Reproduction

(redirected from Rutting season)
Also found in: Dictionary, Thesaurus, Medical, Encyclopedia.
Related to Rutting season: Mating season

Reproduction

A woman's right to determine whether she will give birth was not legally recognized until the 1960s and 1970s, when U.S. Supreme Court decisions established that right. Until that time, women in the United States were denied access to Birth Control and to legal abortions by state criminal laws. Since the 1970s, there has been ongoing controversy over legalized Abortion, with the Supreme Court allowing states to impose restrictions on obtaining the procedure. In addition, medical science has developed techniques of Artificial Insemination and in vitro fertilization that enable pregnancy. These advances, in turn, have created opportunities for Surrogate Motherhood, opening up even more legal issues dealing with reproductive rights. Because of the cultural importance placed on motherhood and the intersection of religious beliefs and public policy, the debate over reproductive rights has been contentious.

Historical Background

In the nineteenth century, the average size of the U.S. family declined dramatically. A white woman in 1800 gave birth to an average of seven children. By the end of the century, the average was three-and-a-half children. In part, the decline was caused by the dissemination of scientific information on birth control. Many of the nineteenth-century proponents of family planning were radical social reformers who offended church and community leaders with their graphic descriptions of human Reproduction.

Conservatives sought to curtail this information on birth control and abortion. The most prominent conservative watchdog was Anthony Comstock, a New York businessman who led a national reform effort against obscene materials. His work resulted in the federal Comstock Law of 1873, which criminalized the transmission and receipt of "obscene," "lewd," or "lascivious" publications through the U.S. mail. The law specified that materials designed, adapted, or intended "for preventing conception or producing abortion" were included in the list of banned items. Some states passed "little Comstock laws" that prohibited the use of contraceptives.

Until the second half of the nineteenth century, few states had criminal laws against abortion. Women in colonial times had used abortion to dispose of the offspring of rape or seduction. Abortion was not illegal under the Common Law as long as it was performed before "quickening," the period at about four months when the fetus begins to move in the womb.

State legislatures passed laws in the first half of the nineteenth century that adopted the quickening rule, and a few states allowed abortion after quickening to save the life of the mother. Abortions increased markedly in the 1850s and 1860s, especially among middle-class white women.

Religious leaders began to denounce abortion, but the American Medical Association (AMA) proved to be the most successful in ending legalized abortion. The AMA was formed in 1847, and the all-male professional group (women were not allowed to become doctors) made abortion law reform one of its top priorities. The AMA saw abortion reform as a way to increase its influence and to drive out unlicensed practitioners of abortion. By the 1880s, medical and religious leaders had convinced all-male state legislatures (women were not allowed to vote) to impose criminal penalties on persons performing abortions and, in some states, on the women who had abortions. The laws were based on the states' Police Power to regulate public health and safety. This had some justification because abortion procedures of the time were dangerous, subjecting women to sterility and, in many cases, death. In response, women turned to birth control and to illegal abortions. The legal restrictions on birth control and abortion that were created in the late nineteenth century were not be removed until the 1960s and 1970s.

Restricting Antiabortion Protests

The legalization of Abortion resulted in the creation of many groups opposed to the medical procedure. Some groups have sought to take away this reproductive right by Lobbying Congress and state legislatures, and others have picketed outside clinics that offer abortion services. In the 1990s, groups such as Operation Rescue sought to prevent abortions by organizing mass demonstrations outside clinics and blockading their entrances, as well as confronting and impeding women seeking to enter the clinics.

Clinics responded by obtaining court injunctions that restricted how close abortion protestors could get to clinic property. Abortion protestors claimed that these court orders violated their First Amendment rights of assembly and free speech.

