Jump to content

Paternal age effect: Difference between revisions

From Wikipedia, the free encyclopedia
Content deleted Content added
revert Ozzie10aaaa edits - you should not remove references for being dated while leaving the information unreferenced
Line 46: Line 46:


===Monogenetic disorders===
===Monogenetic disorders===
Studies reveal that the following list of congenital disorders, collectively known as paternal age effect (PAE) disorders, are all caused by a small number of dominantly-acting point mutations and almost exclusively originate from unaffected fathers, suggesting that the mutations are taking place during spermatogenesis. Mutations in the fibroblast growth factor receptor genes FGFR2, cause [[Apert syndrome]],<ref name="Tarín-1998"/><ref name="Risch-1987"/><ref name="Yoon-2009">{{Cite journal | author=Yoon SR, Qin J, Glaser RL, Jabs EW, Wexler NS, Sokol R, Arnheim N, Calabrese P | title = The Ups and Downs of Mutation Frequencies during Aging Can Account for the Apert Syndrome Paternal Age Effect | journal = PLoS Genet | volume = 5 | issue = 7 | pages = e1000558 | year = 2009 | url = http://www.plosgenetics.org/article/info%3Adoi%2F10.1371%2Fjournal.pgen.1000558/ | doi = 10.1371/journal.pgen.1000558 | pmid = 19593369 | pmc=2700275 | editor1-last=Walsh | editor1-first=Bruce }}</ref> [[Crouzon syndrome]],<ref name="Tarín-1998"/><ref name="Risch-1987"/> and [[Pfeiffer syndrome]].<ref name="Tarín-1998"/><ref name="Risch-1987"/><ref name="Goriely and Wilkie 2012">{{cite journal | author = Goriely A, Wilkie AOM | title = Paternal Age Effect Mutations and Selfish Spermatogonial Selection: Causes and Consequences For Human Disease | journal = Am. J. Human Genet. | volume = 90 | issue = 2 | pages = 175–200 | year = 2012 | doi=10.1016/j.ajhg.2011.12.017 | pmid=22325359}}</ref> Mutations in the FGFR3 gene lead to the formation of [[achondroplasia]],<ref name="Tarín-1998">{{Cite journal | author = Tarín JJ, Brines J, Cano A | title = Long-term effects of delayed parenthood | journal = Hum Reprod | volume = 13 | issue = 9 | pages = 2371–6 | year = 1998 | doi = 10.1093/humrep/13.9.2371 | url = http://humrep.oxfordjournals.org/cgi/reprint/13/9/2371 | pmid = 9806250 }}</ref><ref name="Risch-1987">{{Cite journal | author = Risch N, Reich EW, Wishnick MM, McCarthy JG | title = Spontaneous mutation and parental age in humans | journal = American Journal of Human Genetics | volume = 41 | issue = 2 | pages = 218–48 | year = 1987 | doi = | pmid = 3618593 | pmc = 1684215 }}</ref> [[thanatophoric dysplasia]],<ref name="Tarín-1998"/> [[hypochondroplasia]], and [[Muenke syndrome]].<ref name="Goriely and Wilkie 2012" /> These disorders occur spontaneously as a result of advanced paternal age and at the rate of 1 in 30,000 for achondroplasia births.<ref name="Orioli et al 1986">{{cite journal | author = Orioli I.M. et al | title = The birth prevalence rates for the skeletal dysplasias | journal = J. Med. Genet. | volume = 23 | pages = 328–332 | year = 1986 | doi=10.1136/jmg.23.4.328}}</ref><ref name="Waller DK 2008">{{cite journal | author = Waller D.K. | title = The population-based prevalence of achondroplasia and thanatophoric dysplasia in selected regions of the U.S. | journal = Am. J. Med. Genet. | volume = 146A | pages = 2385–9 | year = 2008 | doi=10.1002/ajmg.a.32485}}</ref> Other conditions involving the mutations in the RET gene lead to [[multiple endocrine neoplasia type 2A]] and 2B, the PTPN11 gene which leads to [[Noonan syndrome]],<ref name="Tartaglia et al 2004">{{cite journal | author = Tartaglia M. et al | title = Paternal germline origin and sex-ratio distortion in transmission of PTPN11 mutations in Noonan syndrome | journal = Am. J. Hum. Genet. | volume = 75 | pages = 492–7 | year = 2004 | doi=10.1086/423493}}</ref> and the HRAS mutations which cause [[Costello syndrome]].<ref name="Tarín-1998"/>{{Cite journal | author = Carlson KM, Bracamontes J, Jackson CE, Clark R, Lacroix A, Wells SA Jr, Goodfellow PJ | title = Parent-of-origin effects in multiple endocrine neoplasia type 2B | journal = American Journal of Human Genetics | volume = 55 | issue = 6 | pages = 1076–82 | year = 1994 | pmid = 7977365 | pmc = 1918453 }}</ref><ref name="Schuffenecker-1997">{{Cite journal | author = Schuffenecker I, Ginet N, Goldgar D, Eng C, Chambe B, Boneu A, Houdent C, Pallo D, Schlumberger M, Thivolet C, Lenoir GM; Le Groupe d'Etude des Tumeurs a Calcitonine | title = Prevalence and parental origin of de novo RET mutations in multiple endocrine neoplasia type 2A and familial medullary thyroid carcinoma. Le Groupe d'Etude des Tumeurs a Calcitonine | journal = American Journal of Human Genetics | volume = 60 | issue = 1 | pages = 233–7 | year = 1997 | pmid = 8981969 | pmc = 1712555 }}</ref><ref name="Sol-Church et al 2006">{{cite journal | author = Sol-Church K. et al | title = Paternal bias in parental origin of HRAS mutations in Costello syndrome | journal = Hum. Mutat. | volume = 27 | pages = 736–741 | year = 2006 | doi=10.1002/humu.20381}}</ref>
Studies reveal that the following list of congenital disorders, collectively known as paternal age effect (PAE) disorders, are all caused by a small number of dominantly-acting point mutations and almost exclusively originate from unaffected fathers, suggesting that the mutations are taking place during spermatogenesis. Mutations in the fibroblast growth factor receptor genes FGFR2, cause [[Apert syndrome]],<ref name="Tarín-1998"/><ref name="Risch-1987"/><ref name="Yoon-2009">{{Cite journal | author=Yoon SR, Qin J, Glaser RL, Jabs EW, Wexler NS, Sokol R, Arnheim N, Calabrese P | title = The Ups and Downs of Mutation Frequencies during Aging Can Account for the Apert Syndrome Paternal Age Effect | journal = PLoS Genet | volume = 5 | issue = 7 | pages = e1000558 | year = 2009 | url = http://www.plosgenetics.org/article/info%3Adoi%2F10.1371%2Fjournal.pgen.1000558/ | doi = 10.1371/journal.pgen.1000558 | pmid = 19593369 | pmc=2700275 | editor1-last=Walsh | editor1-first=Bruce }}</ref> [[Crouzon syndrome]],<ref name="Tarín-1998"/><ref name="Risch-1987"/> and [[Pfeiffer syndrome]].<ref name="Tarín-1998"/><ref name="Risch-1987"/><ref name="Goriely and Wilkie 2012">{{cite journal | author = Goriely A, Wilkie AOM | title = Paternal Age Effect Mutations and Selfish Spermatogonial Selection: Causes and Consequences For Human Disease | journal = Am. J. Human Genet. | volume = 90 | issue = 2 | pages = 175–200 | year = 2012 | doi=10.1016/j.ajhg.2011.12.017 | pmid=22325359}}</ref> Mutations in the FGFR3 gene lead to the formation of [[achondroplasia]],<ref name="Tarín-1998">{{Cite journal | author = Tarín JJ, Brines J, Cano A | title = Long-term effects of delayed parenthood | journal = Hum Reprod | volume = 13 | issue = 9 | pages = 2371–6 | year = 1998 | doi = 10.1093/humrep/13.9.2371 | url = http://humrep.oxfordjournals.org/cgi/reprint/13/9/2371 | pmid = 9806250 }}</ref><ref name="Risch-1987">{{Cite journal | author = Risch N, Reich EW, Wishnick MM, McCarthy JG | title = Spontaneous mutation and parental age in humans | journal = American Journal of Human Genetics | volume = 41 | issue = 2 | pages = 218–48 | year = 1987 | doi = | pmid = 3618593 | pmc = 1684215 }}</ref> [[thanatophoric dysplasia]],<ref name="Tarín-1998"/> [[hypochondroplasia]], and [[Muenke syndrome]].<ref name="Goriely and Wilkie 2012" /> These disorders occur spontaneously as a result of advanced paternal age and at the rate of 1 in 30,000 for achondroplasia births.<ref name="Orioli et al 1986">{{cite journal | author = Orioli I.M. et al | title = The birth prevalence rates for the skeletal dysplasias | journal = J. Med. Genet. | volume = 23 | pages = 328–332 | year = 1986 | doi=10.1136/jmg.23.4.328}}</ref><ref name="Waller DK 2008">{{cite journal | author = Waller D.K. | title = The population-based prevalence of achondroplasia and thanatophoric dysplasia in selected regions of the U.S. | journal = Am. J. Med. Genet. | volume = 146A | pages = 2385–9 | year = 2008 | doi=10.1002/ajmg.a.32485}}</ref> Other conditions involving the mutations in the RET gene lead to [[multiple endocrine neoplasia type 2A]] and 2B, the PTPN11 gene which leads to [[Noonan syndrome]],<ref name="Tartaglia et al 2004">{{cite journal | author = Tartaglia M. et al | title = Paternal germline origin and sex-ratio distortion in transmission of PTPN11 mutations in Noonan syndrome | journal = Am. J. Hum. Genet. | volume = 75 | pages = 492–7 | year = 2004 | doi=10.1086/423493}}</ref> and the HRAS mutations which cause [[Costello syndrome]].<ref name="Tarín-1998"/>{{Cite journal | author = Carlson KM, Bracamontes J, Jackson CE, Clark R, Lacroix A, Wells SA Jr, Goodfellow PJ | title = Parent-of-origin effects in multiple endocrine neoplasia type 2B | journal = American Journal of Human Genetics | volume = 55 | issue = 6 | pages = 1076–82 | year = 1994 | pmid = 7977365 | pmc = 1918453 }}</ref><ref name="Schuffenecker-1997">{{Cite journal | author = Schuffenecker I, Ginet N, Goldgar D, Eng C, Chambe B, Boneu A, Houdent C, Pallo D, Schlumberger M, Thivolet C, Lenoir GM; Le Groupe d'Etude des Tumeurs a Calcitonine | title = Prevalence and parental origin of de novo RET mutations in multiple endocrine neoplasia type 2A and familial medullary thyroid carcinoma. Le Groupe d'Etude des Tumeurs a Calcitonine | journal = American Journal of Human Genetics | volume = 60 | issue = 1 | pages = 233–7 | year = 1997 | pmid = 8981969 | pmc = 1712555 }}</ref><ref name="Sol-Church et al 2006">{{cite journal | author = Sol-Church K. et al | title = Paternal bias in parental origin of HRAS mutations in Costello syndrome | journal = Hum. Mutat. | volume = 27 | pages = 736–741 | year = 2006 | doi=10.1002/humu.20381}}</ref>


