Paternal age effect

Last updated

The paternal age effect is the statistical relationship between the father's age at conception and biological effects on the child. [1] Such effects can relate to birthweight, congenital disorders, life expectancy and psychological outcomes. [2] A 2017 review found that while severe health effects are associated with higher paternal age, the total increase in problems caused by paternal age is low. [3] While paternal age has increased since 1960–1970, this is not seen as a major public health concern. [3]

Contents

The genetic quality of sperm, as well as its volume and motility, may decrease with age, [4] [5] leading the population geneticist James F. Crow to claim that the "greatest mutational health hazard to the human genome is fertile older males". [6]

The paternal age effect was first proposed implicitly by physician Wilhelm Weinberg in 1912 [7] and explicitly by psychiatrist Lionel Penrose in 1955. [8] DNA-based research started more recently, in 1998, in the context of paternity testing.

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 behavioural differences. [9] [3] Conditions proposed to show correlation with paternal age include the following: [10]

Single-gene disorders

Advanced paternal age may be associated with a higher risk for certain single-gene disorders caused by mutations of the FGFR2 , FGFR3 and RET genes. [11] These conditions are Apert syndrome, Crouzon syndrome, Pfeiffer syndrome, achondroplasia, thanatophoric dysplasia, multiple endocrine neoplasia type 2, and multiple endocrine neoplasia type 2b. [11] The most significant effect concerns achondroplasia (a form of dwarfism), which might occur in about 1 in 1,875 children fathered by men over 50, compared to 1 in 15,000 in the general population. [12] However, the risk for achondroplasia is still considered clinically negligible. [13] The FGFR genes may be particularly prone to a paternal age effect due to selfish spermatogonial selection, whereby the influence of spermatogonial mutations in older men is enhanced because cells with certain mutations have a selective advantage over other cells (see § DNA mutations). [14]

Pregnancy effects

Several studies have reported that advanced paternal age is associated with an increased risk of miscarriage. [15] The strength of the association differs between studies. [16] It has been suggested that these miscarriages are caused by chromosome abnormalities in the sperm of aging men. [15] An increased risk for stillbirth has also been suggested for pregnancies fathered by men over 45. [16]

Birth outcomes

A systematic review published in 2010 concluded that the graph of the risk of low birthweight in infants with paternal age is "saucer-shaped" (U-shaped); that is, the highest risks occur at low and at high paternal ages. [17] 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. [17] There was no association of paternal age with preterm births or with small for gestational age births. [17]

Mental illness

Schizophrenia is associated with advanced paternal age. [18] [19] [20] Some 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. [21]

In one study, the risk of bipolar disorder, particularly for early-onset disease, is J-shaped, with the lowest risk for children of 20- to 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. [22] A second study also found a risk of schizophrenia in both fathers above age 50 and fathers below age 25. The risk in younger fathers was noted to affect only male children. [23]

A 2010 study found the relationship between parental age and psychotic disorders to be stronger with maternal age than paternal age. [24]

A 2016 review concluded that the mechanism behind the reported associations was still not clear, with evidence both for selection of individuals liable to psychiatric illness into late fatherhood and evidence for causative mutations. The mechanisms under discussion are not mutually exclusive. [25]

A 2017 review concluded that the vast majority of studies supported a relationship between older paternal age and autism and schizophrenia but that there is less convincing and also inconsistent evidence for associations with other psychiatric illnesses. [3]

Cancers

Paternal age may be associated with an increased risk of breast cancer, [26] but the association is weak and there are confounding effects. [10]

According to a 2017 review, there is consistent evidence of an increase in incidence of childhood acute lymphoblastic leukemia with paternal age. Results for associations with other childhood cancers are more mixed (e.g. retinoblastoma) or generally negative. [3]

Diabetes mellitus

High paternal age has been suggested as a risk factor for type 1 diabetes, [27] but research findings are inconsistent, and a clear association has not been established. [28] [29]

Down syndrome

It appears that a paternal-age effect might exist with respect to Down syndrome, but it is very small in comparison to the maternal-age effect. [30] [31]

Intelligence

A review in 2005 found a U-shaped relationship between paternal age and low intelligence quotients (IQs). [32] The highest IQ was found at paternal ages of 25–29; fathers younger than 25 and older than 29 tended to have children with lower IQs. [32] It also found that "at least a half dozen other studies ... have demonstrated significant associations between paternal age and human intelligence." [32] A 2009 study examined children at 8 months, 4 years and 7 years and found that higher paternal age was associated with poorer scores in almost all neurocognitive tests used but that higher maternal age was associated with better scores on the same tests; [33] this was a reverse effect to that observed in the 2005 review, which found that maternal age began to correlate with lower intelligence at a younger age than paternal age, [32] however two other past studies were in agreement with the 2009 study's results. [24] An editorial accompanying the 2009 paper emphasized the importance of controlling for socioeconomic status in studies of paternal age and intelligence. [34] A 2010 study from Spain also found an association between advanced paternal age and intellectual disability. [24]