The U.S. Supreme Court, in Schenck v. Pro-Choice Network of Western New York, 519 U.S. 357, 117 S. Ct. 855, 137 L. Ed. 2d 1 (1997), clarified what types of restrictions a judge could impose on abortion clinic protests. The Court upheld an Injunction provision that imposed a fixed buffer zone around the abortion clinic. In this case the buffer zone affected protests within 15 feet from either side or edge of, or in front of, doorways or doorway entrances, parking lot entrances, and driveways and driveway entrances. Chief Justice william h. rehnquist ruled that the government had an interest in ensuring public safety and order, promoting free flow of traffic, protecting property rights, and protecting a woman's freedom to seek pregnancy-related services.

The Court did strike down a provision concerning floating buffer zones. These zones, which prohibited demonstrations within 15 feet of any person or vehicle seeking access to or leaving abortion facilities, "burdened more speech than was necessary" to serve the government interests cited in support of fixed zones. Thus, protestors were free to approach persons outside the 15-foot fixed buffer zone.

In 2000, though, the Court again considered the issue of a buffer zone in Hill v. Colorado, 530 U.S. 703, 120 S. Ct. 2480, 147 L. Ed. 2d 597 (2000). The Court upheld Colorado's 1993 statute, which prevented anyone from counseling, distributing leaflets, or displaying signs within eight feet of others without their consent whenever they are within 100 feet of a health-clinic entrance. The Colorado law was enacted, according to attorneys for the state, after abortion patients complained of being spat on, kicked, and harassed outside clinics. Those who challenged the law claimed it was a violation of their Freedom of Speech under the First Amendment. The court found sufficient public and State Interest to uphold the restriction.

Further readings

Briant, Keith. 1962. Marie Stopes: A Biography. London: Hogarth.

Korn, Peter. 1996. A Year in the Life of an Abortion Clinic. New York: Grove/Atlantic.

Cross-references

Abortion; Roe v. Wade.

Birth Control

In the early twentieth century, a group of reformers sought to legally provide birth control information. The most prominent of these reformers was margaret sanger, who coined the term birth control. Sanger challenged state laws restricting birth control information, seeking to draw public support. Though the courts generally rebuffed her efforts, Sanger helped build a national movement. In 1921, she founded the American Birth Control League, which, in 1942, became the Planned Parenthood Federation of America.

Renewed legal challenges to restrictive state laws began in the 1950s. By 1960, almost every state had legalized birth control. Nevertheless, laws remained on the books that prevented the distribution of birth control information and contraceptives. A specific target was the 1879 Connecticut little Comstock law that made the sale and possession of birth control devices a misdemeanor. The law also prohibited anyone from assisting, abetting, or counseling another in the use of birth control devices.

The Supreme Court reviewed the Connecticut law in griswold v. state of connecticut, 381 U.S. 479, 85 S. Ct. 1678, 14 L. Ed. 2d 510 (1965). Estelle Griswold was the director of Planned Parenthood in Connecticut. Just three days after Planned Parenthood opened a clinic in New Haven, Griswold was arrested. She was convicted and fined $100. The Connecticut courts upheld her conviction, rejecting the contention that the state law was unconstitutional.

The Supreme Court struck down the Connecticut birth control law on a vote of 7 to 2. In his majority opinion, Justice william o. douglas announced that the law was unconstitutional because it violated an individual's right to privacy. Douglas asserted that "specific guarantees in the Bill of Rights have penumbras, formed by emanations from those guarantees that help give them life and substance. Various guarantees create zones of privacy." Thus, these "penumbras" (things on the fringe of a major region) and "emanations" added up to a general, independent right of privacy. In Douglas's view, this general right was infringed by the state of Connecticut when it outlawed birth control. He said that the state cannot be permitted "to search the sacred precincts of marital bedrooms for telltale signs of the use of contraceptives."

The Griswold decision invalidated the Connecticut law only insofar as it invaded marital privacy, leaving open the question of whether states could prohibit the use of birth control devices by unmarried persons. In Eisenstadt v. Baird, 405 U.S. 438, 92 S. Ct. 1029, 31 L. Ed. 2d 349 (1972), the Court reviewed a Massachusetts law that prohibited unmarried persons from obtaining and using contraceptives. William Baird was arrested after giving a lecture on birth control to a college group and providing contraceptive foam to a female student. The Court struck down the law, establishing that the right of privacy is an individual right, not a right enjoyed only by married couples. Justice william j. brennan jr., in his majority opinion, stated, "If the right of privacy means anything, it is the right of the individual, married or single, to be free from unwarranted governmental intrusion into matters so fundamentally affecting a person as the decision whether or not to beget a child."