===Other conditions===
===Other conditions===
Other conditions and diseases which have been suggested as having a possible correlation with paternal age include: [[chondrodystrophy]],<ref name="Lian-1986">{{Cite journal | author = Lian ZH, Zack MM, Erickson JD | title = Paternal age and the occurrence of birth defects | journal = American Journal of Human Genetics | volume = 39 | issue = 5 | pages = 648–60 | year = 1986 | pmid = 3788977 | pmc = 1684057 }}</ref> [[acrodysostosis]],<ref name="Tarín-1998"/> [[aniridia]],<ref name="Tarín-1998"/>[[Nevoid basal cell carcinoma syndrome|basal cell nevus syndrome]],<ref name="Tarín-1998"/> [[cataract]]s,<ref name="Tarín-1998"/> [[Cerebral palsy]], athetoid/dystonic,<ref name="Tarín-1998"/> [[CHARGE syndrome]],<ref name="Blake-2006">{{Cite journal | author = Blake KD, Prasad C | title = CHARGE syndrome | journal = Orphanet J Rare Dis | volume = 1 | issue = | pages = 34 | year = 2006 | doi = 10.1186/1750-1172-1-34 | url = http://www.biomedcentral.com/content/pdf/1750-1172-1-34.pdf | pmid = 16959034 | pmc = 1586184 }}</ref> [[cleft palate]],<ref name="Tarín-1998"/><ref name="Green-2010">{{Cite journal | author = Green RF, Devine O, Crider KS, Olney RS, Archer N, Olshan AF, Shapira SK | title = Association of Paternal Age and Risk for Major Congenital Anomalies from the National Birth Defects Prevention Study, 1997–2004 | journal = Annals of Epidemiology | volume = 20 | issue = 3 | pages = 241–9 | year = 2010 | doi = 10.1016/j.annepidem.2009.10.009 | url = http://www.annalsofepidemiology.org/article/S1047-2797%2809%2900367-6/abstract | pmid = 20056435 | pmc = 2824069 }}</ref> [[cleidocranial dysostosis]],<ref name="Tarín-1998"/> [[craniosynostosis]],<ref name="Rasmussen-2008">{{Cite journal | author = Rasmussen SA, Yazdy MM, Frías JL, Honein MA | title = Priorities for public health research on craniosynostosis: summary and recommendations from a Centers for Disease Control and Prevention-sponsored meeting | journal = American Journal of Medical Genetics | volume = 146A | issue = 2 | pages = 149–58 | year = 2008 | doi = 10.1002/ajmg.a.32106 | url = http://www.ameriface.org/Craniosynostosis_AJMG.pdf | pmid = 18080327 }}</ref>[[diaphragmatic hernia]],<ref name="Green-2010"/> [[Duchenne muscular dystrophy]],<ref name="Tarín-1998"/> [[exostosis|exostoses]], multiple,<ref name="Tarín-1998"/> congenital malformations in extremities,<ref name="Tarín-1998"/><ref name="Zhu-2005">{{Cite journal | author = Zhu JL, Madsen KM, Vestergaard M, Olesen AV, Basso O, Olsen J | title = Paternal age and congenital malformations | journal = Hum Reprod | volume = 20 | issue = 11 | pages = 3173–7 | year = 2005 | doi = 10.1093/humrep/dei186 | url=http://humrep.oxfordjournals.org/cgi/content/full/20/11/3173 | pmid = 16006461 }}</ref> [[fibrodysplasia ossificans progressiva]],<ref name="Tarín-1998"/><ref name="Risch-1987"/> Heart defects,<ref name="Tarín-1998"/><ref name="Lian-1986"/> [[Hemiplegia]],<ref name="Tarín-1998"/> [[Hemophilia A]],<ref name="Tarín-1998"/> [[Klinefelter's syndrome]],<ref>BRS Genetics - Dudek 2009</ref> [[Lesch-Nyhan syndrome]],<ref name="Tarín-1998"/> [[Marfan syndrome]],<ref name="Risch-1987"/> nasal [[aplasia]],<ref name="Tarín-1998"/>[[neural tube defect]]s,<ref name="Tarín-1998"/> [[oculodentodigital syndrome]],<ref name="Tarín-1998"/> [[osteogenesis imperfecta]] type IIA,<ref name="Tarín-1998"/>[[polycystic kidney disease]],<ref name="Tarín-1998"/> Polyposis coli,<ref name="Tarín-1998"/> [[Preauricular sinus and cyst|Preauricular cyst]],<ref name="Tarín-1998"/> [[Progeria]],<ref name="Tarín-1998"/> [[Psychosis|psychotic disorders]],<ref name="Tarín-1998"/><ref name="Miller-2010">{{Cite journal | author = Miller B, Pihlajamaa J, Haukka J, Cannon M, Henriksson M, Heilä H, Huttunen M, Tanskanen A, Lönnqvist J, Suvisaari J, Kirkpatrick B | title = Paternal age and mortality in nonaffective psychosis | journal = Schizophr Res | volume = 121 | issue = 1–3 | pages = 218–26 |date=16 Feb 2010| doi = 10.1016/j.schres.2010.01.020 | pmid = 20163936 }}</ref> [[Neurofibromatosis type I|von Recklinghausen neurofibromatosis]],<ref name="Tarín-1998"/><ref name="Risch-1987"/> [[retinitis pigmentosa]],<ref name="Tarín-1998"/>[[retinoblastoma]], bilateral,<ref name="Tarín-1998"/> [[situs inversus]],<ref name="Lian-1986"/> Soto's basal cell nevus,<ref name="Tarín-1998"/>[[Treacher-Collins Syndrome]],<ref name="Tarín-1998"/> [[tuberous sclerosis]],<ref name="Tarín-1998"/> [[Meatal stenosis|Urethral stenosis]],<ref name="Tarín-1998"/> [[Waardenburg syndrome]],<ref name="Tarín-1998"/> and [[Wilms' tumor]].<ref name="Tarín-1998"/>
Other conditions and diseases which have been suggested as having a possible correlation with paternal age include: [[chondrodystrophy]],<ref name="Lian-1986">{{Cite journal | author = Lian ZH, Zack MM, Erickson JD | title = Paternal age and the occurrence of birth defects | journal = American Journal of Human Genetics | volume = 39 | issue = 5 | pages = 648–60 | year = 1986 | pmid = 3788977 | pmc = 1684057 }}</ref> [[acrodysostosis]],<ref name="Tarín-1998"/> [[aniridia]],<ref name="Tarín-1998"/>[[Nevoid basal cell carcinoma syndrome|basal cell nevus syndrome]],<ref name="Tarín-1998"/> [[cataract]]s,<ref name="Tarín-1998"/> [[Cerebral palsy]], athetoid/dystonic,<ref name="Tarín-1998"/> [[CHARGE syndrome]],<ref name="Blake-2006">{{Cite journal | author = Blake KD, Prasad C | title = CHARGE syndrome | journal = Orphanet J Rare Dis | volume = 1 | issue = | pages = 34 | year = 2006 | doi = 10.1186/1750-1172-1-34 | url = http://www.biomedcentral.com/content/pdf/1750-1172-1-34.pdf | pmid = 16959034 | pmc = 1586184 }}</ref> [[cleft palate]],<ref name="Tarín-1998"/><ref name="Green-2010">{{Cite journal | author = Green RF, Devine O, Crider KS, Olney RS, Archer N, Olshan AF, Shapira SK | title = Association of Paternal Age and Risk for Major Congenital Anomalies from the National Birth Defects Prevention Study, 1997–2004 | journal = Annals of Epidemiology | volume = 20 | issue = 3 | pages = 241–9 | year = 2010 | doi = 10.1016/j.annepidem.2009.10.009 | url = http://www.annalsofepidemiology.org/article/S1047-2797%2809%2900367-6/abstract | pmid = 20056435 | pmc = 2824069 }}</ref> [[cleidocranial dysostosis]],<ref name="Tarín-1998"/> [[craniosynostosis]],<ref name="Rasmussen-2008">{{Cite journal | author = Rasmussen SA, Yazdy MM, Frías JL, Honein MA | title = Priorities for public health research on craniosynostosis: summary and recommendations from a Centers for Disease Control and Prevention-sponsored meeting | journal = American Journal of Medical Genetics | volume = 146A | issue = 2 | pages = 149–58 | year = 2008 | doi = 10.1002/ajmg.a.32106 | url = http://www.ameriface.org/Craniosynostosis_AJMG.pdf | pmid = 18080327 }}</ref>[[diaphragmatic hernia]],<ref name="Green-2010"/> [[Duchenne muscular dystrophy]],<ref name="Tarín-1998"/> [[exostosis|exostoses]], multiple,<ref name="Tarín-1998"/> congenital malformations in extremities,<ref name="Tarín-1998"/><ref name="Zhu-2005">{{Cite journal | author = Zhu JL, Madsen KM, Vestergaard M, Olesen AV, Basso O, Olsen J | title = Paternal age and congenital malformations | journal = Hum Reprod | volume = 20 | issue = 11 | pages = 3173–7 | year = 2005 | doi = 10.1093/humrep/dei186 | url=http://humrep.oxfordjournals.org/cgi/content/full/20/11/3173 | pmid = 16006461 }}</ref> [[fibrodysplasia ossificans progressiva]],<ref name="Tarín-1998"/><ref name="Risch-1987"/> Heart defects,<ref name="Tarín-1998"/><ref name="Lian-1986"/> [[Hemiplegia]],<ref name="Tarín-1998"/> [[Hemophilia A]],<ref name="Tarín-1998"/> [[Klinefelter's syndrome]],<ref>BRS Genetics - Dudek 2009</ref> [[Lesch-Nyhan syndrome]],<ref name="Tarín-1998"/> [[Marfan syndrome]],<ref name="Risch-1987"/> nasal [[aplasia]],<ref name="Tarín-1998"/>[[neural tube defect]]s,<ref name="Tarín-1998"/> [[oculodentodigital syndrome]],<ref name="Tarín-1998"/> [[osteogenesis imperfecta]] type IIA,<ref name="Tarín-1998"/>[[polycystic kidney disease]],<ref name="Tarín-1998"/> Polyposis coli,<ref name="Tarín-1998"/> [[Preauricular sinus and cyst|Preauricular cyst]],<ref name="Tarín-1998"/> [[Progeria]],<ref name="Tarín-1998"/> [[Psychosis|psychotic disorders]],<ref name="Tarín-1998"/><ref name="Miller-2010">{{Cite journal | author = Miller B, Pihlajamaa J, Haukka J, Cannon M, Henriksson M, Heilä H, Huttunen M, Tanskanen A, Lönnqvist J, Suvisaari J, Kirkpatrick B | title = Paternal age and mortality in nonaffective psychosis | journal = Schizophr Res | volume = 121 | issue = 1–3 | pages = 218–26 |date=16 Feb 2010| doi = 10.1016/j.schres.2010.01.020 | pmid = 20163936 }}</ref> [[Neurofibromatosis type I|von Recklinghausen neurofibromatosis]],<ref name="Tarín-1998"/><ref name="Risch-1987"/> [[retinitis pigmentosa]],<ref name="Tarín-1998"/>[[retinoblastoma]], bilateral,<ref name="Tarín-1998"/> [[situs inversus]],<ref name="Lian-1986"/> Soto's basal cell nevus,<ref name="Tarín-1998"/>[[Treacher-Collins Syndrome]],<ref name="Tarín-1998"/> [[tuberous sclerosis]],<ref name="Tarín-1998"/> [[Meatal stenosis|Urethral stenosis]],<ref name="Tarín-1998"/> [[Waardenburg syndrome]],<ref name="Tarín-1998"/> and [[Wilms' tumor]].<ref name="Tarín-1998"/>


