USA flag icon

An official website of the United States government

Here’s how you know

Official websites use .gov A .gov website belongs to an official government organization in the United States.

Secure .gov websites use HTTPS A lock ( Lock Locked padlock icon ) or https:// means you’ve safely connected to the .gov website. Share sensitive information only on official, secure websites.

Anabolic Steroids and Other Appearance and Performance Enhancing Drugs (APEDs)

  • Anabolic-androgenic steroids are the best-studied class of appearance and performance enhancing drugs (APEDs). APEDs are used to improve appearance by building muscle mass or to enhance athletic performance.
  • Although anabolic steroids and other APEDs may directly and indirectly have effects on a person’s mood, they do not typically produce a euphoric high. However, people who use these substances may develop a substance use disorder, defined as continued use despite adverse consequences.
  • Anabolic steroids can cause severe, long-lasting, and in some cases, irreversible damage. They can lead to early heart attacks, strokes, liver tumors, kidney failure, and psychiatric problems. In addition, stopping steroid use can cause depression, often leading to resumption of use.

What are anabolic steroids and other appearance and performance enhancing drugs (APEDs)?

Anabolic-androgenic steroids , often shortened to "anabolic steroids," "steroids," or "androgens," 2,3  are the most widely misused APED. These are synthetic substances similar to the male sex hormone testosterone. They promote the growth of skeletal muscle (anabolic effects) and the development of male sexual characteristics (androgenic effects) in both males and females. 2

These compounds are sometimes used medically to treat delayed puberty and muscle loss due to disease 4  and to treat low levels of testosterone in men with an associated medical condition. 5  Anabolic androgenic steroids can also improve feelings of well-being and increase bone strength, but are not approved for these purposes. However, testosterone-supplementation therapy is an increasingly common treatment for mood and sexual performance problems associated with male aging, and it is controversially being prescribed even for younger men. 6

Note that in the context of this report, anabolic steroids refer only to the non-prescribed use (misuse) of testosterone and testosterone-like substances by athletes and non-athlete bodybuilders. This research report will not cover image enhancers, such as  dermal fillers ,  Botox , or the skin tanner,  melanotan . 7

Non-steroidal anabolics , include insulin, insulin-like growth hormone (IGF), and human growth hormone (HGH)—substances that are produced by the human body and are prescribed for legitimate medical uses but also sometimes misused for performance enhancement.

Ergo/thermogenics  are compounds used to decrease body fat or to promote leanness versus muscle mass in endurance athletes. 8  The three main categories of ergo/thermogenics are:

  • Xanthines : compounds that increase attention and wakefulness and suppress appetite. Examples are caffeine, the asthma drug theophylline, and theobromine—a substance found in chocolate, coffee, and tea. 9
  • Sympathomimetics : drugs that are similar in structure and action to epinephrine and norepinephrine—natural chemicals in the body that increase heart rate, constrict blood vessels, and raise blood pressure. An example is ephedrine, which is derived from the ephedra plant. Ephedrine/ephedra used to be included in dietary supplements that promoted weight loss, increased energy, and enhanced athletic performance. 10  In 2004, the FDA banned the U.S. sale of dietary supplements containing ephedrine/ephedra due to various possible health risks including cardiovascular and nervous system effects. 11
  • Thyroid hormones : substances that regulate metabolism by altering the function of the thyroid. 12  Cytomel is an example.

Nutritional/dietary supplements are substances purchased legally from nutritional stores or via the internet that are often taken in combination with other APEDs. Creatine, which boosts exercise capacity, is one common example.

In the United States, dietary supplements containing steroid precursors such as tetrahydrogestrinone (THG) and androstenedione previously could be purchased legally without a prescription. Athletes took steroid precursors in an effort to boost testosterone levels. Less is known about the side effects of steroid precursors, but if large quantities of these compounds substantially increase testosterone levels in the body, then they also are likely to produce the same side effects as anabolic steroids themselves. 13  The purchase of these supplements, with the notable exception of dehydroepiandrosterone (DHEA), became illegal after the passage of the Anabolic Steroid Control Act of 2004, which amended the Controlled Substances Act. 14

What is the scope of anabolic steroid use in the United States?