With Griswold and Eisenstadt, state prohibition of birth control information and devices came to an end. These decisions also enabled schools to give more information to students concerning sex education. Some schools even dispense contraceptives.

In 1997, the Food and Drug Administration (FDA) approved the use of emergency contraceptive, known popularly as the "morning-after pill." Developed by Canadian professor Albert Yuzpe and known as the Yuzpe Regimen, the pill contains heavy dosages of hormones that can prevent pregnancy if taken 72 hours after sexual intercourse. Proponents, including pro-choice advocates in the abortion debate, claim that it is a safe and effective method of birth control. Pro-life advocates and others denounce the pill as a form of abortion. Some critics in the medical field also claimed that repeated use of the pill could have unknown long-term effects due to its high level of hormones.

Abortion

The establishment in Eisenstadt of an individual's right to privacy soon had dramatic implications for state laws that criminalized abortions. Until the 1960s, abortion was illegal in every state, except to save the mother's life. The growth of the modern feminist movement in the 1960s led to calls for the legalization of abortion, and many state legislatures began to amend their laws to permit abortion when the pregnancy resulted from a rape or when the child was likely to suffer from a serious birth defect. However, these laws generally required that a committee of doctors approve the abortion.

State legislation was swept away with the Supreme Court's controversial decision in Roe v. Wade, 410 U.S. 113, 93 S. Ct. 705, 35 L. Ed. 2d 147 (1973). A Class Action lawsuit challenged the state of Texas's abortion law. sarah weddington, the attorney for "Jane Roe," argued that the Constitution allows a woman to control her own body, including the decision to terminate an unwanted pregnancy.

The Supreme Court, on a 7–2 vote, struck down the Texas law. Justice harry a. black-mun, in his majority opinion, relied on the prior right to privacy decisions to justify the Court's action. Blackmun concluded that the right to privacy "is broad enough to encompass a woman's decision whether or not to terminate her pregnancy." More importantly, he stated that the right of privacy is a fundamental right. This meant that the state of Texas had to meet the Strict Scrutiny test of constitutional review. Texas showed a compelling state interest because it had a strong interest in protecting maternal health that justified reasonable state regulation of abortions performed after the first trimester (three months) of pregnancy. However, Texas also sought to proscribe all abortions and claimed a compelling State Interest in protecting unborn human life. Though the Court acknowledged that this was a legitimate interest, it held that it does not become compelling until that point in pregnancy when the fetus becomes "viable," capable of "meaningful life outside the mother's womb." Beyond the point of viability, the Court held that the state may prohibit abortion, except in cases in which it is necessary to preserve the life or health of the mother.

The Court rejected the argument that a fetus is a "person" as that term is used in the Constitution and thus possesses a right to life. To find a fetus to be a person would make any abortion a Homicide, which would prevent a state from allowing abortions in cases of rape or in which the pregnancy endangers the life of the mother.

The Roe decision elicited a hostile reaction from opponents of abortion. The creation of a "pro-life" movement that sought to overturn Roe was immediate, becoming a new fixture in U.S. politics. Pro-life forces sought a constitutional amendment to undo the decision, but it fell one vote short in the U.S. Senate in 1983. Over time, as the composition of the Supreme Court has changed, the Court has modified its views, without overturning Roe.

In the 1970s, a majority of the Court resisted efforts by some states to put restrictions on a woman's right to have an abortion. In Planned Parenthood of Central Missouri v. Danforth, 428 U.S. 52, 96 S. Ct. 2831, 49 L. Ed. 2d 788 (1976), the Court struck down a Missouri law that required minors to obtain the consent of their husbands or parents before obtaining an abortion. In 1979, in Bellotti v. Baird, 443 U.S. 622, 99 S. Ct. 3035, 61 L. Ed. 2d 797, the Court invalidated a similar Massachusetts law. Both opinions emphasized the personal nature of abortion decisions and the fact that the state cannot give someone else a Veto over the exercise of one's constitutional rights.