===Paternal mortality before adulthood of child===
===Paternal mortality before adulthood of child===
Line 63: Line 63:


===Fertility===
===Fertility===
Older men have decreased pregnancy rates, increased time to pregnancy, and increased infertility at a given point in time. Increasing paternal age may also increase the risk of reproductive failure, which has led some researchers to compare age 40 to the "Amber Light" in a man's reproductive life.
Older men have decreased pregnancy rates, increased time to pregnancy, and increased infertility at a given point in time. Increasing paternal age may also increase the risk of reproductive failure, which has led some researchers to compare age 40 to the "Amber Light" in a man's reproductive life.


==Mechanisms==
==Mechanisms==
Line 69: Line 69:


===DNA mutations===
===DNA mutations===
In contrast to [[oogenesis]], which involves 22 [[mitosis|mitotic]] divisions before birth and 2 [[meiosis|meiotic]] divisions after birth, [[spermatogenesis]] involves 30 mitotic divisions before puberty, and 4 mitotic and 2 meiotic divisions after puberty. Advanced paternal age may therefore lead to "copy error" in replication or the accumulation of [[mutagen]]s, thereby leading to ''[[Mutation|de novo]]'' mutations in sperm DNA. A study of 78 [[Iceland]]ic families found that each additional year in the age of the father causes about two new mutations in the child.<ref name="Kong-2012">{{Cite journal | author = Kong A, Frigge ML, Masson G, Besenbacher S, Sulem P, Magnusson G, Gudjonsson SA, Sigurdsson A, Jonasdottir A, Jonasdottir A, Wong WS, Sigurdsson G, Walters GB, Steinberg S, Helgason H, Thorleifsson G, Gudbjartsson DF, Helgason A, Magnusson OT, Thorsteinsdottir U, Stefansson K | title = Rate of de novo mutations and the importance of father's age to disease risk | journal = Nature | volume = 488 | issue = 7412 | pages = 471–5 | year = 2012 | doi = 10.1038/nature11396 | url = http://www.nature.com/nature/journal/v488/n7412/full/nature11396.html | pmid = 22914163 | pmc = 3548427 }}</ref> . Regarding the increased risk at very young paternal ages, an international study indicates that the DNA mutation rate in very young fathers may also be elevated.<ref name="Forster-2015">{{Cite journal | author = Forster P, Hohoff C, Dunkelmann B, Schürenkamp M, Pfeiffer H, Neuhuber F, Brinkmann B | title = Elevated germline mutation rate in teenage fathers | journal = Proc R Soc B | volume = 282 | issue = 1803 | pages = 1–6 | year = 2015 | doi = 10.1098/rspb.2014.2898 | url = http://rspb.royalsocietypublishing.org/content/282/1803/20142898 | pmid = 25694621 | pmc = 4345458 }}</ref>
In contrast to [[oogenesis]], which involves 22 [[mitosis|mitotic]] divisions before birth and 2 [[meiosis|meiotic]] divisions after birth, [[spermatogenesis]] involves 30 mitotic divisions before puberty, and 4 mitotic and 2 meiotic divisions after puberty. Advanced paternal age may therefore lead to "copy error" in replication or the accumulation of [[mutagen]]s, thereby leading to ''[[Mutation|de novo]]'' mutations in sperm DNA. A study of 78 [[Iceland]]ic families found that each additional year in the age of the father causes about two new mutations in the child.<ref name="Kong-2012">{{Cite journal | author = Kong A, Frigge ML, Masson G, Besenbacher S, Sulem P, Magnusson G, Gudjonsson SA, Sigurdsson A, Jonasdottir A, Jonasdottir A, Wong WS, Sigurdsson G, Walters GB, Steinberg S, Helgason H, Thorleifsson G, Gudbjartsson DF, Helgason A, Magnusson OT, Thorsteinsdottir U, Stefansson K | title = Rate of de novo mutations and the importance of father's age to disease risk | journal = Nature | volume = 488 | issue = 7412 | pages = 471–5 | year = 2012 | doi = 10.1038/nature11396 | url = http://www.nature.com/nature/journal/v488/n7412/full/nature11396.html | pmid = 22914163 | pmc = 3548427 }}</ref> . Regarding the increased risk at very young paternal ages, an international study indicates that the DNA mutation rate in very young fathers may also be elevated.<ref name="Forster-2015">{{Cite journal | author = Forster P, Hohoff C, Dunkelmann B, Schürenkamp M, Pfeiffer H, Neuhuber F, Brinkmann B | title = Elevated germline mutation rate in teenage fathers | journal = Proc R Soc B | volume = 282 | issue = 1803 | pages = 1–6 | year = 2015 | doi = 10.1098/rspb.2014.2898 | url = http://rspb.royalsocietypublishing.org/content/282/1803/20142898 | pmid = 25694621 | pmc = 4345458 }}</ref>


===DNA methylation===
===DNA methylation===
Line 78: Line 78:


===Clonal expansion of spermatogonial cells===
===Clonal expansion of spermatogonial cells===
A distinct set of monogenetic congenital disorders, collectively known as paternal age effect (PAE) disorders, are all caused by a small number of dominantly-acting point mutations and almost exclusively originate from unaffected fathers, suggesting that the mutations are taking place during spermatogenesis. Mutations in the fibroblast growth factor receptor genes FGFR2 cause Apert syndrome, Crouzon syndrome, and Pfeiffer syndrome. Mutations in the FGFR3 gene lead to the formation of achondroplasia, thanatophoric dysplasia, hypochondroplasia, and Muenke syndrome. In recent studies of multiple endocrine neoplasia Type 2A and 2B and Apert syndrome, a total of 92 new mutations were discovered and all were found to be paternal in origin.<ref name="Schuffenecker-1997" /> These studies which show an extreme paternal bias for PAE mutations is argued to be caused by the distinct phenomenon of clonal expansion of spermatogonial cells with gain-of-function protein properties. This mechanism known as “selfish selection”, results in an enrichment of mutant sperm over time and may preferentially carry alterations in genes that could have far-reaching consequences for the health of future generations.<ref name="Goriely and Wilkie 2013">{{cite journal | author = Goriely A, Wilkie AOM | title = "Selfish Spermatogonial Selection": A Novel mechanism for the association between advanced paternal age and neurodevelopmental disorders | journal = Am. J. Psychiatry | volume = 170 | pages = 599–608 | year = 2013 | doi=10.1176/appi.ajp.2013.12101352}}</ref>
A distinct set of monogenetic congenital disorders, collectively known as paternal age effect (PAE) disorders, are all caused by a small number of dominantly-acting point mutations and almost exclusively originate from unaffected fathers, suggesting that the mutations are taking place during spermatogenesis. Mutations in the fibroblast growth factor receptor genes FGFR2 cause Apert syndrome, Crouzon syndrome, and Pfeiffer syndrome. Mutations in the FGFR3 gene lead to the formation of achondroplasia, thanatophoric dysplasia, hypochondroplasia, and Muenke syndrome. In recent studies of multiple endocrine neoplasia Type 2A and 2B and Apert syndrome, a total of 92 new mutations were discovered and all were found to be paternal in origin.<ref name="Schuffenecker-1997" /> These studies which show an extreme paternal bias for PAE mutations is argued to be caused by the distinct phenomenon of clonal expansion of spermatogonial cells with gain-of-function protein properties. This mechanism known as “selfish selection”, results in an enrichment of mutant sperm over time and may preferentially carry alterations in genes that could have far-reaching consequences for the health of future generations.<ref name="Goriely and Wilkie 2013">{{cite journal | author = Goriely A, Wilkie AOM | title = "Selfish Spermatogonial Selection": A Novel mechanism for the association between advanced paternal age and neurodevelopmental disorders | journal = Am. J. Psychiatry | volume = 170 | pages = 599–608 | year = 2013 | doi=10.1176/appi.ajp.2013.12101352}}</ref>


