMed
Fantuzzi, G. & Faggioni, R. (2000). Leptin in the regulation of immunity, inflammation, and hematopoiesis. J. Leukocyte Biol., 68, 437–46.Google ScholarPubMed
Felmeden, D. C., Spencer, C. G., Belgore, F. M.et al. (2003). Endothelial damage and angiogenesis in hypertensive patients: relationship to cardiovascular risk factors and risk factor management. Am. J. Hypertens., 16, 11–20.CrossRefGoogle ScholarPubMed
Furui, T., Kurauchi, O., Tanaka, M., et al. (1994). Decrease in cytochrome c oxidase and cytochrome oxidase subunit I messenger RNA levels in preeclamptic pregnancies. Obstet. Gynecol., 84, 283–8.Google ScholarPubMed
Gaber, L. W., Spargo, B. H. & Lindheimer, M. D. (1994). Renal pathology in pre-eclampsia. Builliere's Clin. Obstet. Gynecol., 8, 443–68.CrossRefGoogle ScholarPubMed
Garner, P. R., D'Alton, M. E., Dudley, D. K., Huard, P. & Hardie, M. (1990). Preeclampsia in diabetic pregnancies. Am. J. Obstet. Gynecol., 163, 505–8.CrossRefGoogle ScholarPubMed
Gavrilova, O., Barr, V., Marcus, S. B. & Reitman, M. (1997). Hyperleptinemia of pregnancy associated with the appearance of a circulating form of the leptin receptor. J. Biol. Chem., 272, 30546–51.CrossRefGoogle ScholarPubMed
Germain, S. J., Knight, M., Sooranna, S. R., Redman, C. W. G. & Sargent, I. L. (2002). Interaction of circulating syncytiotrophoblast microvillous fragments with maternal monocytes in normal and pre-eclamptic pregnancies. J. Soc. Gynecol. Invest., 9 (Suppl), 259A.Google Scholar
Gratacos, E., Casals, E., Deulofeu, R., et al. (1998). Lipid peroxide and vitamin E patterns in pregnant women with different types of hypertension in pregnancy. Am. J. Obstet. Gynecol., 178, 1072–6.CrossRefGoogle ScholarPubMed
Grosfeld, A., Turban, S., Andre, J.et al. (2001). Transcriptional effect of hypoxia on placental leptin. FEBS Lett, 502, 122–6.CrossRefGoogle ScholarPubMed
Hansson, G. K., Libby, P., Schönbeck, U.Yan, Z. Q. (2002). Innate and adaptive immunity in the pathogenesis of atherosclerosis. Circ. Res., 91, 281–91.CrossRefGoogle ScholarPubMed
Hardy, D. B. & Yang, K. (2002). The expression of 11 β-hydroxysteroid dehydrogenase type 2 is induced during trophoblast differentiation: effects of hypoxia. J. Clin. Endocrinol. Metab., 87, 3696–701.Google ScholarPubMed
He, L., Wang, Z. & Sun, Y. (2004). Reduced amount of cytochrome c oxidase subunit I messenger RNA in placentae from pregnancies complicated by preeclampsia. Acta Obstet. Gynecol. Scand., 83, 144–8.CrossRefGoogle Scholar
Hertig, A. T. (1945). Vascular pathology in the hypertensive albuminuric toxemias of pregnancy. Clinics, 4, 602–14.Google Scholar
Ho, S., Winkler-Lowen, L. B., Morrish, D. W.et al. (1999). The role of Bcl-2 expression in EGF inhibition of TNF-alpha/IFN-gamma-induced villous trophoblast apoptosis. Placenta, 20, 423–30.CrossRefGoogle ScholarPubMed
Holland, E. (1909). Recent work on the aetiology of eclampsia. J. Obstet. Gynaecol. Br. Emp., 16, 255–73.CrossRefGoogle Scholar
Hornig, C., Barleon, B., Ahmad, S.et al. (2000). Release and complex formation of soluble VEGFR-1 from endothelial cells and biological fluids. Lab. Invest., 80, 443–54.CrossRefGoogle ScholarPubMed
Huang, L., Wang, Z. & Li, C. (2001). Modulation of circulating leptin levels by its soluble receptor. J. Biol. Chem., 276, 6343–9.CrossRefGoogle ScholarPubMed
Huppertz, B., Frank, H. G., Kingdom, J. C., Reister, F. & Kaufmann, P. (1998). Villous cytotrophoblast regulation of the syncytial apoptotic cascade in the human placenta. Histochem. Cell Biol., 110, 495–508.CrossRefGoogle ScholarPubMed
Huppertz, B., Kingdom, J., Caniggia, I.et al. (2003). Hypoxia favours necrotic versus apoptotic shedding of placental syncytiotrophoblast into the maternal circulation. Placenta, 24, 181–90.CrossRefGoogle ScholarPubMed