All data refer to the United States population.

It is difficult to estimate the prevalence of steroid misuse in the United States because many national surveys that ask about drug use do not include questions about steroids. However, data on steroid misuse among young students are available from the NIDA-supported  Monitoring the Future Survey .

How many young students use steroids?

In 2022, an estimated 0.8% of 8th graders, 0.5% of 10th graders, and 1.3% of 12th graders reported misusing steroids in the past 12 months. Source:  2022 Monitoring the Future Survey .

Why are anabolic steroids misused?

Anabolic steroids increase lean muscle mass when used in conjunction with weight training. The aim, for non-athlete weightlifters, is typically improvement of appearance. Steroid use is often associated with a form of male body dysmorphic disorder called muscle dysmorphia, a preoccupation with the perceived inadequate size of their muscles. 19

As a result, some users report taking anabolic steroids to increase confidence and because they feel that they are at a point where they can no longer get bigger through weight training alone. Most users report that anabolic steroids help them achieve their ideal body. 28

Increasing muscle mass may also promote strength, which can improve performance in certain types of sports. More benefit is seen for strength-dependent sports (weightlifting, shot-put throwing, football) than for sports that require speed, agility, flexibility, and/or endurance. 29

Anabolic steroid users also report that their muscles recover faster from intense strain and muscle injury. 30  Research in animals has not conclusively supported this belief, with some showing that anabolic steroids can enhance recovery from certain types of muscle damage, 31,32  but others finding no benefit in taking anabolic steroids to enhance muscle recovery. 33

Anabolic steroid users report using an average of about 11 APEDs per year. They are also more likely than non-steroid users to take supplements such as protein powders and creatine; estrogen blockers; ergo/thermogenics, such as caffeine or ephedrine; medications for erectile dysfunction; and other hormones such as insulin, thyroid hormones, and human growth hormone. 26

How are anabolic steroids used?

Commonly misused steroids.

  • Anadrol (oxymetholone)
  • Anavar (oxandrolone)
  • Dianabol (methandienone )
  • Winstrol (stanozolol)
  • Restandol (testosterone undecanoate)

Injectable Steroids

  • Deca-Durabolin (nandrolone decanoate)
  • Durabolin (nandrolone phenpropionate)
  • Depo-Testosterone (testosterone cypionate)
  • Agovirin (testosterone propionate)
  • Retandrol (testosterone phenylpropionate)
  • Equipoise (boldenone undecylenate) 29,34

Cycling, stacking, pyramiding, and plateauing

Steroids are often used in patterns called "cycling." This involves taking multiple doses of steroids over a specific period of time, stopping for a period, and starting again. People who misuse steroids also typically "stack" the drugs, meaning that they take two or more different anabolic steroids, mix oral and/or injectable types, and sometimes even take compounds that are designed for veterinary use. 37,38 The belief is that different steroids interact to produce an effect on muscle size that is greater than the effects of each drug individually, 36 a theory that has not been tested scientifically.

Another common mode of steroid misuse is referred to as "pyramiding," which typically involves taking them in a cycle of six to 12 weeks, tapering gradually rather than starting and finishing a cycle abruptly. At the beginning of a cycle, the person starts with low doses of the drugs being stacked and then slowly increases the doses. In the second half of the cycle, the doses are slowly decreased to zero. This is sometimes followed by a second cycle in which the person continues to train but without drugs. Steroid users believe that pyramiding allows the body time to adjust to the high doses, and the drug-free cycle allows the body's hormonal system time to recuperate. 2

A technique called "plateauing" may also be used, whereby steroids are staggered, overlapped, or substituted with another type of steroid to avoid developing tolerance. 36 As with stacking, the effects of pyramiding, cycling, and plateauing have not been substantiated scientifically.