===Social associations===
===Social associations===
Line 84: Line 84:


===Semen===
===Semen===
A 2001 review on variation in semen quality and fertility by male age concluded that older men had lower semen volume, lower sperm motility, and a decreased percent of normal sperm. One common factor is the abnormal regulation of sperm once a mutation arises. It has been seen that once taking place, the mutation will almost always be positively selected for and over time will lead to the mutant sperm replacing all non-mutant sperm. In younger males, this process is corrected and regulated by the growth factor receptor-RAS signal transduction pathway.<ref>{{cite journal|last1=Goriely|first1=Anne|last2=Wilkie|first2=Andrew|title=Paternal Age Effect Mutations and Selfish Spermatogonial Selection: Causes and Consequences for Human Disease|journal=The American Journal of Human Genetics|date=2012|volume=90|issue=2|pages=175–200|doi=10.1016/j.ajhg.2011.12.017|pmid=22325359}}</ref>
A 2001 review on variation in semen quality and fertility by male age concluded that older men had lower semen volume, lower sperm motility, and a decreased percent of normal sperm. One common factor is the abnormal regulation of sperm once a mutation arises. It has been seen that once taking place, the mutation will almost always be positively selected for and over time will lead to the mutant sperm replacing all non-mutant sperm. In younger males, this process is corrected and regulated by the growth factor receptor-RAS signal transduction pathway.<ref>{{cite journal|last1=Goriely|first1=Anne|last2=Wilkie|first2=Andrew|title=Paternal Age Effect Mutations and Selfish Spermatogonial Selection: Causes and Consequences for Human Disease|journal=The American Journal of Human Genetics|date=2012|volume=90|issue=2|pages=175–200|doi=10.1016/j.ajhg.2011.12.017|pmid=22325359}}</ref>


A 2014 review indicated that increasing male age is associated with declines in many semen traits, including semen volume and percentage motility. However, this review also found that sperm concentration did not decline as male age increased.<ref>{{cite journal|last1=Johnson|first1=Sheri L.|last2=Dunleavy|first2=Jessica|last3=Gemmell|first3=Neil J.|last4=Nakagawa|first4=Shinichi|title=Consistent age-dependent declines in human semen quality: A systematic review and meta-analysis|journal=Ageing Research Reviews|date=January 2015|volume=19|pages=22–33|doi=10.1016/j.arr.2014.10.007}}</ref>
A 2014 review indicated that increasing male age is associated with declines in many semen traits, including semen volume and percentage motility. However, this review also found that sperm concentration did not decline as male age increased.<ref>{{cite journal|last1=Johnson|first1=Sheri L.|last2=Dunleavy|first2=Jessica|last3=Gemmell|first3=Neil J.|last4=Nakagawa|first4=Shinichi|title=Consistent age-dependent declines in human semen quality: A systematic review and meta-analysis|journal=Ageing Research Reviews|date=January 2015|volume=19|pages=22–33|doi=10.1016/j.arr.2014.10.007}}</ref>
Line 94: Line 94:
In 1912, [[Wilhelm Weinberg]], a German physician, was the first person to hypothesize that non-inherited cases of [[achondroplasia]] could be more common in last-born children than in children born earlier to the same set of parents.<ref name="Crow-2000">{{Cite journal | author = Crow JF | title = The origins, patterns and implications of human spontaneous mutation | journal = Nature Reviews Genetics | volume = 1 | issue = 1 | pages = 40–7 | year = 2000 | doi = 10.1038/35049558 | url = http://ender.bu.edu/~tgardner/be209/lectures/12/articles/Crow.JF_NatRevGen_00.pdf | pmid = 11262873 }}</ref> Although Weinberg "made no distinction between paternal age, maternal age and [[birth order]]" in his hypothesis, by 1953 the term "paternal age effect" had occurred in the medical literature on achondroplasia.<ref name="Crow-2000"/><ref name="Krooth-1953">{{Cite journal | author = Krooth RS | title = Comments on the estimation of the mutation rate for achondroplasia | journal = American Journal of Human Genetics | volume = 5 | issue = 4 | pages = 373–6 | year = 1953 | pmid = 13104383 | pmc = 1716528 }}</ref>{{rp|375}}
In 1912, [[Wilhelm Weinberg]], a German physician, was the first person to hypothesize that non-inherited cases of [[achondroplasia]] could be more common in last-born children than in children born earlier to the same set of parents.<ref name="Crow-2000">{{Cite journal | author = Crow JF | title = The origins, patterns and implications of human spontaneous mutation | journal = Nature Reviews Genetics | volume = 1 | issue = 1 | pages = 40–7 | year = 2000 | doi = 10.1038/35049558 | url = http://ender.bu.edu/~tgardner/be209/lectures/12/articles/Crow.JF_NatRevGen_00.pdf | pmid = 11262873 }}</ref> Although Weinberg "made no distinction between paternal age, maternal age and [[birth order]]" in his hypothesis, by 1953 the term "paternal age effect" had occurred in the medical literature on achondroplasia.<ref name="Crow-2000"/><ref name="Krooth-1953">{{Cite journal | author = Krooth RS | title = Comments on the estimation of the mutation rate for achondroplasia | journal = American Journal of Human Genetics | volume = 5 | issue = 4 | pages = 373–6 | year = 1953 | pmid = 13104383 | pmc = 1716528 }}</ref>{{rp|375}}


Scientific interest in paternal age effects increased in the late 20th and early 21st centuries because the average paternal age increased in countries such as the United Kingdom,<ref name="Bray-2006">{{cite journal | author = Bray I, Gunnell D, Smith GD | title = Advanced paternal age: How old is too old? | journal = J Epidemiol Community Health| volume = 60 | issue = 10 | pages = 851–3 | year = 2006 | doi = 10.1136/jech.2005.045179 | pmid = 16973530 | pmc = 2566050}}</ref> Australia,<ref>{{cite web |url=http://www.abs.gov.au/ausstats/abs@.nsf/Products/92A449676DCE3145CA25766A00120E3D?opendocument |title= 3301.0 - Births, Australia, 2008. Summary of findings. Births|author= Australian Bureau of Statistics|date=11 November 2009 |accessdate=25 February 2010}}</ref> and Germany, and because birth rates for fathers aged 30–54 years have risen between 1980 and 2006 in the United States.<ref>{{cite journal |title= Births: final data for 2006 |author= Martin JA, Hamilton BE, Sutton PD, Ventura SJ, Menacker F, Kirmeyer S, Mathews TJ |publisher= National Center for Health Statistics |location= Hyattsville, MD | journal = National Vital Statistics Reports | volume = 57 | issue = 7 | pages = 1–104 | year=2009 |url= http://www.cdc.gov/nchs/data/nvsr/nvsr57/nvsr57_07.pdf |accessdate=25 February 2010}}</ref> Possible reasons for the increases in average paternal age include increasing life expectancy and increasing rates of divorce and remarriage. Despite recent increases in average paternal age, however, the [[Ramjit Raghav|oldest father]] documented in the medical literature was born in 1840: George Isaac Hughes was 94 years old at the time of the birth of his son by his second wife, a 1935 article in the ''[[Journal of the American Medical Association]]'' stated that his fertility "has been definitely and affirmatively checked up medically," and he fathered a daughter in 1936 at age 96.<ref name="Seymour-1935">{{Cite journal | author = Seymour FI, Duffy C, Koerner A | title = A case of authenticated fertility in a man, aged 94 | journal = J Am Med Assoc | volume = 105 | issue = 18 | pages = 1423–4| year = 1935 | url = http://jama.ama-assn.org/cgi/reprint/105/18/1423 | doi=10.1001/jama.1935.92760440002009a}}</ref><ref>{{cite news |title= A father again at 96; North Carolinan's baby a sister to boy born two years ago |newspaper= New York Times |date= 4 June 1936 | page = 10}}</ref>
Scientific interest in paternal age effects increased in the late 20th and early 21st centuries because the average paternal age increased in countries such as the United Kingdom,<ref name="Bray-2006">{{cite journal | author = Bray I, Gunnell D, Smith GD | title = Advanced paternal age: How old is too old? | journal = J Epidemiol Community Health| volume = 60 | issue = 10 | pages = 851–3 | year = 2006 | doi = 10.1136/jech.2005.045179 | pmid = 16973530 | pmc = 2566050}}</ref> Australia,<ref>{{cite web |url=http://www.abs.gov.au/ausstats/abs@.nsf/Products/92A449676DCE3145CA25766A00120E3D?opendocument |title= 3301.0 - Births, Australia, 2008. Summary of findings. Births|author= Australian Bureau of Statistics|date=11 November 2009 |accessdate=25 February 2010}}</ref> and Germany, and because birth rates for fathers aged 30–54 years have risen between 1980 and 2006 in the United States.<ref>{{cite journal |title= Births: final data for 2006 |author= Martin JA, Hamilton BE, Sutton PD, Ventura SJ, Menacker F, Kirmeyer S, Mathews TJ |publisher= National Center for Health Statistics |location= Hyattsville, MD | journal = National Vital Statistics Reports | volume = 57 | issue = 7 | pages = 1–104 | year=2009 |url= http://www.cdc.gov/nchs/data/nvsr/nvsr57/nvsr57_07.pdf |accessdate=25 February 2010}}</ref> Possible reasons for the increases in average paternal age include increasing life expectancy and increasing rates of divorce and remarriage. Despite recent increases in average paternal age, however, the [[Ramjit Raghav|oldest father]] documented in the medical literature was born in 1840: George Isaac Hughes was 94 years old at the time of the birth of his son by his second wife, a 1935 article in the ''[[Journal of the American Medical Association]]'' stated that his fertility "has been definitely and affirmatively checked up medically," and he fathered a daughter in 1936 at age 96.<ref name="Seymour-1935">{{Cite journal | author = Seymour FI, Duffy C, Koerner A | title = A case of authenticated fertility in a man, aged 94 | journal = J Am Med Assoc | volume = 105 | issue = 18 | pages = 1423–4| year = 1935 | url = http://jama.ama-assn.org/cgi/reprint/105/18/1423 | doi=10.1001/jama.1935.92760440002009a}}</ref><ref>{{cite news |title= A father again at 96; North Carolinan's baby a sister to boy born two years ago |newspaper= New York Times |date= 4 June 1936 | page = 10}}</ref>
In 2012, two 96-year-old men, [[List of people with the most children#fathers|Nanu Ram Jogi]] and [[Ramjit Raghav]], both from India, claimed to have fathered children that year.,<ref>[http://articles.timesofindia.indiatimes.com/2012-10-16/india/34497611_1_oldest-man-nanu-ram-jogi-first-child Nanu Ram Jogi fathers another child aged 96], article in the Times of India, 16 October 2012</ref><ref name="Daily Mail 1">{{cite news|newspaper=[[The Daily Mail]]|url=http://www.dailymail.co.uk/news/article-2442492/Worlds-oldest-dad-97-devastated-wife-leaves-following-disappearance-son.html|title=World's oldest dad, 97, devastated after wife leaves him following disappearance of their son|date=3 October 2013|location=London}}</ref>
In 2012, two 96-year-old men, [[List of people with the most children#fathers|Nanu Ram Jogi]] and [[Ramjit Raghav]], both from India, claimed to have fathered children that year.,<ref>[http://articles.timesofindia.indiatimes.com/2012-10-16/india/34497611_1_oldest-man-nanu-ram-jogi-first-child Nanu Ram Jogi fathers another child aged 96], article in the Times of India, 16 October 2012</ref><ref name="Daily Mail 1">{{cite news|newspaper=[[The Daily Mail]]|url=http://www.dailymail.co.uk/news/article-2442492/Worlds-oldest-dad-97-devastated-wife-leaves-following-disappearance-son.html|title=World's oldest dad, 97, devastated after wife leaves him following disappearance of their son|date=3 October 2013|location=London}}</ref>