On the other hand, later research concluded that previously reported negative associations might be explained by confounding factors, especially parental intelligence and education. A re-analysis of the 2009 study found that the paternal age effect could be explained by adjusting for maternal education and number of siblings. [35] A 2012 Scottish study found no significant association between paternal age and intelligence, after adjusting what was initially an inverse-U association for both parental education and socioeconomic status as well as number of siblings. [36] A 2013 study of half a million Swedish men adjusted for genetic confounding by comparing brothers and found no association between paternal age and offspring IQ. [37] Another study from 2014 found an initially positive association between paternal age and offspring IQ that disappeared when adjusting for parental IQs. [38]

Life expectancy

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). [39] 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. [40] The researchers also found a correlation between paternal age and offspring death by injury or poisoning, indicating the need to control for social and behavioral confounding factors. [41]

In 2012, a study 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 rate of aging 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. [42]

Associated social and genetic characteristics

Father's age versus father's risk of death
(among French population) [43]
Father's age
at birth
Risk of father's death
before child's 18th birthday
201.5%
252.2%
303.3%
355.4%
408.3%
4512.1%

Parents do not decide when to reproduce randomly. This implies that paternal age effects may be confounded by social and genetic predictors of reproductive timing.

A simulation study concluded that reported paternal age effects on psychiatric disorders in the epidemiological literature are too large to be explained only by mutations. They conclude that a model in which parents with a genetic liability to psychiatric illness tend to reproduce later better explains the literature. [9]

Later age at parenthood is also associated with a more stable family environment, with older parents being less likely to divorce or change partners. [43] Older parents also tend to occupy a higher socio-economic position and report feeling more devoted to their children and satisfied with their family. [43] On the other hand, the risk of the father dying before the child becomes an adult increases with paternal age. [43]

To adjust for genetic liability, some studies compare full siblings. Additionally, or alternatively, studies statistically adjust for some or all of these confounding factors. Using sibling comparisons or adjusting for more covariates frequently changes the direction or magnitude of paternal age effects. For example, one study drawing on Finnish census data concluded that increases in offspring mortality with paternal age could be explained completely by parental loss. [44] On the other hand, a population-based cohort study drawing on 2.6 million records from Sweden found that risk of attention deficit hyperactivity disorder was only positively associated with paternal age when comparing siblings. [45]

Mechanisms

Several hypothesized chains of causality exist whereby increased paternal age may lead to health effects. [16] [46] There are different types of genome mutations, with distinct mutation mechanisms:

Telomere length

Telomeres are repetitive genetic sequences at both ends of each chromosome that protect the structure of the chromosome. [47] As men age, most telomeres shorten, but sperm telomeres increase in length. [16] The offspring of older fathers have longer telomeres in both their sperm and white blood cells. [16] [47] A large study showed a positive paternal, but no independent maternal age effect on telomere length. Because the study used twins, it could not compare siblings who were discordant for paternal age. It found that telomere length was 70% heritable. [48]

DNA point mutations

In contrast to oogenesis, the production of sperm cells is a lifelong process. [16] Each year after puberty, spermatogonia (precursors of the spermatozoa) divide meiotically about 23 times. [46] By the age of 40, the spermatogonia will have undergone about 660 such divisions, compared to 200 at age 20. [46] Copying errors might sometimes happen during the DNA replication preceding these cell divisions, which may lead to new (de novo) mutations in the sperm DNA. [14]

The selfish spermatogonial selection hypothesis proposes that the influence of spermatogonial mutations in older men is further enhanced because cells with certain mutations have a selective advantage over other cells. [46] [49] Such an advantage would allow the mutated cells to increase in number through clonal expansion. [46] [49] In particular, mutations that affect the RAS pathway, which regulates spermatogonial proliferation, appear to offer a competitive advantage to spermatogonial cells, while also leading to diseases associated with paternal age. [49]

DNA fragmentation

During the past two decades evidence has accumulated that pregnancy loss as well as reduced rate of success with assisted reproductive technologies is linked to impaired sperm chromosome integrity and DNA fragmentation [50] . Advanced paternal age was shown to be associated with a significant increase in DNA fragmentation in a recent systematic review (where 17 out of the 19 studies considered showed such an association) [51] .

Epigenetic changes

DNA methylation DNA methylation.jpg
DNA methylation

The production of sperm cells involves DNA methylation, an epigenetic process that regulates the expression of genes. [46] Improper genomic imprinting and other errors sometimes occur during this process, which can affect the expression of genes related to certain disorders, increasing the offspring's susceptibility. The frequency of these errors appears to increase with age. This could explain the association between paternal age and schizophrenia.; [52] Paternal age affects offspring's behavior, possibly via an epigenetic mechanism recruiting a transcriptional repressor REST. [53]

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, a decreased percent of normal sperm, as well as decreased pregnancy rates, increased time to pregnancy and increased infertility at a given point in time. [54] When controlling for the age of the female partner, comparisons between men under 30 and men over 50 found relative decreases in pregnancy rates between 23% and 38%. [54]

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. [55]