Ishihara, N., Matsuo, H., Murakoshi, H.et al. (2002). Increased apoptosis in the syncytiotrophoblast in human term placentae complicated by either preeclampsia or intrauterine growth retardation. Am. J. Obstet. Gynecol., 186, 158–66.CrossRefGoogle ScholarPubMed
Jones, C. J. & Fox, H. (1980). An ultrastructural and ultrahistochemical study of the human placenta in maternal preeclampsia. Placenta, 1, 61–76.CrossRefGoogle Scholar
Kabbinavar, F., Hurwitz, H. I., Fehrenbacher, L.et al. (2003). Phase II, randomized trial comparing bevacizumab plus fluorouracil (FU)/leucovorin (LV) with FU/LV alone in patients with metastatic colorectal cancer. J. Clin. Oncol., 21, 60–5.CrossRefGoogle ScholarPubMed
Karalis, K., Muglia, J. L., Bae, D., Hilderbrand, H. & Majzoub, J. A. (1997). CRH and the immune system. J. Neuroimmunol., 172, 131–6.CrossRefGoogle Scholar
Khong, T. Y. (1991). Acute atherosis in pregnancies complicated by hypertension, small-for-gestational-age infants, and diabetes mellitus. Arch. Pathol. Lab. Med., 115, 722–5.Google ScholarPubMed
Kliman, H. J., Nestler, J. E., Sermasi, E., Sanger, J. M. & Strauss, J. F. (1986). Purification, characterization, and in vitro differentiation of cytotrophoblasts from human term placentae. Endocrinology, 118, 1567–82.CrossRefGoogle ScholarPubMed
Kluck, R. M., Bossy, W. E., Green, D. R. & Newmeyer, D. D. (1997). The release of cytochrome c from mitochondria: a primary site for Bcl-2 regulation of apoptosis. Science, 275, 1132–6.CrossRefGoogle ScholarPubMed
Knapen, M. F., Peters, W. H., Mulder, T. P.et al. (1999). Glutathione and glutathione-related enzymes in decidua and placenta of controls and women with pre-eclampsia. Placenta, 20, 541–6.CrossRefGoogle ScholarPubMed
Knerr, I., Beinder, E. & Rascher, W. (2002). Syncytin, a novel human endogenous retroviral gene in human placenta: evidence for its dysregulation in pre-eclampsia and HELLPsyndrome. Am. J. Obstet. Gynecol., 186, 210–13.CrossRefGoogle ScholarPubMed
Knerr, I., Weigel, C., Linnemann, K.et al. (2003). Transcriptional effects of hypoxia on fusiogenic syncytin and its receptor ASCT2 in human cytotrophoblast BeWo cells and in ex vivo perfused placental cotyledons. Am. J. Obstet. Gynecol., 189, 583–8.CrossRefGoogle ScholarPubMed
Knight, M., Redman, C. W., Linton, E. A. & Sargent, I. L. (1998). Shedding of syncytiotrophoblast microvilli into the maternal circulation in pre-eclamptic pregnancies. Br. J. Obstet. Gynaecol., 105, 632–40.CrossRefGoogle ScholarPubMed
Kudo, Y. & Boyd, C. A. (1990). Characterization of amino acid transport systems in human placental basal membrane vesicles. Biochim. Biophys. Acta, 1021, 169–74.CrossRefGoogle ScholarPubMed
Kudo, Y., Boyd, C. A., Sargent, I. L. & Redman, C. W. (2003). Hypoxia alters expression and function of syncytin and its receptor during trophoblast cell fusion of human placental BeWo cells: implications for impaired trophoblast syncytialization in pre-eclampsia. Biochim. Biophys. Acta, 1638, 63–71.CrossRefGoogle Scholar
Kumar, S., Lo, D. Y. M., Smarason, A. K.et al. (2000). Pre-eclampsia is associated with increased levels of circulating apoptotic microparticles and fetal cell-free DNA. J. Soc. Gynecol. Invest., 7(Suppl), 181a.Google Scholar
Lammert, A., Kiess, W., Bottner, A., Glasow, A. & Kratzsch, J. (2001). Soluble leptin receptor represents the main leptin binding activity in human blood. Biochem. Biophys. Res. Commun., 283, 982–8.CrossRefGoogle ScholarPubMed
Lavillette, D., Marin, M., Ruggieri, A.et al. (2002). The envelope glycoprotein of human endogenous retrovirus type W uses a divergent family of amino acid transporters/cell surface receptors. J. Virol., 76, 6442–52.CrossRefGoogle ScholarPubMed
Lee, X., Keith, J.-C. J., Stumm, N.et al. (2001). Downregulation of placental syncytin expression and abnormal protein localization in preeclampsia. Placenta, 22, 808–12.CrossRefGoogle Scholar