Revision as of 15:01, 28 May 2015

The paternal age effect is the statistical relationship between paternal age at conception and biological effects on the child. Such effects can relate to miscarriage risk, birthweight, congenital disorders and health-related conditions including mortality and longevity, or risk of psychological outcomes. The age-risk relationship appears to be "saucer-shaped" (U-shaped) or J-shaped, with a higher risk in very young fathers, the lowest risk at 20–40 years of age, and increased risk in older men.

A 2009 review concludes that the absolute risk for genetic anomalies in offspring is low, and states that "There is no clear association between adverse health outcome and paternal age but longitudinal studies are needed."[1] Some research even indicates a longevity advantage for offspring of older fathers.[2]

On the other hand, the genetic quality of sperm, as well as its volume and motility, all typically decrease with age,[3] leading the population geneticist James F. Crow to claim that the "greatest mutational health hazard to the human genome is fertile older males".[4]

Because paternity did not become provable until 1970, and the cost of definitively establishing it only recently became low enough to do it on widespread basis, this has meant that only limited scientific research into paternal age effect problems of degraded DNA has been done. Harry Fisch, a physician who has done research in this area, says that research into paternal age effect degradation of DNA is "in its infancy".[5]

Health effects

Evidence for a paternal age effect has been proposed for a number of conditions, diseases and other effects. In many of these, the statistical evidence of association is weak, and the association may be related by confounding factors, or behavioral differences.[6] Conditions proposed to show correlation with paternal age include the following:[1]

Pregnancy effects

Studies published between 2002 and 2008 have been consistent in associating advanced paternal age with miscarriage,[7][8][9][10] stillbirth,[11] and fetal death (which includes both miscarriage and stillbirth).[12] A 2002 study linked paternal age with pre-eclampsia, a complication of pregnancy that can be associated with adverse health outcomes for both the pregnant woman and the fetus.[13]

Birth outcomes

A systematic review published in 2010 concluded the risk of low birthweight in infants with paternal age is "saucer-shaped"; that is, the highest risks occur at low and at high paternal ages.[14] Compared with a paternal age of 25–28 years as a reference group, the odds ratio for low birthweight was approximately 1.1 at a paternal age of 20 and approximately 1.2 at a paternal age of 50.[14] There was no association of paternal age with preterm births or with small for gestational age births.[14]

In a 2008 retrospective study found a paternal age of 40 years or greater was not associated with infant death in the first year of life.[15] However, the risks of were elevated for infants whose fathers were less than 20 years old.[15][16]

Mental illness

Schizophrenia is associated with advanced paternal age with 12 out of 14 studies supporting a relationship.[17] Paternal age older than 55 is a moderate risk factor for schizophrenia.[18]

Most studies examining autism spectrum disorder (ASD) and advanced paternal age have demonstrated an association between the two, although there also appears to be an increase with maternal age.[19]

The risk of bipolar disorder ("manic depression") particularly for early-onset disease, is J-shaped, with the lowest risk for children of 20-24-year old fathers, a twofold risk for younger fathers, and a threefold risk for fathers >50 years old. There is no similar relationship with maternal age.[20]

Cancers

Paternal age may be associated with an increased risk of breast cancer,[21] but the association is weak and there are confounding effects.[1]

Diabetes mellitus

A higher paternal age is a possible risk factor for type 1 diabetes.[22]

Down syndrome

It appears that a paternal-age effect exists with respect to Down syndrome, but is very small in comparison to maternal-age effect.[23]

Intelligence

By 1998, "Intellectual disability or decreased learning capacity of unknown aetiology" was thought to be associated with increased paternal age.[24] In 2005, Malaspina and colleagues detected an "inverted U-shaped relationship" between paternal age and intelligence quotients (IQs) in 44,175 people from Israel.[25] There was a peak at paternal ages of 25-44; fathers younger than 25 and older than 44 tended to have children with lower IQs.[25] Malaspina et al. also reviewed the literature and found that "at least a half dozen other studies ... have demonstrated significant associations between paternal age and human intelligence."[25]

A 2009 study by Saha et al. examined 33,437 children at 8 months, 4 years, and 7 years.[26][27] The researchers found that paternal age was associated with poorer scores in almost all neurocognitive tests used, but that maternal age was associated with better scores on the same tests.[26] An editorial accompanying the paper by Saha et al. emphasized the importance of controlling for socioeconomic status in studies of paternal age and intelligence.[28] A 2010 paper from Spain provided further evidence that average paternal age is elevated in cases of intellectual disability.[29]

Mortality and longevity of offspring

A 2008 paper found a U-shaped association between paternal age and the overall mortality rate in children (i.e., mortality rate up to age 18).[30] Although the relative mortality rates were higher, the absolute numbers were low, because of the relatively low occurrence of genetic abnormality. The study has been criticized for not adjusting for maternal health, which could have a large effect on child mortality.[31] Surprisingly, the researchers found a correlation between paternal age and offspring death by injury or poisoning, indicating the need to control for social and behavioral confounding factors.[32]

In 2012, Eisenberg et al. published a study which showed that greater age at paternity tends to increase telomere length in offspring for up to two generations. Since telomere length has effects on health and mortality, this may have effects on health and the "pace of senescence" in these offspring. The authors speculated that this effect may provide a mechanism by which populations have some plasticity in adapting longevity to different social and ecological contexts.[2]

Monogenetic disorders

Studies reveal that the following list of congenital disorders, collectively known as paternal age effect (PAE) disorders, are all caused by a small number of dominantly-acting point mutations and almost exclusively originate from unaffected fathers, suggesting that the mutations are taking place during spermatogenesis. Mutations in the fibroblast growth factor receptor genes FGFR2, cause Apert syndrome,[24][33][34][35] Crouzon syndrome,[24][33] and Pfeiffer syndrome.[24][33][36] Mutations in the FGFR3 gene lead to the formation of achondroplasia,[24][33] thanatophoric dysplasia,[24] hypochondroplasia, and Muenke syndrome.[36] These disorders occur spontaneously as a result of advanced paternal age and at the rate of 1 in 30,000 for achondroplasia births.[37][38] Other conditions involving the mutations in the RET gene lead to multiple endocrine neoplasia type 2A and 2B, the PTPN11 gene which leads to Noonan syndrome,[39] and the HRAS mutations which cause Costello syndrome.[24][40][41][42]

Other conditions

Other conditions and diseases which have been suggested as having a possible correlation with paternal age include: chondrodystrophy,[43] acrodysostosis,[24] aniridia,[24]basal cell nevus syndrome,[24] cataracts,[24] Cerebral palsy, athetoid/dystonic,[24] CHARGE syndrome,[44][45] cleft palate,[24][46] cleidocranial dysostosis,[24] craniosynostosis,[47]diaphragmatic hernia,[46] Duchenne muscular dystrophy,[24] exostoses, multiple,[24] congenital malformations in extremities,[24][48] fibrodysplasia ossificans progressiva,[24][33] Heart defects,[24][43] Hemiplegia,[24] Hemophilia A,[24] Klinefelter's syndrome,[49] Lesch-Nyhan syndrome,[24] Marfan syndrome,[33] nasal aplasia,[24]neural tube defects,[24] oculodentodigital syndrome,[24] osteogenesis imperfecta type IIA,[24]polycystic kidney disease,[24] Polyposis coli,[24] Preauricular cyst,[24] Progeria,[24] psychotic disorders,[24][50] von Recklinghausen neurofibromatosis,[24][33] retinitis pigmentosa,[24]retinoblastoma, bilateral,[24] situs inversus,[43] Soto's basal cell nevus,[24]Treacher-Collins Syndrome,[24] tuberous sclerosis,[24] Urethral stenosis,[24] Waardenburg syndrome,[24] and Wilms' tumor.[24]

Paternal mortality before adulthood of child

The risk of the father dying before the child becomes an adult increases by increased paternal age, such as can be demonstrated by the following data from France in 2007:[51]

Paternal age at childbirth 25 30 35 40 45
Risk of father not surviving until child's 18th birthday (in %)[51] 2.2 3.3 5.4 8.3 12.1

Fertility

Older men have decreased pregnancy rates, increased time to pregnancy, and increased infertility at a given point in time.[52] Increasing paternal age may also increase the risk of reproductive failure, which has led some researchers to compare age 40 to the "Amber Light" in a man's reproductive life.[53]

Mechanisms

Several hypothesized chains of causality exist whereby increased paternal age may lead to health effects.