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. [56]

History

Birth defects were acknowledged in the children of older men and women even in antiquity. In book six of Plato's Republic, Socrates states that men and women should have children in the "prime of their life" which is stated to be twenty in a woman and thirty in a man. He states that in his proposed society men should be forbidden to father children in their fifties and that the offspring of such unions should be considered "the offspring of darkness and strange lust." He suggests appropriate punishments be administered to the offenders and their offspring. [57] [58]

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. [59] Weinberg "made no distinction between paternal age, maternal age and birth order" in his hypothesis. In 1953, Krooth used the term "paternal age effect" in the context of achondroplasia, but mistakenly thought the condition represented a maternal age effect. [59] [60] :375 The paternal age effect for achondroplasia was described by Lionel Penrose in 1955. At a DNA level, the paternal age effect was first reported in 1998 in routine paternity tests. [61]

Scientific interest in paternal age effects is relevant because the average paternal age increased in countries such as the United Kingdom, [62] Australia [63] and Germany, [64] and because birth rates for fathers aged 30–54 years have risen between 1980 and 2006 in the United States. [65] Possible reasons for the increases in average paternal age include increasing life expectancy and increasing rates of divorce and remarriage. [64] 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. [64] [66] [67]

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 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; [11] it is commonly defined as age 40 or above, but the effect increases linearly with paternal age, rather than appearing at any particular age. [68] According to a 2006 review, 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." [62]

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, great-grandchildren and later generations. [6]

See also

Related Research Articles

<span class="mw-page-title-main">Inbreeding</span> Reproduction by closely related organisms

Inbreeding is the production of offspring from the mating or breeding of individuals or organisms that are closely related genetically. By analogy, the term is used in human reproduction, but more commonly refers to the genetic disorders and other consequences that may arise from expression of deleterious recessive traits resulting from incestuous sexual relationships and consanguinity. Animals avoid incest only rarely.

Achondroplasia is a genetic disorder with an autosomal dominant pattern of inheritance whose primary feature is dwarfism. It is the most common cause of dwarfism and affects about 1 in 27,500 people. In those with the condition, the arms and legs are short, while the torso is typically of normal length. Those affected have an average adult height of 131 centimetres for males and 123 centimetres (4 ft) for females. Other features can include an enlarged head with prominent forehead and underdevelopment of the midface. Complications can include sleep apnea or recurrent ear infections. Achondroplasia includes the extremely rare short-limb skeletal dysplasia with severe combined immunodeficiency.

Infertility is the inability of an animal or plant to reproduce by natural means. It is usually not the natural state of a healthy adult, except notably among certain eusocial species. It is the normal state of a human child or other young offspring, because they have not undergone puberty, which is the body's start of reproductive capacity.

Fertility is the ability to conceive a child. The fertility rate is the average number of children born during an individual's lifetime and is quantified demographically. Conversely, infertility is the difficulty or inability to reproduce naturally. In general, infertility is defined as not being able to conceive a child after one year of unprotected sex. Infertility is widespread, with fertility specialists available all over the world to assist parents and couples who experience difficulties conceiving a baby.

<span class="mw-page-title-main">Birth defect</span> Condition present at birth regardless of cause; human disease or disorder developed prior to birth

A birth defect, also known as a congenital disorder, is an abnormal condition that is present at birth regardless of its cause. Birth defects may result in disabilities that may be physical, intellectual, or developmental. The disabilities can range from mild to severe. Birth defects are divided into two main types: structural disorders in which problems are seen with the shape of a body part and functional disorders in which problems exist with how a body part works. Functional disorders include metabolic and degenerative disorders. Some birth defects include both structural and functional disorders.

A maternal effect is a situation where the phenotype of an organism is determined not only by the environment it experiences and its genotype, but also by the environment and genotype of its mother. In genetics, maternal effects occur when an organism shows the phenotype expected from the genotype of the mother, irrespective of its own genotype, often due to the mother supplying messenger RNA or proteins to the egg. Maternal effects can also be caused by the maternal environment independent of genotype, sometimes controlling the size, sex, or behaviour of the offspring. These adaptive maternal effects lead to phenotypes of offspring that increase their fitness. Further, it introduces the concept of phenotypic plasticity, an important evolutionary concept. It has been proposed that maternal effects are important for the evolution of adaptive responses to environmental heterogeneity.

<span class="mw-page-title-main">Germline mutation</span> Inherited genetic variation

A germline mutation, or germinal mutation, is any detectable variation within germ cells. Mutations in these cells are the only mutations that can be passed on to offspring, when either a mutated sperm or oocyte come together to form a zygote. After this fertilization event occurs, germ cells divide rapidly to produce all of the cells in the body, causing this mutation to be present in every somatic and germline cell in the offspring; this is also known as a constitutional mutation. Germline mutation is distinct from somatic mutation.