Leu, S. J. & Singh, V. K. (1991). Modulation of natural killer cell-mediated lysis by corticotropin-releasing neurohormone. J. Neuroimmunol., 33, 253–60.CrossRefGoogle ScholarPubMed
Leung, D. N., Smith, S. C., To, K. F., Sahota, D. S. & Baker, P. N. (2001). Increased placental apoptosis in pregnancies complicated by preeclampsia. Am. J. Obstet. Gynecol., 184, 1249–50.CrossRefGoogle ScholarPubMed
Levy, R., Smith, S. D., Chandler, K., Sadovsky, Y. & Nelson, D. M. (2000). Apoptosis in human cultured trophoblasts is enhanced by hypoxia and diminished by epidermal growth factor. Am. J. Physiol. Cell. Physiol., 278, C982–8.CrossRefGoogle ScholarPubMed
Li, H., Dakour, J., Kaufman, S.et al. (2003). Adrenomedullin is decreased in preeclampsia because of failed response to epidermal growth factor and impaired syncytialization. Hypertension, 42, 895–900.CrossRefGoogle ScholarPubMed
Linton, E. A., Behan, D. P., Saphier, P. W. & Lowry, P. J. (1990). Corticotropin-releasing hormone binding protein: reduction in the ACTH releasing activity of placental but not hypothalamic CRF. J. Clin. Endocrinol. Metab., 70, 1574–80.CrossRefGoogle Scholar
Linton, E. A., Perkins, A. V., Woods, R. J.et al. (1993). Corticotropin-releasing hormone-binding protein (CRH-BP) plasma levels decrease during the third trimester of normal human pregnancy. J. Clin. Endocrinol. Metab., 76, 260–2.Google ScholarPubMed
Little, W. A. (1960). Placental infarction. Obstet. Gynecol., 15, 109–30.Google ScholarPubMed
Lo, Y. M., Leung, T. N., Tein, M. S.et al. (1999). Quantitative abnormalities of fetal DNA in maternal serum in preeclampsia. Clin. Chem., 45, 184–8.Google ScholarPubMed
Masuzaki, H., Ogawa, Y., Sagawa, N.et al. (1997). Nonadipose tissue production of leptin: leptin as a novel placenta-derived hormone in humans. Nat. Med., 3, 1029–33.CrossRefGoogle ScholarPubMed
Matsubara, S., Minakami, H., Sato, I. & Saito, T. (1997). Decrease in cytochrome c oxidase activity detected cytochemically in the placental trophoblast of patients with preeclampsia. Placenta, 18, 255–9.CrossRefGoogle Scholar
Mayhew, T. M., Leach, L., McGee, R.et al. (1999). Proliferation, differentiation and apoptosis in villous trophoblast at 13–41 weeks of gestation (including observations on annulate lamellae and nuclear pore complexes). Placenta, 20, 407–22.CrossRefGoogle Scholar
Mayne, S. T. (2003). Antioxidant nutrients and chronic disease: use of biomarkers of exposure and oxidative stress status in epidemiologic research. J. Nutr., 133, (Suppl 3), 933S–40S.CrossRefGoogle ScholarPubMed
Maynard, S. E, Min, J. Y., Merchan, J.et al. (2003). Excess placental soluble fms-like tyrosine kinase 1 (sFlt1) may contribute to endothelial dysfunction, hypertension, and proteinuria in preeclampsia. J. Clin. Invest., 111, 649–58.CrossRefGoogle ScholarPubMed
McCarthy, S. A. & Bicknell, R. (1993). Inhibition of vascular endothelial cell growth by activin-A. J. Biol. Chem., 268, 23066–71.Google ScholarPubMed
McCarthy, J. F., Misra, D. N. & Roberts, J. M. (1999). Maternal plasma leptin is increased in preeclampsia and positively correlates with fetal cord concentration. Am. J. Obstet. Gynecol., 180, 731–6.CrossRefGoogle ScholarPubMed
Mi, S., Lee, X., Li, X.et al. (2000). Syncytin is a captive retroviral envelope protein involved in human placental morphogenesis. Nature, 403, 785–9.CrossRefGoogle ScholarPubMed
Michel, U., Ebert, S., Phillips, D. & Nau, R. (2003). Serum concentrations of activin and follistatin are elevated and run in parallel in patients with septicemia. Eur. J. Endocrinol. 148, 559–64.CrossRefGoogle ScholarPubMed