DNA mutations

In contrast to oogenesis, which involves 22 mitotic divisions before birth and 2 meiotic divisions after birth, spermatogenesis involves 30 mitotic divisions before puberty, and 4 mitotic and 2 meiotic divisions after puberty.[44] Advanced paternal age may therefore lead to "copy error" in replication or the accumulation of mutagens, thereby leading to de novo mutations in sperm DNA.[44] A study of 78 Icelandic families found that each additional year in the age of the father causes about two new mutations in the child.[54] . Regarding the increased risk at very young paternal ages, an international study indicates that the DNA mutation rate in very young fathers may also be elevated.[55]

DNA methylation

Epigenetic processes such as parental imprinting could explain the association between paternal age and schizophrenia.[56]

Telomere length

In 2012, Eisenberg et al. published a study which showed that greater age at paternity tends to increase telomere length in offspring for up to two generations. Since telomere length has effects on health and mortality, this may have effects on health and the "pace of senescence" in these offspring. The authors speculated that this effect may provide a mechanism by which populations have some plasticity in adapting longevity to different social and ecological contexts.[2]

Clonal expansion of spermatogonial cells

A distinct set of monogenetic congenital disorders, collectively known as paternal age effect (PAE) disorders, are all caused by a small number of dominantly-acting point mutations and almost exclusively originate from unaffected fathers, suggesting that the mutations are taking place during spermatogenesis. Mutations in the fibroblast growth factor receptor genes FGFR2 cause Apert syndrome, Crouzon syndrome, and Pfeiffer syndrome. Mutations in the FGFR3 gene lead to the formation of achondroplasia, thanatophoric dysplasia, hypochondroplasia, and Muenke syndrome. In recent studies of multiple endocrine neoplasia Type 2A and 2B and Apert syndrome, a total of 92 new mutations were discovered and all were found to be paternal in origin.[40][41][57] These studies which show an extreme paternal bias for PAE mutations is argued to be caused by the distinct phenomenon of clonal expansion of spermatogonial cells with gain-of-function protein properties. This mechanism known as “selfish selection”, results in an enrichment of mutant sperm over time and may preferentially carry alterations in genes that could have far-reaching consequences for the health of future generations.[58]

Social associations

Later age at parenthood is associated with a more stable family environment, higher socio-economic position, higher income and better living conditions, as well as better parenting practices,[51] but it is more or less uncertain whether these entities are effects of advanced parental age, are contributors to advanced parental age, or common effects of a certain state such as personality type.

Semen

A 2001 review on variation in semen quality and fertility by male age concluded that older men had lower semen volume, lower sperm motility, and a decreased percent of normal sperm.[52] One common factor is the abnormal regulation of sperm once a mutation arises. It has been seen that once taking place, the mutation will almost always be positively selected for and over time will lead to the mutant sperm replacing all non-mutant sperm. In younger males, this process is corrected and regulated by the growth factor receptor-RAS signal transduction pathway.[59]

A 2014 review indicated that increasing male age is associated with declines in many semen traits, including semen volume and percentage motility. However, this review also found that sperm concentration did not decline as male age increased.[60]

X-linked effects

Some classify the paternal age effect as one of two different types. One effect is directly related to advanced paternal age and autosomal mutations in the offspring. The other effect is an indirect effect in relation to mutations on the X chromosome which are passed to daughters who are then at risk for having sons with X-linked diseases.[61]

History

In 1912, Wilhelm Weinberg, a German physician, was the first person to hypothesize that non-inherited cases of achondroplasia could be more common in last-born children than in children born earlier to the same set of parents.[44] Although Weinberg "made no distinction between paternal age, maternal age and birth order" in his hypothesis, by 1953 the term "paternal age effect" had occurred in the medical literature on achondroplasia.[44][62]: 375 

Scientific interest in paternal age effects increased in the late 20th and early 21st centuries because the average paternal age increased in countries such as the United Kingdom,[63] Australia,[64] and Germany,[65] and because birth rates for fathers aged 30–54 years have risen between 1980 and 2006 in the United States.[66] Possible reasons for the increases in average paternal age include increasing life expectancy and increasing rates of divorce and remarriage.[65] Despite recent increases in average paternal age, however, the oldest father documented in the medical literature was born in 1840: George Isaac Hughes was 94 years old at the time of the birth of his son by his second wife, a 1935 article in the Journal of the American Medical Association stated that his fertility "has been definitely and affirmatively checked up medically," and he fathered a daughter in 1936 at age 96.[65]: 329 [67][68] In 2012, two 96-year-old men, Nanu Ram Jogi and Ramjit Raghav, both from India, claimed to have fathered children that year.,[69][70]

Medical assessment

The American College of Medical Genetics recommends obstetric ultrasonography at 18–20 weeks gestation in cases of advanced paternal age "to evaluate fetal growth and development," but it notes that this procedure "is unlikely to detect many of the conditions of interest." They also note that there is no standard definition of "advanced paternal age."[71] Bray et al. (2006) expressed an opinion that any adverse effects of advanced paternal age "should be weighed up against potential social advantages for children born to older fathers who are more likely to have progressed in their career and to have achieved financial security."[63]

Geneticist James F. Crow described mutations that have a direct visible effect on the child's health and also mutations that can be latent or have minor visible effects on the child's health; many such minor or latent mutations allow the child to reproduce, but cause more serious problems for grandchildren, greatgrandchildren and later generations.[4]