In Akron v. Akron Center for Reproductive Health, 462 U.S. 416, 103 S. Ct. 2481, 76 L. Ed. 2d 687 (1983), the Court struck down a city ordinance that required that all abortions be performed in hospitals; a twenty-four-hour waiting period must pass before an abortion could be performed; certain specified statements be made by a doctor to a woman seeking an abortion to ensure that she made a truly informed decision; and all fetal remains be disposed in a humane and sanitary manner. The Court held that these requirements imposed significant burdens on a woman's exercise of her constitutional right without substantially furthering the state's legitimate interests.

Opponents of abortion were successful, however, in preventing the payment of public funds for abortions not deemed medically necessary. In Maher v. Roe, 432 U.S. 464, 97 S. Ct. 2376, 53 L. Ed. 2d 484 (1977), the Court upheld a Connecticut state regulation that denied Medicaid benefits to indigent women seeking to have abortions, unless their physicians certified that their abortions were medically necessary. The Court found the law permissible because poor women were not a "suspect class" entitled to strict scrutiny review and because the regulation did not unduly burden the exercise of fundamental rights. In 1980, the Court upheld a provision of federal law, commonly known as the Hyde amendment, forbidding federal funds to support nontherapeutic abortions (Harris v. McRae, 448 U.S. 297, 100 S. Ct. 2671, 65 L. Ed. 2d 784).

During the 1980s and 1990s, the conservative majority on the Court showed more deference to state regulation of abortions. In Webster v. Reproductive Health Services, 492 U.S. 490, 109 S. Ct. 3040, 106 L. Ed. 2d 410 (1989), the Court upheld a Missouri law restricting abortions that contained the statement, "the life of each human being begins at conception." On a 5–4 vote, the Court upheld a law that forbids state employees from performing, assisting in, or counseling women to have abortions. It also prohibited the use of any state facilities for these purposes and required all doctors who would perform abortions to conduct viability tests on fetuses at or beyond 20 weeks' gestation.

In 1991, the Court upheld federal regulations imposed by the Reagan administration that barred birth control clinics that received federal funds from providing information about abortion services to their clients (Rust v. Sullivan, 500 U.S. 173, 111 S. Ct. 1759, 114 L. Ed. 2d 233). The Supreme Court found the regulation to be a legitimate condition imposed on the receipt of federal financial assistance.

The Court appeared to be ready to overturn the Roe precedent, but it surprised observers when it upheld Roe in Planned Parenthood v. Casey, 505 U.S. 833, 112 S. Ct. 2791, 120 L. Ed. 2d 674 (1992). The Pennsylvania law restricting abortions required spousal notification, parental consent in cases of minors, and a 24-hour waiting period before the abortion could be performed. Similar requirements had been struck down by the Court before.

On a 5–4 vote, the Court reaffirmed the essential holding of Roe that the constitutional right of privacy is broad enough to include a woman's decision to terminate her pregnancy. Though there was no majority opinion, the controlling opinion by Justice anthony m. kennedy, joined by Justices Sandra Day O'Connor and david h. souter, defended the reasoning of Roe and the line of cases that followed it. However, the joint opinion abandoned the trimester framework and declared a new "undue burden" test for judging regulations of abortion. Using this test, the joint opinion upheld the parental consent, waiting period, and record-keeping and reporting provisions, but invalidated the spousal notification requirement.

Pregnancy and Medical Developments

Artificial insemination, in vitro fertilization, and embryo transplants have created new opportunities for conceiving children. With artificial insemination, sperm from a donor is introduced into the vagina or through the cervix of a woman by any method other than sexual inter-course. Originally this technique was used when a husband was sterile or impotent, but it is now available to women regardless of whether they are married. For example, a lesbian couple could use artificial insemination to start a biological family.