Mise, H., Sagawa, N., Matsumoto, T.et al. (1998). Augmented placental production of leptin in preeclampsia: possible involvement of placental hypoxia. J. Clin. Endocrinol. Metab., 83, 3225–9.Google ScholarPubMed
Mohan, A., Asselin, J., Sargent, I. L., Groome, N. P. & Muttukrishna, S. (2001). Effect of cytokines and growth factors on the secretion of inhibin A, activin A and follistatin by term placental villous trophoblasts in culture. Eur. J. Endocrinol., 145, 505–11.CrossRefGoogle ScholarPubMed
Morikawa, S., Kurauchi, O., Tanaka, M.et al. (1997). Increased mitochondrial damage by lipid peroxidation in trophoblast cells of preeclamptic placentae. Biochem. Mol. Biol. Int., 41, 767–75.Google Scholar
Muttukrishna, S., Knight, P. G., Groome, N. P., Redman, C. W. & Ledger, W. L. (1997). Activin A and inhibin A as possible endocrine markers for preeclampsia. Lancet, 349, 1285–8.CrossRefGoogle Scholar
Muttukrishna, S., North, R. A., Morris, J.et al. (2000). Serum inhibin A and activin A are elevated prior to the onset of preeclampsia. Hum. Reprod., 15, 1640–5.CrossRefGoogle Scholar
Myatt, L., Rosenfield, R. B., Eis, A. L.et al. (1996). Nitrotyrosine residues in placenta. Evidence of peroxynitrite formation and action. Hypertension, 28, 488–93.CrossRefGoogle ScholarPubMed
Myatt, L., Eis, A. L., Brockman, D. E.et al. (1997). Differential localization of superoxide dismutase isoforms in placental villous tissue of normotensive, pre-eclamptic, and intrauterine growth-restricted pregnancies. J. Histochem. Cytochem., 45, 1433–8.CrossRefGoogle ScholarPubMed
Naicker, T., Khedun, S. M., Moodley, J. & Pijnenborg, R. (2003). Quantitative analysis of trophoblast invasion in preeclampsia. Acta Obstet. Gynecol. Scand., 82, 722–9.CrossRefGoogle ScholarPubMed
Navarra, P., Miceli, F., Tringali, G.et al. (2001). Evidence for a functional link between the heme oxygenase-carbon monoxide pathway and corticotropin-releasing hormone release from primary cultures of human trophoblast cells. J. Clin. Endocrinol. Metab., 86, 317–23.Google ScholarPubMed
Nelson, D. M. (1996). Apoptotic changes occur in syncytiotrophoblast of human placental villi where fibrin type fibrinoid is deposited at discontinuities in the villous trophoblast. Placenta, 17, 387–91.CrossRefGoogle ScholarPubMed
Ness, R. B. & Roberts, J. M. (1996). Heterogeneous causes constituting the single syndrome of preeclampsia, a hypothesis and its implications. Am. J. Obstet. Gynecol., 175, 1365–70.CrossRefGoogle ScholarPubMed
Ng, E. K., Leung, T. N., Tsui, N. B.et al. (2003a). The concentration of circulating corticotropin-releasing hormone mRNA in maternal plasma is increased in preeclampsia. Clin. Chem., 49, 727–31.CrossRefGoogle Scholar
Ng, E. K. O., Tsui, N. B. Y., Lau, T. K.et al. (2003b). mRNA of placental origin is readily detectable in maternal plasma. Proc. Natl. Acad. Sci. U.S.A., 100, 4748–53.CrossRefGoogle Scholar
Norman, M., Eriksson, C. G. & Eneroth, P. (1989). A comparison between the composition of maternal peripheral plasma and plasma collected from the retroplacental compartment at Caesarean section. A study on protein and steroid hormones and binding proteins. Arch. Gynecol. Obstet., 244, 215–26.CrossRefGoogle ScholarPubMed
Nuamah, M. A., Sagawa, N., Yura, S.et al. (2003). Free-to-total leptin ratio in maternal plasma is constant throughout human pregnancy. Endocr. J., 50, 421–8.CrossRefGoogle ScholarPubMed
Ohguchi, M., Yamato, K., Ishihara, Y.et al. (1998). Activin A regulates the production of mature interleukin-1beta and interleukin-1 receptor antagonist in human monocytic cells. J. Interferon Cytokine Res., 18, 491–8.CrossRefGoogle ScholarPubMed
Pereda, Paez M., Sauer, J., Perez-Castro, C.et al. (1995). Corticotropin-releasing hormone differentially modulates the interleukin-1 system according to the level of monocyte activation by endotoxin. Endocrinology, 136, 5504–10.CrossRefGoogle Scholar