Recurrent miscarriage or recurrent pregnancy loss (RPL) is the spontaneous loss of 2-3 pregnancies that is estimated to affect up to 5% of women. The exact number of pregnancy losses and gestational weeks used to define RPL differs among medical societies. In the majority of cases, the exact cause of pregnancy loss is unexplained despite genetic testing and a thorough evaluation. When a cause for RPL is identified, almost half are attributed to a chromosomal abnormality. RPL has been associated with several risk factors including parental and genetic factors, congenital and acquired anatomical conditions, lifestyle factors, endocrine disorders, thrombophila, immunological factors, and infections. The American Society of Reproductive Medicine recommends a thorough evaluation after 2 consecutive pregnancy losses, however, this can differ from recommendations by other medical societies. RPL evaluation be evaluated by numerous tests and imaging studies depending on the risk factors. These range from cytogenetic studies, blood tests for clotting disorders, hormone levels, diabetes screening, thyroid function tests, sperm analysis, antibody testing, and imaging studies. Treatment is typically tailored to the relevant risk factors and test findings. RPL can have a significant impact on the psychological well-being of couples and has been associated with higher levels of depression, anxiety, and stress. Therefore, it is recommended that appropriate screening and management be considered by medical providers.  

<span class="mw-page-title-main">Grandmother hypothesis</span> Hypothesis concerning the existence of menopause

The grandmother hypothesis is a hypothesis to explain the existence of menopause in human life history by identifying the adaptive value of extended kin networking. It builds on the previously postulated "mother hypothesis" which states that as mothers age, the costs of reproducing become greater, and energy devoted to those activities would be better spent helping her offspring in their reproductive efforts. It suggests that by redirecting their energy onto those of their offspring, grandmothers can better ensure the survival of their genes through younger generations. By providing sustenance and support to their kin, grandmothers not only ensure that their genetic interests are met, but they also enhance their social networks which could translate into better immediate resource acquisition. This effect could extend past kin into larger community networks and benefit wider group fitness.

<span class="mw-page-title-main">Virility</span> Traits of masculinity viewed in a positive light

Virility refers to any of a wide range of masculine characteristics viewed positively. Virile means "marked by strength or force". Virility is commonly associated with vigour, health, sturdiness, and constitution, especially in the fathering of children. In this last sense, virility is to men as fertility is to women. Virile has become obsolete in referring to a "nubile" young woman, or "a maid that is Marriageable or ripe for a Husband, or Virill".

<span class="mw-page-title-main">Causes of autism</span> Proposed causes of autism

The causes of autism are environmental or genetic factors that predispose an individual to develop autism, also known as autism spectrum disorder (ASD). Many causes of autism have been proposed, but understanding of the theory of causation of autism is incomplete. Attempts have been made to incorporate the known genetic and environmental causes into a comprehensive causative framework. ASD is a neurodevelopmental disorder marked by impairments in communicative ability and social interaction and restricted/repetitive behaviors, interests, or activities not suitable for the individual's developmental stage. The severity of symptoms and functional impairment vary between individuals.

Advanced maternal age, in a broad sense, is the instance of a woman being of an older age at a stage of reproduction, although there are various definitions of specific age and stage of reproduction. The variability in definitions is in part explained by the effects of increasing age occurring as a continuum rather than as a threshold effect.

Metabolic imprinting refers to the long-term physiological and metabolic effects that an offspring's prenatal and postnatal environments have on them. Perinatal nutrition has been identified as a significant factor in determining an offspring's likelihood of it being predisposed to developing cardiovascular disease, obesity, and type 2 diabetes amongst other conditions.

Prenatal stress is exposure of an expectant mother to psychosocial or physical stress, which can be caused by daily life events or by environmental hardships. This psychosocial or physical stress that the expectant mother is experiencing has an effect on the fetus. According to the Developmental Origins of Health and Disease (DOHaD), a wide range of environmental factors a woman may experience during the perinatal period can contribute to biological impacts and changes in the fetus that then causes health risks later in the child's life.

<span class="mw-page-title-main">Health effects of pesticides</span> How pesticides affect human health

Health effects of pesticides may be acute or delayed in those who are exposed. Acute effects can include pesticide poisoning, which may be a medical emergency. Strong evidence exists for other, long-term negative health outcomes from pesticide exposure including birth defects, fetal death, neurodevelopmental disorder, cancer, and neurologic illness including Parkinson's disease. Toxicity of pesticides depend on the type of chemical, route of exposure, dosage, and timing of exposure.

<span class="mw-page-title-main">Imprinted brain hypothesis</span> Conjecture on the causes of autism and psychosis

The imprinted brain hypothesis is an unsubstantiated hypothesis in evolutionary psychology regarding the causes of autism spectrum and schizophrenia spectrum disorders, first presented by Bernard Crespi and Christopher Badcock in 2008. It claims that certain autistic and schizotypal traits are opposites, and that this implies the etiology of the two conditions must be at odds.

The epigenetics of schizophrenia is the study of how inherited epigenetic changes are regulated and modified by the environment and external factors and how these changes influence the onset and development of, and vulnerability to, schizophrenia. Epigenetics concerns the heritability of those changes, too. Schizophrenia is a debilitating and often misunderstood disorder that affects up to 1% of the world's population. Although schizophrenia is a heavily studied disorder, it has remained largely impervious to scientific understanding; epigenetics offers a new avenue for research, understanding, and treatment.