The technique of in vitro fertilization gained international attention with the 1978 birth in England of Louise Brown, the first child conceived by in vitro fertilization. This technique involves the fertilization of the egg outside the womb. The embryo is then transferred to a woman's uterus.

Because sperm and eggs can be frozen and stored indefinitely, there are occasional legal disputes over the rights to these genetic materials when a Husband and Wife divorce. For example, in Kass v. Kass, 696 N.E. 2d 174 (N.Y. 1998), the New York Court of Appeals determined that the custody of five frozen embryos should be determined by the terms of a contract signed by a couple with a hospital that stored the embryos. The couple had sought to become pregnant through in virto fertilization, but, after several failed attempts, decided to divorce. The husband and wife initially agreed to the terms of a Consent Decree with the hospital whereby the hospital could retain the right to keep the embryos for research purposes. The wife later changed her mind and wanted custody of the embryos. The court held that the consent agreements constituted valid contracts and must be enforced. The court ruled that under the terms of the contract, the hospital should be awarded the embryos for use in research.Other courts have considered disputes whereby one spouse wishes to use embryos for the purpose of procreation while the other wants the embryos destroyed. Several state supreme courts have held that the right of a spouse who wishes to avoid procreation is superior to the wishes of spouse who wishes to procreate. In J. B. v. M. B., 783 A. 2d 707 (N.J. 2001), for example, the New Jersey Supreme Court determined that a husband's right to procreate was not disturbed by its ruling that remaining frozen embryos from the husband and wife be destroyed according to the wishes of the wife.

Developments in in vitro fertilization led to surrogate motherhood, which has caused legal battles as well. In these cases, a woman agrees to be either artificially inseminated by a spermdonor father or have a fertilized ovum inserted into her uterus. After giving birth, the surrogate mother legally surrenders the infant to the person or couple who will adopt and rear the child. The idea of surrogate motherhood is attractive to some couples because a child born of a surrogate mother will share half or all the genetic material of the parents who will raise the child.

One of the most publicized cases regarding surrogate motherhood is that of Baby M. In 1985, Mary Beth Whitehead agreed to be inseminated with the sperm of William Stern and, upon the birth of the child, relinquish her parental rights to Stern. But once the child was born, Whitehead found that she did not wish to give up the child, a girl who she named Sara. A court battle ensued, during which Stern, along with his wife, Elizabeth, were granted temporary custody of the child they had named Melissa. The court decided that Whitehead's parental rights were to be terminated, and Elizabeth Stern was granted the right to immediately adopt the child. The New Jersey Supreme Court overturned this verdict in part on February 2, 1988, restoring Whitehead's parental rights and invalidating Elizabeth Stern's Adoption, but granting William Stern custody of the infant.

Many surrogate mothers are close friends or relatives of the childless couple. However, the practice of commercial surrogate arrangements has increased greatly since the late 1980s. Many major cities have surrogate agencies, which are often run by doctors and lawyers who maintain lists of potential surrogate mothers and help match a woman with a couple wanting to have a baby. Commercial surrogate agencies typically charge a fee of $10,000 or more to make the arrangements, which is in addition to the surrogate mother's expenses and fees, which may range from $10,000 to $100,000.

Commercial surrogate arrangements are not legal in all states, and there is little case law on the subject. Some states declare surrogacy contracts null, void, and unenforceable because they are against public policy. Opponents of commercial surrogacy believe that such arrangements exploit the surrogate mother and turn children into a commodity. They also are concerned that if a child is born with a disability, the adoptive parents may decline to take the child. Finally, there is the issue of the surrogate mother who may not wish to surrender the child after birth.

Other medical developments have also stirred controversy. In 1997, scientists successfully cloned the first adult animal, leading to speculation that the process could be used to clone human beings. Scientists first successfully inserted DNA from one human cell into another human egg, but they do not expect successful human cloning to be possible for several years. The issue has caused heated debates focusing on the scientific, moral, and religious concerns over the possibility that an adult human could be cloned.