See also

References

  1. ^ a b c H. Tournaye, "Male Reproductive Ageing," in Bewley, Ledger, and Nikolaou, eds., Reproductive Ageing, Cambridge University Press (2009), ISBN 9781906985134 (accessed 15 November 2013)
  2. ^ a b c Eisenberg, Dan T.A.; Hayes, M. Geoffrey; Kuzawa, Christopher W. (June 11, 2012). "Delayed paternal age of reproduction in humans is associated with longer telomeres across two generations of descendants". Proc Natl Acad Sci U S A. 109 (26): 10251–10256. doi:10.1073/pnas.1202092109. Retrieved 28 June 2014.
  3. ^ Gurevich, Rachel (June 10, 2008). "Does Age Affect Male Fertility?". About.com:Fertility. About.com. Retrieved 14 February 2010.
  4. ^ a b Crow, James F. (August 5, 1997). "The high spontaneous mutation rate: Is it a health risk?". Proceedings of the National Academy of Sciences. 94 (16): 8380–8386. doi:10.1073/pnas.94.16.8380. PMC 33757. PMID 9237985. {{cite journal}}: |access-date= requires |url= (help)
  5. ^ Vanderbes, Jennifer (June 25, 2011). "What's That Ticking Sound? The Male Biological Clock". Wall Street Journal. Retrieved 3 December 2013.
  6. ^ see Correlation does not imply causation
  7. ^ de la Rochebrochard E, Thonneau P (2002). "Paternal age and maternal age are risk factors for miscarriage; results of a multicentre European study". Hum Reprod. 17 (6): 1649–56. doi:10.1093/humrep/17.6.1649. PMID 12042293.
  8. ^ Slama R, Bouyer J, Windham G, Fenster L, Werwatz A, Swan SH (2005). "Influence of paternal age on the risk of spontaneous abortion". Am J Epidemiol. 161 (9): 816–23. doi:10.1093/aje/kwi097. PMID 15840613.{{cite journal}}: CS1 maint: multiple names: authors list (link)
  9. ^ Kleinhaus K, Perrin M, Friedlander Y, Paltiel O, Malaspina D, Harlap S (2006). "Paternal age and spontaneous abortion" (PDF). Obstet Gynecol. 108 (2): 369–77. doi:10.1097/01.AOG.0000224606.26514.3a. PMID 16880308.{{cite journal}}: CS1 maint: multiple names: authors list (link)
  10. ^ Belloc S, Cohen-Bacrie P, Benkhalifa M, Cohen-Bacrie M, De Mouzon J, Hazout A, Ménézo Y (2008). "Effect of maternal and paternal age on pregnancy and miscarriage rates after intrauterine insemination". Reprod Biomed Online. 17 (3): 392–7. doi:10.1016/S1472-6483(10)60223-4. PMID 18765010.{{cite journal}}: CS1 maint: multiple names: authors list (link) [dead link]
  11. ^ Astolfi P, De Pasquale A, Zonta LA (2004). "Late paternity and stillbirth risk". Hum Reprod. 19 (11): 2497–501. doi:10.1093/humrep/deh449. PMID 15319387.{{cite journal}}: CS1 maint: multiple names: authors list (link)
  12. ^ Nybo Andersen AM, Hansen KD, Andersen PK, Davey Smith G (2004). "Advanced paternal age and risk of fetal death: a cohort study". Am J Epidemiol. 160 (12): 1214–22. doi:10.1093/aje/kwh332. PMID 15583374.{{cite journal}}: CS1 maint: multiple names: authors list (link)
  13. ^ Harlap S, Paltiel O, Deutsch L, Knaanie A, Masalha S, Tiram E, Caplan LS, Malaspina D, Friedlander Y (2002). "Paternal age and preeclampsia". Epidemiology. 13 (6): 660–7. doi:10.1097/01.EDE.0000031708.99480.70. PMID 12410007.{{cite journal}}: CS1 maint: multiple names: authors list (link)
  14. ^ a b c Shah PS; Knowledge Synthesis Group on determinants of preterm/low birthweight births (2010). "Paternal factors and low birthweight, preterm, and small for gestational age births: a systematic review". Am J Obstet Gynecol. 202 (2): 103–23. doi:10.1016/j.ajog.2009.08.026. PMID 20113689.{{cite journal}}: CS1 maint: multiple names: authors list (link)
  15. ^ a b Chen XK, Wen SW, Krewski D, Fleming N, Yang Q, Walker MC (2008). "Paternal age and adverse birth outcomes: teenager or 40+, who is at risk?". Hum Reprod. 23 (6): 1290–6. doi:10.1093/humrep/dem403. PMID 18256111.{{cite journal}}: CS1 maint: multiple names: authors list (link)
  16. ^ Laurance J (7 February 2008). "Older fathers have healthier offspring". The Independent. London. Retrieved 25 February 2010.
  17. ^ Kirkpatrick, B; Messias, E; Harvey, PD; Fernandez-Egea, E; Bowie, CR (November 2008). "Is schizophrenia a syndrome of accelerated aging?". Schizophrenia bulletin. 34 (6): 1024–32. doi:10.1093/schbul/sbm140. PMID 18156637.
  18. ^ Torrey EF, Buka S, Cannon TD, Goldstein JM, Seidman LJ, Liu T, Hadley T, Rosso IM, Bearden C, Yolken RH (2009). "Paternal age as a risk factor for schizophrenia: how important is it?". Schizophr Res. 114 (1–3): 1–5. doi:10.1016/j.schres.2009.06.017. PMID 19683417.{{cite journal}}: CS1 maint: multiple names: authors list (link)
  19. ^ Kolevzon A, Gross R, Reichenberg A (2007). "Prenatal and perinatal risk factors for autism: a review and integration of findings". Arch Pediatr Adolesc Med. 161 (4): 326–333. doi:10.1001/archpedi.161.4.326. PMID 17404128.{{cite journal}}: CS1 maint: multiple names: authors list (link)
  20. ^ Frans EM, Sandin S, Reichenberg A, Lichtenstein P, Långström N, Hultman CM (2008). "Advancing Paternal Age and Bipolar Disorder". Arch Gen Psychiatry. 65 (9): 1034–1040. doi:10.1001/archpsyc.65.9.1034. PMID 18762589.{{cite journal}}: CS1 maint: multiple names: authors list (link)
  21. ^ Xue F, Michels KB (2007). "Intrauterine factors and risk of breast cancer: a systematic review and meta-analysis of current evidence". Lancet Oncol. 8 (12): 1088–100. doi:10.1016/S1470-2045(07)70377-7. PMID 18054879.
  22. ^ Chronic disease epidemiology and control (3rd ed. ed.). Washington, DC: American Public Health Association. 2010. p. 301. ISBN 9780875531922. {{cite book}}: |edition= has extra text (help)
  23. ^ Girirajan S (2009). "Parental-age effects in Down syndrome" (PDF). J Genet. 88 (1): 1–7. doi:10.1007/s12041-009-0001-6. PMID 19417538.
  24. ^ a b c d e f g h i j k l m n o p q r s t u v w x y z aa ab ac ad ae af ag ah ai aj ak al am an Tarín JJ, Brines J, Cano A (1998). "Long-term effects of delayed parenthood". Hum Reprod. 13 (9): 2371–6. doi:10.1093/humrep/13.9.2371. PMID 9806250.{{cite journal}}: CS1 maint: multiple names: authors list (link)
  25. ^ a b c Malaspina D, Reichenberg A, Weiser M, Fennig S, Davidson M, Harlap S, Wolitzky R, Rabinowitz J, Susser E, Knobler HY (2005). "Paternal age and intelligence: implications for age-related genomic changes in male germ cells". Psychiatr Genet. 15 (2): 117–25. doi:10.1097/00041444-200506000-00008. PMID 15900226.{{cite journal}}: CS1 maint: multiple names: authors list (link)
  26. ^ a b Saha S, Barnett AG, Foldi C, Burne TH, Eyles DW, Buka SL, McGrath JJ (2009). Brayne, Carol (ed.). "Advanced Paternal Age Is Associated with Impaired Neurocognitive Outcomes during Infancy and Childhood". PLoS Med. 6 (3): e40. doi:10.1371/journal.pmed.1000040. PMC 2653549. PMID 19278291.{{cite journal}}: CS1 maint: multiple names: authors list (link) CS1 maint: unflagged free DOI (link)
  27. ^ Dayton L (10 March 2009). "Not the sharpest? Blame old dad". The Australian. Retrieved 25 February 2010.
  28. ^ Cannon M (2009). "Contrasting Effects of Maternal and Paternal Age on Offspring Intelligence: The clock ticks for men too". PLoS Med. 6 (3): e42. doi:10.1371/journal.pmed.1000042. PMC 2653550. PMID 19278293.{{cite journal}}: CS1 maint: unflagged free DOI (link)
  29. ^ Lopez-Castroman J, Gómez DD, Belloso JJ, Fernandez-Navarro P, Perez-Rodriguez MM, Villamor IB, Navarrete FF, Ginestar CM, Currier D, Torres MR, Navio-Acosta M, Saiz-Ruiz J, Jimenez-Arriero MA, Baca-Garcia E (2010). "Differences in maternal and paternal age between schizophrenia and other psychiatric disorders". Schizophr Res. 116 (2–3): 184–90. doi:10.1016/j.schres.2009.11.006. PMID 19945257.{{cite journal}}: CS1 maint: multiple names: authors list (link)
  30. ^ Zhu JL, Vestergaard M, Madsen KM, Olsen J (2008). "Paternal age and mortality in children". Eur J Epidemiol. 23 (7): 443–7. doi:10.1007/s10654-008-9253-3. PMID 18437509.{{cite journal}}: CS1 maint: multiple names: authors list (link)
  31. ^ "In this particular study, no adjustment was made for the health of the mother, and this could have had a large effect on child mortality." National Health Service (UK), "Older Dads and the Death of Children," (accessed 15 November 2013)
  32. ^ Tournaye 2009, p. 102
  33. ^ a b c d e f g Risch N, Reich EW, Wishnick MM, McCarthy JG (1987). "Spontaneous mutation and parental age in humans". American Journal of Human Genetics. 41 (2): 218–48. PMC 1684215. PMID 3618593.{{cite journal}}: CS1 maint: multiple names: authors list (link)
  34. ^ Tolarova MM, Harris JA, Ordway DE, Vargervik K (1997). "Birth prevalence, mutation rate, sex ratio, parents' age, and ethnicity in Apert syndrome". American Journal of Medical Genetics. 72 (4): 394–8. doi:10.1002/(SICI)1096-8628(19971112)72:4<394::AID-AJMG4>3.0.CO;2-R. PMID 9375719.{{cite journal}}: CS1 maint: multiple names: authors list (link)
  35. ^ Yoon SR, Qin J, Glaser RL, Jabs EW, Wexler NS, Sokol R, Arnheim N, Calabrese P (2009). Walsh, Bruce (ed.). "The Ups and Downs of Mutation Frequencies during Aging Can Account for the Apert Syndrome Paternal Age Effect". PLoS Genet. 5 (7): e1000558. doi:10.1371/journal.pgen.1000558. PMC 2700275. PMID 19593369.{{cite journal}}: CS1 maint: multiple names: authors list (link) CS1 maint: unflagged free DOI (link)
  36. ^ a b Goriely A, Wilkie AOM (2012). "Paternal Age Effect Mutations and Selfish Spermatogonial Selection: Causes and Consequences For Human Disease". Am. J. Human Genet. 90 (2): 175–200. doi:10.1016/j.ajhg.2011.12.017. PMID 22325359.