The development of an animal model of autism is one approach researchers use to study potential causes of autism. Given the complexity of autism and its etiology, researchers often focus only on single features of autism when using animal models.

Obesity is defined as an abnormal accumulation of body fat, usually 20% or more over an individual's ideal body weight. This is often described as a body mass index (BMI) over 30. However, BMI does not account for whether the excess weight is fat or muscle, and is not a measure of body composition. For most people, however, BMI is an indication used worldwide to estimate nutritional status. Obesity is usually the result of consuming more calories than the body needs and not expending that energy by doing exercise. There are genetic causes and hormonal disorders that cause people to gain significant amounts of weight but this is rare. People in the obese category are much more likely to suffer from fertility problems than people of normal healthy weight.

Anne Goriely is a British geneticist who is a professor of human genetics at the University of Oxford. Her research investigates the molecular mechanisms that underpin genetic variation, particularly mutations in the male germline.

References

  1. "paternal age effect" . Retrieved 28 May 2015.
  2. Amaral, David; Dawson, Geraldine; Geschwind, Daniel (17 June 2011). Autism Spectrum Disorders. Oxford University Press, USA. ISBN   9780195371826.
  3. 1 2 3 4 5 Nybo Andersen AM, Urhoj SK (February 2017). "Is advanced paternal age a health risk for the offspring?". Fertility and Sterility. 107 (2): 312–318. doi: 10.1016/j.fertnstert.2016.12.019 . PMID   28088314.
  4. Gurevich, Rachel (10 June 2008). "Does Age Affect Male Fertility?". About.com:Fertility. About.com. Retrieved 14 February 2010.
  5. Kovac JR, Addai J, Smith RP, Coward RM, Lamb DJ, Lipshultz LI (November 2013). "The effects of advanced paternal age on fertility". Asian Journal of Andrology. 15 (6): 723–8. doi: 10.1038/aja.2013.92 . PMC   3854059 . PMID   23912310.
  6. 1 2 Crow JF (August 1997). "The high spontaneous mutation rate: is it a health risk?". Proceedings of the National Academy of Sciences of the United States of America. 94 (16): 8380–6. Bibcode:1997PNAS...94.8380C. doi: 10.1073/pnas.94.16.8380 . PMC   33757 . PMID   9237985.
  7. Weinberg, W (1912). "Zur Vererbung des Zwergwuchses" [On the inheritance of dwarfism]. Arch Rassen-u Gesell Biol (in German). 9: 710–718. NAID   10017956735.
  8. Penrose LS (August 1955). "Parental age and mutation". Lancet. 269 (6885): 312–3. doi:10.1016/s0140-6736(55)92305-9. PMID   13243724.
  9. 1 2 Gratten J, Wray NR, Peyrot WJ, McGrath JJ, Visscher PM, Goddard ME (July 2016). "Risk of psychiatric illness from advanced paternal age is not predominantly from de novo mutations". Nature Genetics. 48 (7): 718–24. doi:10.1038/ng.3577. PMID   27213288. S2CID   19816925.
  10. 1 2 Tournaye, Herman (June 2009). "Male Reproductive Ageing". In Bewley, Susan; Ledger, William; Nikolaou, Dimitrios (eds.). Reproductive Ageing. Cambridge University Press. pp. 95–104. ISBN   978-1-906985-13-4.
  11. 1 2 3 Toriello HV, Meck JM (June 2008). "Statement on guidance for genetic counseling in advanced paternal age". Genetics in Medicine. 10 (6): 457–60. doi: 10.1097/GIM.0b013e318176fabb . PMC   3111019 . PMID   18496227.
  12. Kovac JR, Addai J, Smith RP, Coward RM, Lamb DJ, Lipshultz LI (November 2013). "The effects of advanced paternal age on fertility". Asian Journal of Andrology. 15 (6): 723–8. doi:10.1038/aja.2013.92. PMC   3854059 . PMID   23912310.
  13. Czeizel AE, Czeizel B, Vereczkey A (January 2013). "The participation of prospective fathers in preconception care". Clinical Medicine Insights. Reproductive Health. 7: 1–9. doi:10.4137/CMRH.S10930. PMC   3888083 . PMID   24453513.
  14. 1 2 Ramasamy R, Chiba K, Butler P, Lamb DJ (June 2015). "Male biological clock: a critical analysis of advanced paternal age". Fertility and Sterility. 103 (6): 1402–6. doi: 10.1016/j.fertnstert.2015.03.011 . PMC   4955707 . PMID   25881878.
  15. 1 2 Abbas HA, Rafei RE, Charafeddine L, Yunis K (2015). "Effects of Advanced Paternal Age on Reproduction and Outcomes in Offspring". NeoReviews. 16 (2): e69–e83. doi:10.1542/neo.16-2-e69.