Page, N. M., Woods, R. J., Gardiner, S. M.et al. (2000). Excessive placental secretion of neurokinin B during the third trimester causes preeclampsia. Nature, 405, 797–800.CrossRefGoogle Scholar
Park, H. Y., Kwon, H. M., Lim, H. J.et al. (2001). Potential role of leptin in angiogenesis: leptin induces endothelial cell proliferation and expression of matrix metalloproteinases in vivo and in vitro. Exp. Mol. Med., 33, 95–102.CrossRefGoogle ScholarPubMed
Patel, R. P., McAndrew, J., Sellak, H.et al. (1999). Biological aspects of reactive nitrogen species. Biochim. Biophys. Acta, 1411, 385–400.CrossRefGoogle ScholarPubMed
Perkins, A. V. & Linton, E. A. (1995). Identification and isolation of corticotrophin-releasing hormone-positive cells from the human placenta. Placenta, 16, 233–43.CrossRefGoogle ScholarPubMed
Perkins, A. V., Linton, E. A., Eben, F.et al. (1995). Corticotrophin-releasing hormone and corticotrophin-releasing hormone binding protein in normal and pre-eclamptic human pregnancies. Br. J. Obstet. Gynaecol., 102, 118–22.CrossRefGoogle ScholarPubMed
Piering, W. F., Garancis, J. G., Becker, C. G., Beres, J. A. & Lemann, J. (1993). Preeclampsia related to a functioning extrauterine placenta, report of a case and 25-year follow-up. Am. J. Kidney Dis., 21, 310–13.CrossRefGoogle ScholarPubMed
Pijnenborg, R., Anthony, J., Davey, D. A.et al. (1991). Placental bed spiral arteries in the hypertensive disorders of pregnancy. Br. J. Obstet. Gynaecol., 98, 648–55.CrossRefGoogle ScholarPubMed
Pijnenborg, R., Luyten, C., Vercruysse, L. & Assche, F. A. (1996). Attachment and differentiation in vitro of trophoblast from normal and preeclamptic human placentae. Am. J. Obstet. Gynecol., 175, 30–6.CrossRefGoogle Scholar
Poranen, A. K., Ekblad, U., Uotila, P. & Ahotupa, M. (1996). Lipid peroxidation and antioxidants in normal and pre-eclamptic pregnancies. Placenta, 17, 401–5.CrossRefGoogle ScholarPubMed
Poston, L. (2003). The role of oxidative stress. In Walker, J. J.Poston, L., eds., Preeclampsia, Proceedings of RCOG Study Group. London: RCOG Press, pp. 134–46.Google Scholar
Redman, C. W. G. (1991). Current topic. Preeclampsia and the placenta. Placenta, 12, 301–8.CrossRefGoogle ScholarPubMed
Redman, C. W. G. (1992). The placenta and preeclampsia. In , C. W. G. Redman, , I. L. Sargent &, , P. M. Starkey, eds., The Human Placenta. Oxford: Blackwell Scientific Publications, pp. 433–67.Google Scholar
Redman, C. W. G.Sargent, I. L. (2000). Placental debris, oxidative stress and preeclampsia. Placenta, 21, 597–602.CrossRefGoogle Scholar
Redman, C. W. G., Sacks, G. P. & Sargent, I. L. (1999). Preeclampsia, an excessive maternal inflammatory response to pregnancy. Am. J. Obstet. Gynecol., 180, 499–506.CrossRefGoogle Scholar
Reimer, T., Koczan, D., Gerber, B.et al. (2002). Microarray analysis of differentially expressed genes in placental tissue of preeclampsia: upregulation of obesity-related genes. Mol. Hum. Reprod., 8, 674–80.CrossRefGoogle ScholarPubMed
Roberts, J. M., Taylor, R. N., Musci, T. J.et al. (1989). Preeclampsia, an endothelial cell disorder. Am. J. Obstet. Gynecol., 161, 1200–4.CrossRefGoogle ScholarPubMed
Robertson, W. B., Brosens, I. & Dixon, G. (1976). Maternal uterine vascular lesions in the hypertensive complications of pregnancy. Perspect. Nephrol. Hypertens., 5, 115–27.Google ScholarPubMed
Robertson, W. B., Brosens, I. & Landells, W. N. (1985). Abnormal placentation. Obstet. Gynecol. Annu., 14, 411–26.Google ScholarPubMed
Robinson, B. G., Arbiser, J. L., Emanuel, R. L. & Majzoub, J. A. (1989). Species-specific placental corticotropin releasing hormone messenger RNA and peptide expression. Mol. Cell. Endocrinol., 62, 337–41.CrossRefGoogle ScholarPubMed