  37. ^ Orioli I.M.; et al. (1986). "The birth prevalence rates for the skeletal dysplasias". J. Med. Genet. 23: 328–332. doi:10.1136/jmg.23.4.328. {{cite journal}}: Explicit use of et al. in: |author= (help)
  38. ^ Waller D.K. (2008). "The population-based prevalence of achondroplasia and thanatophoric dysplasia in selected regions of the U.S.". Am. J. Med. Genet. 146A: 2385–9. doi:10.1002/ajmg.a.32485.
  39. ^ Tartaglia M.; et al. (2004). "Paternal germline origin and sex-ratio distortion in transmission of PTPN11 mutations in Noonan syndrome". Am. J. Hum. Genet. 75: 492–7. doi:10.1086/423493. {{cite journal}}: Explicit use of et al. in: |author= (help)
  40. ^ a b Carlson KM, Bracamontes J, Jackson CE, Clark R, Lacroix A, Wells SA Jr, Goodfellow PJ (1994). "Parent-of-origin effects in multiple endocrine neoplasia type 2B". American Journal of Human Genetics. 55 (6): 1076–82. PMC 1918453. PMID 7977365.{{cite journal}}: CS1 maint: multiple names: authors list (link)
  41. ^ a b Schuffenecker I, Ginet N, Goldgar D, Eng C, Chambe B, Boneu A, Houdent C, Pallo D, Schlumberger M, Thivolet C, Lenoir GM; Le Groupe d'Etude des Tumeurs a Calcitonine (1997). "Prevalence and parental origin of de novo RET mutations in multiple endocrine neoplasia type 2A and familial medullary thyroid carcinoma. Le Groupe d'Etude des Tumeurs a Calcitonine". American Journal of Human Genetics. 60 (1): 233–7. PMC 1712555. PMID 8981969.{{cite journal}}: CS1 maint: multiple names: authors list (link)
  42. ^ Sol-Church K.; et al. (2006). "Paternal bias in parental origin of HRAS mutations in Costello syndrome". Hum. Mutat. 27: 736–741. doi:10.1002/humu.20381. {{cite journal}}: Explicit use of et al. in: |author= (help)
  43. ^ a b c Lian ZH, Zack MM, Erickson JD (1986). "Paternal age and the occurrence of birth defects". American Journal of Human Genetics. 39 (5): 648–60. PMC 1684057. PMID 3788977.{{cite journal}}: CS1 maint: multiple names: authors list (link)
  44. ^ a b c d e Crow JF (2000). "The origins, patterns and implications of human spontaneous mutation" (PDF). Nature Reviews Genetics. 1 (1): 40–7. doi:10.1038/35049558. PMID 11262873.
  45. ^ Blake KD, Prasad C (2006). "CHARGE syndrome" (PDF). Orphanet J Rare Dis. 1: 34. doi:10.1186/1750-1172-1-34. PMC 1586184. PMID 16959034.{{cite journal}}: CS1 maint: unflagged free DOI (link)
  46. ^ a b Green RF, Devine O, Crider KS, Olney RS, Archer N, Olshan AF, Shapira SK (2010). "Association of Paternal Age and Risk for Major Congenital Anomalies from the National Birth Defects Prevention Study, 1997–2004". Annals of Epidemiology. 20 (3): 241–9. doi:10.1016/j.annepidem.2009.10.009. PMC 2824069. PMID 20056435.{{cite journal}}: CS1 maint: multiple names: authors list (link)
  47. ^ Rasmussen SA, Yazdy MM, Frías JL, Honein MA (2008). "Priorities for public health research on craniosynostosis: summary and recommendations from a Centers for Disease Control and Prevention-sponsored meeting" (PDF). American Journal of Medical Genetics. 146A (2): 149–58. doi:10.1002/ajmg.a.32106. PMID 18080327.{{cite journal}}: CS1 maint: multiple names: authors list (link)
  48. ^ Zhu JL, Madsen KM, Vestergaard M, Olesen AV, Basso O, Olsen J (2005). "Paternal age and congenital malformations". Hum Reprod. 20 (11): 3173–7. doi:10.1093/humrep/dei186. PMID 16006461.{{cite journal}}: CS1 maint: multiple names: authors list (link)
  49. ^ BRS Genetics - Dudek 2009
  50. ^ Miller B, Pihlajamaa J, Haukka J, Cannon M, Henriksson M, Heilä H, Huttunen M, Tanskanen A, Lönnqvist J, Suvisaari J, Kirkpatrick B (16 Feb 2010). "Paternal age and mortality in nonaffective psychosis". Schizophr Res. 121 (1–3): 218–26. doi:10.1016/j.schres.2010.01.020. PMID 20163936.{{cite journal}}: CS1 maint: multiple names: authors list (link)
  51. ^ a b c Attention: This template ({{cite doi}}) is deprecated. To cite the publication identified by doi:10.1093/humupd/dmr040, please use {{cite journal}} (if it was published in a bona fide academic journal, otherwise {{cite report}} with |doi=10.1093/humupd/dmr040 instead.
  52. ^ a b Kidd SA, Eskenazi B, Wyrobek AJ (2001). "Effects of male age on semen quality and fertility: a review of the literature". Fertil Steril. 75 (2): 237–48. doi:10.1016/S0015-0282(00)01679-4. PMID 11172821.{{cite journal}}: CS1 maint: multiple names: authors list (link)
  53. ^ De La Rochebrochard, E; McElreavey, K; Thonneau, P (2003). "Paternal age over 40 years: the "amber light" in the reproductive life of men?". Journal of andrology. 24 (4): 459–65. PMID 12826682.
  54. ^ Kong A, Frigge ML, Masson G, Besenbacher S, Sulem P, Magnusson G, Gudjonsson SA, Sigurdsson A, Jonasdottir A, Jonasdottir A, Wong WS, Sigurdsson G, Walters GB, Steinberg S, Helgason H, Thorleifsson G, Gudbjartsson DF, Helgason A, Magnusson OT, Thorsteinsdottir U, Stefansson K (2012). "Rate of de novo mutations and the importance of father's age to disease risk". Nature. 488 (7412): 471–5. doi:10.1038/nature11396. PMC 3548427. PMID 22914163.{{cite journal}}: CS1 maint: multiple names: authors list (link)
  55. ^ Forster P, Hohoff C, Dunkelmann B, Schürenkamp M, Pfeiffer H, Neuhuber F, Brinkmann B (2015). "Elevated germline mutation rate in teenage fathers". Proc R Soc B. 282 (1803): 1–6. doi:10.1098/rspb.2014.2898. PMC 4345458. PMID 25694621.{{cite journal}}: CS1 maint: multiple names: authors list (link)
  56. ^ Perrin MC, Brown AS, Malaspina D (2007). "Aberrant Epigenetic Regulation Could Explain the Relationship of Paternal Age to Schizophrenia". Schizophr Bull. 33 (6): 1270–3. doi:10.1093/schbul/sbm093. PMC 2779878. PMID 17712030.{{cite journal}}: CS1 maint: multiple names: authors list (link)
  57. ^ Moloney D.; et al. (1996). "Exclusive paternal origin of new mutations in Apert syndrome". Nat. Genet. 13: 48–53. doi:10.1038/ng0596-48. PMID 8673103. {{cite journal}}: Explicit use of et al. in: |author= (help)
  58. ^ Goriely A, Wilkie AOM (2013). ""Selfish Spermatogonial Selection": A Novel mechanism for the association between advanced paternal age and neurodevelopmental disorders". Am. J. Psychiatry. 170: 599–608. doi:10.1176/appi.ajp.2013.12101352.
  59. ^ Goriely, Anne; Wilkie, Andrew (2012). "Paternal Age Effect Mutations and Selfish Spermatogonial Selection: Causes and Consequences for Human Disease". The American Journal of Human Genetics. 90 (2): 175–200. doi:10.1016/j.ajhg.2011.12.017. PMID 22325359.
  60. ^ Johnson, Sheri L.; Dunleavy, Jessica; Gemmell, Neil J.; Nakagawa, Shinichi (January 2015). "Consistent age-dependent declines in human semen quality: A systematic review and meta-analysis". Ageing Research Reviews. 19: 22–33. doi:10.1016/j.arr.2014.10.007.
  61. ^ "Definition of Advanced paternal age".
  62. ^ Krooth RS (1953). "Comments on the estimation of the mutation rate for achondroplasia". American Journal of Human Genetics. 5 (4): 373–6. PMC 1716528. PMID 13104383.
  63. ^ a b Bray I, Gunnell D, Smith GD (2006). "Advanced paternal age: How old is too old?". J Epidemiol Community Health. 60 (10): 851–3. doi:10.1136/jech.2005.045179. PMC 2566050. PMID 16973530.{{cite journal}}: CS1 maint: multiple names: authors list (link)
  64. ^ Australian Bureau of Statistics (11 November 2009). "3301.0 - Births, Australia, 2008. Summary of findings. Births". Retrieved 25 February 2010.
  65. ^ a b c Kühnert B, Nieschlag E (2004). "Reproductive functions of the ageing male". Hum Reprod Update. 10 (4): 327–39. doi:10.1093/humupd/dmh030. PMID 15192059.
  66. ^ Martin JA, Hamilton BE, Sutton PD, Ventura SJ, Menacker F, Kirmeyer S, Mathews TJ (2009). "Births: final data for 2006" (PDF). National Vital Statistics Reports. 57 (7). Hyattsville, MD: National Center for Health Statistics: 1–104. Retrieved 25 February 2010.{{cite journal}}: CS1 maint: multiple names: authors list (link)
  67. ^ Seymour FI, Duffy C, Koerner A (1935). "A case of authenticated fertility in a man, aged 94". J Am Med Assoc. 105 (18): 1423–4. doi:10.1001/jama.1935.92760440002009a.{{cite journal}}: CS1 maint: multiple names: authors list (link)
  68. ^ "A father again at 96; North Carolinan's baby a sister to boy born two years ago". New York Times. 4 June 1936. p. 10.
  69. ^ Nanu Ram Jogi fathers another child aged 96, article in the Times of India, 16 October 2012
  70. ^ "World's oldest dad, 97, devastated after wife leaves him following disappearance of their son". The Daily Mail. London. 3 October 2013.
  71. ^ Toriello HV, Meck JM; Professional Practice and Guidelines Committee, American College of Medical Genetics (2008). "Statement on guidance for genetic counseling in advanced paternal age". Genet Med. 10 (6): 457–60. doi:10.1097/GIM.0b013e318176fabb. PMC 3111019. PMID 18496227.{{cite journal}}: CS1 maint: multiple names: authors list (link)

Further reading

  • Fisch H, Braun S (2005). The male biological clock: the startling news about aging, sexuality, and fertility in men. New York: Free Press. ISBN 0-7432-5991-2.
  • Gavrilov, L.A., Gavrilova, N.S. Human longevity and parental age at conception. In: J.-M.Robine, T.B.L. Kirkwood, M. Allard (eds.) Sex and Longevity: Sexuality, Gender, Reproduction, Parenthood, Berlin, Heidelberg: Springer-Verlag, 2000, 7-31.
  • Gavrilov, L.A., Gavrilova, N.S. Parental age at conception and offspring longevity. Reviews in Clinical Gerontology, 1997, 7: 5-12.
  • Gavrilov, L.A., Gavrilova, N.S. When Fatherhood Should Stop? Letter. Science, 1997, 277(5322): 17-18.