  16. 1 2 3 4 5 6 Sharma R, Agarwal A, Rohra VK, Assidi M, Abu-Elmagd M, Turki RF (April 2015). "Effects of increased paternal age on sperm quality, reproductive outcome and associated epigenetic risks to offspring". Reproductive Biology and Endocrinology. 13 (1): 35. doi: 10.1186/s12958-015-0028-x . PMC   4455614 . PMID   25928123.
  17. 1 2 3 Shah PS (February 2010). "Paternal factors and low birthweight, preterm, and small for gestational age births: a systematic review". American Journal of Obstetrics and Gynecology. 202 (2): 103–23. doi: 10.1016/j.ajog.2009.08.026 . PMID   20113689.
  18. Jaffe AE, Eaton WW, Straub RE, Marenco S, Weinberger DR (March 2014). "Paternal age, de novo mutations and schizophrenia". Molecular Psychiatry. 19 (3): 274–5. doi: 10.1038/mp.2013.76 . PMC   3929531 . PMID   23752248.
  19. Schulz CS, Green MF, Nelson KJ (1 April 2016). Schizophrenia and Psychotic Spectrum Disorders. Oxford University Press. ISBN   9780199378074 via Google Books.
  20. Torrey EF, Buka S, Cannon TD, Goldstein JM, Seidman LJ, Liu T, et al. (October 2009). "Paternal age as a risk factor for schizophrenia: how important is it?". Schizophrenia Research. 114 (1–3): 1–5. doi:10.1016/j.schres.2009.06.017. PMID   19683417. S2CID   36632150.
  21. Kolevzon A, Gross R, Reichenberg A (April 2007). "Prenatal and perinatal risk factors for autism: a review and integration of findings". Archives of Pediatrics & Adolescent Medicine. 161 (4): 326–33. doi: 10.1001/archpedi.161.4.326 . PMID   17404128.
  22. Frans EM, Sandin S, Reichenberg A, Lichtenstein P, Långström N, Hultman CM (September 2008). "Advancing paternal age and bipolar disorder". Archives of General Psychiatry. 65 (9): 1034–40. doi: 10.1001/archpsyc.65.9.1034 . PMID   18762589.
  23. Miller, Brian; Messias, Erick; Miettunen, Jouko; Alaräisänen, Antti; Järvelin, Marjo-Riita; Koponen, Hannu; Räsänen, Pirkko; Isohanni, Matti; Kirkpatrick, Brian (September 2011). "Meta-analysis of Paternal Age and Schizophrenia Risk in Male Versus Female Offspring". Schizophrenia Bulletin. 37 (5): 1039–1047. doi:10.1093/schbul/sbq011. PMC   3160220 . PMID   20185538.
  24. 1 2 3 Lopez-Castroman J, Gómez DD, Belloso JJ, Fernandez-Navarro P, Perez-Rodriguez MM, Villamor IB, et al. (February 2010). "Differences in maternal and paternal age between schizophrenia and other psychiatric disorders". Schizophrenia Research. 116 (2–3): 184–90. doi:10.1016/j.schres.2009.11.006. PMID   19945257. S2CID   20931376.
  25. de Kluiver H, Buizer-Voskamp JE, Dolan CV, Boomsma DI (April 2017). "Paternal age and psychiatric disorders: A review". American Journal of Medical Genetics. Part B, Neuropsychiatric Genetics. 174 (3): 202–213. doi: 10.1002/ajmg.b.32508 . PMC   5412832 . PMID   27770494.
  26. Xue F, Michels KB (December 2007). "Intrauterine factors and risk of breast cancer: a systematic review and meta-analysis of current evidence". The Lancet. Oncology. 8 (12): 1088–1100. doi:10.1016/S1470-2045(07)70377-7. PMID   18054879.
  27. Bishop DB, O'Connor PJ, Desai J (2010). "Diabetes". Chronic Disease Epidemiology and Control (3rd ed.). Washington, DC: American Public Health Association. p. 301. ISBN   9780875531922.
  28. Cardwell CR, Stene LC, Joner G, Bulsara MK, Cinek O, Rosenbauer J, et al. (February 2010). "Maternal age at birth and childhood type 1 diabetes: a pooled analysis of 30 observational studies". Diabetes. 59 (2): 486–94. doi: 10.2337/db09-1166 . PMC   2809958 . PMID   19875616.
  29. Stene LC, Harjutsalo V, Moltchanova E, Tuomilehto J (2011). "Epidemiology of Type 1 Diabetes". In Holt RIG, Cockram C, Flyvbjerg A, Goldstein BJ (eds.). Textbook of Diabetes. John Wiley & Sons. p. 39. ISBN   9781444348064.
  30. Girirajan S (April 2009). "Parental-age effects in Down syndrome". Journal of Genetics. 88 (1): 1–7. doi:10.1007/s12041-009-0001-6. PMID   19417538. S2CID   32292319.
  31. Dzurova D, Pikhart H (June 2005). "Down syndrome, paternal age and education: comparison of California and the Czech Republic". BMC Public Health. 5: 69. doi: 10.1186/1471-2458-5-69 . PMC   1166564 . PMID   15963229.