Ros, H. S., Cnattingius, S. & Lipworth, L. (1998). Comparison of risk factors for preeclampsia and gestational hypertension in a population-based cohort study. Am. J. Epidemiol., 147, 1062–70.CrossRefGoogle Scholar
Russell, C. E., Hedger, M. P., Brauman, J. N., Krester, D. M. & Phillips, D. J. (1999). Activin A regulates growth and acute phase proteins in the human lives cell line, HepQ2. Mol. Cell. Endocrinol. 148, 129–36.CrossRefGoogle Scholar
Sahlin, L., Ostlund, E., Wang, H., Homgren, A. & Fried, G. (2000). Decreased expression of thioredoxin and glutaredoxin in placentae from pregnancies with preeclampsia and intrauterine growth restriction. Placenta, 21, 603–9.CrossRefGoogle ScholarPubMed
Sakuragi, N., Matsuo, H., Coukos, G.et al. (1994). Differentiation-dependent expression of the BCL-2 proto-oncogene in the human trophoblast lineage. J. Soc. Gynecol. Invest., 1, 164–72.CrossRefGoogle ScholarPubMed
Santoso, D. I., Rogers, P., Wallace, E. M.et al. (2002). Localization of indoleamine 2,3-dioxygenase and 4-hydroxynonenal in normal and pre-eclamptic placentae. Placenta, 23, 373–9.CrossRefGoogle ScholarPubMed
Sasaki, A., Shinkawa, O., Margioris, A. N.et al. (1987). Immunoreactive corticotropin-releasing hormone in human plasma during pregnancy, labor, and delivery. J. Clin. Endocrinol. Metab., 64, 224–9.CrossRefGoogle Scholar
Sattar, N., Greer, I. A., Pirwani, I., Gibson, J. & Wallace, A. M. (1998). Leptin levels in pregnancy: marker for fat accumulation and mobilization?Acta. Obstet. Gynecol. Scand., 77, 278–83.CrossRefGoogle ScholarPubMed
Schneider-Kolsky, M. E., Manuelpillai, U., Waldron, K., Dole, A. & Wallace, E. M. (2002). The distribution of activin and activin receptors in gestational tissues across human pregnancy and during labour. Placenta, 23, 294–302.CrossRefGoogle ScholarPubMed
Schrocksnadel, H., Daxenbichler, G., Artner, E., Steckel-Berger, G. & Dapunt, O. (1993). Tumor markers in hypertensive disorders of pregnancy. Gynecol. Obstet. Invest., 35, 204–8.CrossRefGoogle ScholarPubMed
Senaris, R., Garcia, C. T., Casabiell, X.et al. (1997). Synthesis of leptin in human placenta. Endocrinology, 138, 4501–4.CrossRefGoogle ScholarPubMed
Serdar, Z., Gur, E., Colakoethullary, M., Develioethlu, O. & Sarandol, E. (2003). Lipid and protein oxidation and antioxidant function in women with mild and severe preeclampsia. Arch. Gynecol. Obstet., 268, 19–25.Google ScholarPubMed
Sheppard, B. L. & Bonnar, J. (1976). The ultrastructure of the arterial supply of the human placenta in pregnancy complicated by fetal growth retardation. Br. J. Obstet. Gynaecol., 83, 948–59.CrossRefGoogle ScholarPubMed
Shibata, E., Ejima, K., Nanri, H.et al. (2001). Enhanced protein levels of protein thiol/disulphide oxidoreductases in placentae from pre-eclamptic subjects. Placenta, 22, 566–72.CrossRefGoogle ScholarPubMed
Shibata, E., Nanri, H., Ejima, K.et al. (2003). Enhancement of mitochondrial oxidative stress and upregulation of antioxidant protein peroxiredoxin III/SP-22 in the mitochondria of human pre-eclamptic placentae. Placenta, 24, 698–705.CrossRefGoogle ScholarPubMed
Sibai, B. M., Gordon, T., Thom, E.et al. (1995). Risk factors for preeclampsia in healthy nulliparous women: a prospective multicenter study. The National Institute of Child Health and Human Development Network of Maternal-Fetal Medicine Units. Am. J. Obstet. Gynecol., 172, 642–8.CrossRefGoogle ScholarPubMed
Sierra-Honigmann, M. R., Nath, A. K.Mwakami, C.et al. (1998). Biological action of leptin as an angiogenic factor. Science, 281, 1683–6.CrossRefGoogle ScholarPubMed
Sikkema, J. M., Rijn, B. B., Franx, A.et al. (2001). Placental superoxide is increased in preeclampsia. Placenta, 22, 304–8.CrossRefGoogle Scholar