Reproductive Hazards in the Workplace

Legal disputes have arisen when employers have barred pregnant women and women of childbearing age from jobs that pose potential hazards to the fetus. The Supreme Court, in United Auto Workers v. Johnson Controls, 499 U.S. 187, 111 S. Ct. 1196, 113 L. Ed. 2d 158 (1991), ruled that a female employee cannot be excluded from jobs that expose her to health risks that may harm her fetus. The Court found that the exclusion of the women violated Title VII of the Civil Rights Act of 1964 (42 U.S.C.A. § 2000e et seq.) because the company policy only applied to fertile women, not fertile men. Justice Blackmun, in his majority opinion, noted that the policy singled out women on the basis of gender and childbearing capacity rather than on the basis of fertility alone. Concerns about the health of a child born to a worker at the plant were to be left "to the parents who conceive, bear, support, and raise them [the children] rather than to the employers who hire those parents."

Further readings

"Genetics, Reproduction, and the Law." 1999. Trial 35 (July).

Hollinger, Joan Heifetz. 1985. "From Coitus to Commerce: Legal and Social Consequences of Noncoital Reproduction." University of Michigan Journal of Law Reform 18 (summer).

"In re Baby M" (colloquy). 1988. Georgetown Law Journal 76 (June).

"In re Baby M" (symposium). 1988. Seton Hall Law Review 18 (fall).

Levy, Stephanie. 2001. "Whose Embryo Is It Anyway? Courts Wrestle with Issues of High-tech Reproduction." Trial 37 (December).

Rahman, Anika. 1995. "Toward Government Accountability for Women's Reproductive Rights." St. John's Law Review 69 (winter-spring).

Ridder, Stephanie, and Lisa Woll. 1989. "Transforming the Grounds: Autonomy and Reproductive Freedom." Yale Journal of Law and Feminism 2 (fall).

Robertson, John A. 1988. "Procreative Liberty And The State's Burden Of Proof In Regulating Noncoital Reproduction." Law, Medicine & Health Care 16 (spring-summer).

Cross-references

Abortion; Adoption; Fetal Rights; Fetal Tissue Research; Genetic Engineering; Husband and Wife; Penumbra; Sex Discrimination; Wattleton, Alyce Faye; Women's Rights.

West's Encyclopedia of American Law, edition 2. Copyright 2008 The Gale Group, Inc. All rights reserved.
References in periodicals archive ?
We pooled our group survey data into two physiological periods: non-rutting season (May-August) and rutting season (September-November) (Fedosenko 2000, Fedosenko & Blank 2005).
Belinda Barclay, Knowsley Safari's Animal Manager, said: "The fallow deer rutting season is an exciting time here at Knowsley Safari.
"Although camel bite wounds are small, they may penetrate deeply, causing serious injuries to the neck including the major vessels," they warned, adding that "care should be taken when approaching male camels during the rutting season".
Of course, going back to what we said about rutting season, deer collisions are most likely in November, when approximately 18% of all incidents occur, followed by October and December.
The rutting season is when the fallow deer males compete for the right to mate with the does by furiously clashing antlers to determine the strongest, most powerful buck in the herd; the deep-sounding belches are the first signs of the rut.
Solitary gazelles were noted most often during the birthing season in spring and the rutting season in early winter.
The only concern is a potential encounter between the giant animal -- boasting two metres (almost seven feet) of height and 3.5 metres (over 11 feet) of length, two pointed horns and unpredictable manners especially during the rutting season -- with its most dangerous predator, the human.
The rutting season is triggered by the shorter daylight hours and cooler weather.
Clancy clearly lays out the secrets for finding and taking down trophy quality whitetail bucks outside of their rutting season. Explaining the tactics and techniques needed for a successful pursuit of the quarry, Hunting Tough Bucks offers a wealth of invaluable insights and information that will enhance and improve the reader's hunting skills.
Mr Gritten said if a cull was to begin it would be at the start of the rutting season in August.
NEW signs have been erected on a Midlands road which is notorious for collisions with deer during the rutting season.
scraped their antlers against its trunk in rutting season.