  32. 1 2 3 4 Malaspina D, Reichenberg A, Weiser M, Fennig S, Davidson M, Harlap S, et al. (June 2005). "Paternal age and intelligence: implications for age-related genomic changes in male germ cells". Psychiatric Genetics. 15 (2): 117–25. doi:10.1097/00041444-200506000-00008. PMID   15900226. S2CID   33387858.
  33. Saha S, Barnett AG, Foldi C, Burne TH, Eyles DW, Buka SL, McGrath JJ (March 2009). Brayne C (ed.). "Advanced paternal age is associated with impaired neurocognitive outcomes during infancy and childhood". PLOS Medicine. 6 (3): e40. doi: 10.1371/journal.pmed.1000040 . PMC   2653549 . PMID   19278291.
  34. Cannon M (March 2009). "Contrasting effects of maternal and paternal age on offspring intelligence: the clock ticks for men too". PLOS Medicine. 6 (3): e42. doi: 10.1371/journal.pmed.1000042 . PMC   2653550 . PMID   19278293.
  35. Edwards RD, Roff J (September 2010). "Negative effects of paternal age on children's neurocognitive outcomes can be explained by maternal education and number of siblings". PLOS ONE. 5 (9): e12157. Bibcode:2010PLoSO...512157E. doi: 10.1371/journal.pone.0012157 . PMC   2939033 . PMID   20856853.
  36. Whitley E, Deary IJ, Der G, Batty GD, Benzeval M (13 December 2012). "Paternal age in relation to offspring intelligence in the West of Scotland Twenty-07 prospective cohort study". PLOS ONE. 7 (12): e52112. Bibcode:2012PLoSO...752112W. doi: 10.1371/journal.pone.0052112 . PMC   3521707 . PMID   23272219.
  37. Myrskylä M, Silventoinen K, Tynelius P, Rasmussen F (April 2013). "Is later better or worse? Association of advanced parental age with offspring cognitive ability among half a million young Swedish men". American Journal of Epidemiology. 177 (7): 649–55. doi: 10.1093/aje/kws237 . PMID   23467498.
  38. Arslan RC, Penke L, Johnson W, Iacono WG, McGue M (25 February 2014). "The effect of paternal age on offspring intelligence and personality when controlling for paternal trait level". PLOS ONE. 9 (2): e90097. arXiv: 1309.4625 . Bibcode:2014PLoSO...990097A. doi: 10.1371/journal.pone.0090097 . PMC   3934965 . PMID   24587224.
  39. Zhu JL, Vestergaard M, Madsen KM, Olsen J (2008). "Paternal age and mortality in children". European Journal of Epidemiology. 23 (7): 443–7. doi:10.1007/s10654-008-9253-3. PMID   18437509. S2CID   2092996.
  40. "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)
  41. Tournaye 2009, p. 102
  42. Eisenberg DT, Hayes MG, Kuzawa CW (June 2012). "Delayed paternal age of reproduction in humans is associated with longer telomeres across two generations of descendants". Proceedings of the National Academy of Sciences of the United States of America. 109 (26): 10251–6. Bibcode:2012PNAS..10910251E. doi: 10.1073/pnas.1202092109 . PMC   3387085 . PMID   22689985.
  43. 1 2 3 4 Schmidt L, Sobotka T, Bentzen JG, Nyboe Andersen A (2012). "Demographic and medical consequences of the postponement of parenthood". Human Reproduction Update. 18 (1): 29–43. doi: 10.1093/humupd/dmr040 . PMID   21989171.
  44. Myrskylä M, Elo IT, Kohler IV, Martikainen P (October 2014). "The association between advanced maternal and paternal ages and increased adult mortality is explained by early parental loss". Social Science & Medicine. 119: 215–23. doi:10.1016/j.socscimed.2014.06.008. PMC   4436970 . PMID   24997641.
  45. D'Onofrio BM, Rickert ME, Frans E, Kuja-Halkola R, Almqvist C, Sjölander A, et al. (April 2014). "Paternal age at childbearing and offspring psychiatric and academic morbidity". JAMA Psychiatry. 71 (4): 432–8. doi: 10.1001/jamapsychiatry.2013.4525 . PMC   3976758 . PMID   24577047.
  46. 1 2 3 4 5 6 Malaspina D, Gilman C, Kranz TM (June 2015). "Paternal age and mental health of offspring". Fertility and Sterility. 103 (6): 1392–6. doi: 10.1016/j.fertnstert.2015.04.015 . PMC   4457665 . PMID   25956369.
  47. 1 2 Wiener-Megnazi Z, Auslender R, Dirnfeld M (January 2012). "Advanced paternal age and reproductive outcome". Asian Journal of Andrology. 14 (1): 69–76. doi:10.1038/aja.2011.69. PMC   3735149 . PMID   22157982.