Smarason, A. K., Sargent, I. L., Starkey, P. M. & Redman, C. W. G. (1993). The effect of placental syncytiotrophoblast microvillous membranes from normal and pre-eclamptic women on the growth of endothelial cells in vitro. Br. J. Obstet. Gynaecol., 100, 943–9.CrossRefGoogle ScholarPubMed
Smith, G. C., Pell, J. P. & Walsh, D. (2001). Pregnancy complications and maternal risk of ischaemic heart disease: a retrospective cohort study of 129,290 births. Lancet, 357, 2002–6.CrossRefGoogle ScholarPubMed
Staff, A. C., Ranheim, T., Khoury, J. & Henriksen, T. (1999). Increased contents of phospholipids, cholesterol, and lipid peroxides in decidua basalis in women with preeclampsia. Am. J. Obstet. Gynecol., 180, 587–92.CrossRefGoogle ScholarPubMed
Takacs, P., Kauma, S. W., Sholley, M. M.et al. (2001). Increased circulating lipid peroxides in severe preeclampsia activate NF-KappaB and upregulate ICAM-1 in vascular endothelial cells. FASEB J., 15, 279–81.CrossRefGoogle ScholarPubMed
Tannetta, D. S., Muttukrishna, S., Groome, N. P., Redman, C. W. & Sargent, I. L. (2003). Endothelial cells and peripheral blood mononuclear cells are a potential source of extraplacental activin a in preeclampsia. J. Clin. Endocrinol. Metab., 88, 5995–6001.CrossRefGoogle ScholarPubMed
Tsatsaris, V., Goffin, F., Munaut, C.et al. (2003). Over-expression of the soluble vascular endothelial growth factor receptor in preeclamptic patients: pathophysiological consequences. J. Clin. Endocrinol. Metab., 88, 5555–63.CrossRefGoogle Scholar
Dadelszen, P., Hurst, G. & Redman, C. W. G. (1999). The supernatants from co-cultured endothelial cells and syncytiotrophoblast microvillous membranes activate peripheral blood leucocytes in vitro. Hum. Reprod., 14, 919–24.CrossRefGoogle Scholar
Vuorinen, K., Remes, A., Sormunen, R., Tapanainen, J. & Hassinen, I. E. (1998). Placental mitochondrial DNA and respiratory chain enzymes in the etiology of preeclampsia. Obstet. Gynecol., 91, 950–955.Google ScholarPubMed
Wallenburg, H. C., Stolte, L. A. M. & Janssens, J. (1973). The pathogenesis of placental infarction I. A morphologic study in the human placenta. Am. J. Obstet. Gynecol., 116, 835–40.CrossRefGoogle ScholarPubMed
Walsh, S. W. & Wang, Y. (1993). Deficient glutathione peroxidase activity in preeclampsia is associated with increased placental production of thromboxane and lipid peroxides. Am. J. Obstet. Gynecol., 169, 1456–61.CrossRefGoogle ScholarPubMed
Wang, Y. & Walsh, S. W. (1996). Antioxidant activities and mRNA expression of superoxide dismutase, catalase, and glutathione peroxidase in normal and preeclamptic placentae. J. Soc. Gynecol. Inves., 3, 179–84.CrossRefGoogle Scholar
Wang, Y. & Walsh, S. W. (1998). Placental mitochondria as a source of oxidative stress in preeclampsia. Placenta, 19, 581–6.CrossRefGoogle Scholar
Wang, Y. & Walsh, S. W. (2001). Increased superoxide generation is associated with decreased superoxide dismutase activity and mRNA expression in placental trophoblast cells in preeclampsia. Placenta, 22, 206–12.CrossRefGoogle Scholar
Wang, Z., Zhang, G. & Lin, M. (1999). Mitochondrial tRNA(leu)(UUR) gene mutation and the decreased activity of cytochrome c oxidase in preeclampsia. J. Tongji Med. Univ., 19, 209–11.Google ScholarPubMed
Weinstein, L. (1982). Syndrome of hemolysis, elevated liver enzymes, and low platelet count: a severe consequence of hypertension in pregnancy. Am. J. Obstet. Gynecol., 142, 159–67.CrossRefGoogle ScholarPubMed
Wentworth, P. (1967). Placental infarction and toxemia of pregnancy. Am. J. Obstet. Gynecol., 99, 318–26.CrossRefGoogle ScholarPubMed
Yang, J. C., Haworth, L., Sherry, R. M.et al. (2003). A randomized trial of bevacizumab, an anti-vascular endothelial growth factor antibody, for metastatic renal cancer. New Engl. J. Med., 349, 427–34.CrossRefGoogle ScholarPubMed
Yusuf, K., Smith, S. D., Sadovsky, Y. & Nelson, D. M. (2002). Trophoblast differentiation modulates the activity of caspases in primary cultures of term human trophoblasts. Pediatr. Res., 52, 411–15.CrossRefGoogle ScholarPubMed
Zabriskie, J. R. (1963). Effect of cigarette smoking during pregnancy. Study of 2000 cases. Obstet. Gynecol., 21, 405–11.Google Scholar
Zarkesh-Esfahani, H., Pockley, G., Metcalfe, R. A.et al. (2001). High-dose leptin activates human leucocytes via receptor expression on monocytes. J. Immunol., 167, 4593–9.CrossRefGoogle ScholarPubMed
Zeek, P. M. & Assali, N. S. (1950). Vascular changes with eclamptogenic toxemia of pregnancy. Am. J. Clin. Pathol., 20, 1099–109.CrossRefGoogle ScholarPubMed
Zhong, X. Y., Laivuori, H., Livingston, J. C.et al. (2001). Elevation of both maternal and fetal extracellular circulating deoxyribonucleic acid concentrations in the plasma of pregnant women with preeclampsia. Am. J. Obstet. Gynecol., 184, 414–19.CrossRefGoogle ScholarPubMed
Zusterzeel, P. L., Rutten, H., Roelofs, H. M., Peters, W. H. & Steegers, E. A. (2001). Protein carbonyls in decidua and placenta of pre-eclamptic women as markers for oxidative stress. Placenta, 22, 213–19.CrossRefGoogle ScholarPubMed
Hafner, E., Metzenbauer, M., Hofinger, D.et al. (2003). Placental growth from the first to the second trimester of pregnancy in SGA-foetuses and pre-eclamptic pregnancies compared to normal foetuses. Placenta, 24, 336–42.CrossRefGoogle ScholarPubMed
Hofmeyr, G. J., Wilkins, T. & Redman, C. W. (1991). C4 and plasma protein in hypertension during pregnancy with and without proteinuria. BMJ, 302, 218.CrossRefGoogle ScholarPubMed

Save book to Kindle

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

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

Find out more about the Kindle Personal Document Service.

  • Trophoblast and pre-eclampsia
    • By C. W. G. Redman, Nuffield Department of Obstetrics and Gynaecology, University of Oxford, UK, I. L. Sargent, Nuffield Department of Obstetrics and Gynaecology, University of Oxford, UK, E .A. Linton, Nuffield Department of Obstetrics and Gynaecology, University of Oxford, UK
  • Edited by Ashley Moffett, University of Cambridge, Charlie Loke, University of Cambridge, Anne McLaren, Cancer Research, UK
  • Book: Biology and Pathology of Trophoblast
  • Online publication: 07 August 2009
  • Chapter DOI: https://doi.org/10.1017/CBO9780511545207.015
Available formats
×

Save book to Dropbox

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

  • Trophoblast and pre-eclampsia
    • By C. W. G. Redman, Nuffield Department of Obstetrics and Gynaecology, University of Oxford, UK, I. L. Sargent, Nuffield Department of Obstetrics and Gynaecology, University of Oxford, UK, E .A. Linton, Nuffield Department of Obstetrics and Gynaecology, University of Oxford, UK
  • Edited by Ashley Moffett, University of Cambridge, Charlie Loke, University of Cambridge, Anne McLaren, Cancer Research, UK
  • Book: Biology and Pathology of Trophoblast
  • Online publication: 07 August 2009
  • Chapter DOI: https://doi.org/10.1017/CBO9780511545207.015
Available formats
×

Save book to Google Drive

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

  • Trophoblast and pre-eclampsia
    • By C. W. G. Redman, Nuffield Department of Obstetrics and Gynaecology, University of Oxford, UK, I. L. Sargent, Nuffield Department of Obstetrics and Gynaecology, University of Oxford, UK, E .A. Linton, Nuffield Department of Obstetrics and Gynaecology, University of Oxford, UK
  • Edited by Ashley Moffett, University of Cambridge, Charlie Loke, University of Cambridge, Anne McLaren, Cancer Research, UK
  • Book: Biology and Pathology of Trophoblast
  • Online publication: 07 August 2009
  • Chapter DOI: https://doi.org/10.1017/CBO9780511545207.015
Available formats
×