  48. Broer L, Codd V, Nyholt DR, Deelen J, Mangino M, Willemsen G, et al. (October 2013). "Meta-analysis of telomere length in 19,713 subjects reveals high heritability, stronger maternal inheritance and a paternal age effect". European Journal of Human Genetics. 21 (10): 1163–8. doi:10.1038/ejhg.2012.303. PMC   3778341 . PMID   23321625.
  49. 1 2 3 Goriely A, McGrath JJ, Hultman CM, Wilkie AO, Malaspina D (June 2013). ""Selfish spermatogonial selection": a novel mechanism for the association between advanced paternal age and neurodevelopmental disorders". The American Journal of Psychiatry. 170 (6): 599–608. doi: 10.1176/appi.ajp.2013.12101352 . PMC   4001324 . PMID   23639989.
  50. Chan PTK, Robaire B. Advanced Paternal Age and Future Generations. Front Endocrinol (Lausanne). 2022 Jun 9;13:897101. doi: 10.3389/fendo.2022.897101. PMID: 35757433; PMCID: PMC9218097
  51. Gonzalez DC, Ory J, Blachman-Braun R, Nackeeran S, Best JC, Ramasamy R. Advanced Paternal Age and Sperm DNA Fragmentation: A Systematic Review. World J Mens Health. 2022 Jan;40(1):104-115. doi: 10.5534/wjmh.200195. Epub 2021 Apr 16. PMID: 33987998; PMCID: PMC8761235
  52. Perrin MC, Brown AS, Malaspina D (November 2007). "Aberrant epigenetic regulation could explain the relationship of paternal age to schizophrenia". Schizophrenia Bulletin. 33 (6): 1270–3. doi:10.1093/schbul/sbm093. PMC   2779878 . PMID   17712030.
  53. Yoshizaki, Kaichi; Koike, Tasuku; Kimura, Ryuichi; Kikkawa, Takako; Oki, Shinya; Koike, Kohei; Mochizuki, Kentaro; Inada, Hitoshi; Kobayashi, Hisato; Matsui, Yasuhisa; Kono, Tomohiro; Osumi, Noriko (15 February 2019). "Paternal age affects offspring's behavior possibly via an epigenetic mechanism recruiting a transcriptional repressor REST" (PDF). doi:10.1101/550095. S2CID   91611232.{{cite journal}}: Cite journal requires |journal= (help)
  54. 1 2 Kidd SA, Eskenazi B, Wyrobek AJ (February 2001). "Effects of male age on semen quality and fertility: a review of the literature". Fertility and Sterility. 75 (2): 237–48. doi: 10.1016/S0015-0282(00)01679-4 . PMID   11172821.
  55. Johnson SL, Dunleavy J, Gemmell NJ, Nakagawa S (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. PMID   25462195. S2CID   8864418.
  56. "Definition of Advanced paternal age" . Retrieved 13 June 2023.
  57. "The Internet Classics Archive | the Republic by Plato".
  58. Galton, D J (1 August 1998). "Greek theories on eugenics". Journal of Medical Ethics. 24 (4): 263–267. doi:10.1136/jme.24.4.263. ISSN   0306-6800. PMC   1377679 . PMID   9752630.
  59. 1 2 Crow, James F. (October 2000). "The origins, patterns and implications of human spontaneous mutation". Nature Reviews Genetics. 1 (1): 40–47. doi:10.1038/35049558. PMID   11262873. S2CID   22279735.
  60. Krooth RS (December 1953). "Comments on the estimation of the mutation rate for achondroplasia". American Journal of Human Genetics. 5 (4): 373–6. PMC   1716528 . PMID   13104383.
  61. Brinkmann B, Klintschar M, Neuhuber F, Hühne J, Rolf B (June 1998). "Mutation rate in human microsatellites: influence of the structure and length of the tandem repeat". American Journal of Human Genetics. 62 (6): 1408–15. doi:10.1086/301869. PMC   1377148 . PMID   9585597.
  62. 1 2 Bray I, Gunnell D, Davey Smith G (October 2006). "Advanced paternal age: how old is too old?". Journal of Epidemiology and Community Health. 60 (10): 851–3. doi:10.1136/jech.2005.045179. PMC   2566050 . PMID   16973530.
  63. Australian Bureau of Statistics (11 November 2009). "3301.0 - Births, Australia, 2008. Summary of findings. Births" . Retrieved 25 February 2010.
  64. 1 2 3 Kühnert B, Nieschlag E (2004). "Reproductive functions of the ageing male". Human Reproduction Update. 10 (4): 327–39. doi: 10.1093/humupd/dmh030 . PMID   15192059.
  65. 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): 1–104. Retrieved 25 February 2010.
  66. 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.
  67. "A father again at 96; North Carolinan's baby a sister to boy born two years ago" . The New York Times. Associated Press. 4 June 1936. p. 10. Retrieved 26 April 2019.
  68. Frans E, MacCabe JH, Reichenberg A (February 2015). "Advancing paternal age and psychiatric disorders". World Psychiatry. 14 (1): 91–3. doi:10.1002/wps.20190. PMC   4329902 . PMID   25655163.

Further reading