What are the side effects of anabolic steroid misuse?

A variety of side effects can occur when anabolic steroids are misused, ranging from mild effects to ones that are harmful or even life-threatening. Most are reversible if the user stops taking the drugs. However, others may be permanent or semi-permanent.

Most data on the long-term effects of anabolic steroids in humans come from case reports rather than formal epidemiological studies. Serious and life-threatening adverse effects may be underreported, especially since they may occur many years later. One review found 19 deaths in published case reports related to anabolic steroid use between 1990 and 2012; however, many steroid users also used other drugs, making it difficult to show that the anabolic steroid use caused these deaths. 39 One animal study found that exposing male mice for one fifth of their lifespan to steroid doses comparable to those taken by human athletes caused a high frequency of early deaths. 40

Possible Health Consequences of Anabolic Steroid Misuse
 

Cardiovascular System

Steroid use has been associated with high blood pressure; 41 decreased function of the heart’s ventricles; 23,41,42 and cardiovascular diseases such as heart attacks, 43 artery damage, 44 and strokes, 45,46 even in athletes younger than 30. Steroids contribute to the development of cardiovascular disease partly by increasing the level of low-density lipoprotein (LDL) 47 and decreasing the level of high-density lipoprotein (HDL). 47,48 High LDL and low HDL levels increase the risk of atherosclerosis, a condition in which fatty substances are deposited inside arteries and disrupt blood flow. If blood is prevented from reaching the heart or brain, the result can be a heart attack or stroke, respectively. Steroids also increase the risk that blood clots will form in blood vessels, potentially disrupting blood flow and damaging the heart muscle, so that it does not pump blood effectively. 49

Hormonal System

Steroid use disrupts the normal production of hormones in the body. Changes that can be reversed include decreased sperm production, 56–59 decreased function of the testes (hypogonadism) that leads to low testosterone levels, 60 and shrinking of the testicles (testicular atrophy). 56,61 Irreversible changes include male-pattern baldness and breast development (gynecomastia) in men. 59,62 Anabolic steroids may also act upon the hormone system to increase the risk of testicular cancer, especially when steroids are used in combination with insulin-like growth factor. 63

In females, anabolic steroids cause masculinization. Specifically, breast size and body fat decrease, the skin becomes coarse, and the voice deepens. 64 Women may experience excessive growth of body hair but lose scalp hair. 65 With continued administration of steroids, some of these effects become irreversible. It is commonly believed that anabolic steroids will produce irreversible enlargement of the clitoris in females, although there are no studies on this. 66

Many people who inject anabolic steroids may use nonsterile injection techniques or share contaminated needles with other users. This puts these steroid users at risk for acquiring life threatening viral infections, such as HIV and hepatitis B and C. 76 In addition, animal models indicate that anabolic steroids suppress the immune system, 77 which could worsen infections.

Steroid misuse has been associated with liver damage, 50,51 tumors, 46,52,53 and a rare condition called peliosis hepatis, in which blood-filled cysts form in the liver. 54 The cysts can rupture, causing internal bleeding and even death in rare cases. 55

Musculoskeletal System

Rising levels of testosterone and other sex hormones normally trigger the growth spurt that occurs during puberty and adolescence. These rising levels of testosterone also provide the signals to stop growth. 67 When a child or adolescent takes anabolic steroids, the resulting artificially high sex hormone levels can prematurely signal the bones to stop growing. 68

Evidence suggests that weightlifters who misuse anabolic steroids have stiffer tendons, which could lead to an increased risk for tendon injury. 69

Steroid misuse can cause acne, 70–72 hair loss on the head, cysts, and oily hair and skin. 65 Users who inject steroids may also develop pain and abscess formation at injection sites. 73

Anabolic steroids can also produce jaundice, or yellowing of the skin or eyes, as a result of damage to the liver. 74,75

How does anabolic steroid misuse affect behavior?

Case reports and small studies indicate that anabolic steroids increase irritability and aggression, 75 although findings may be confounded by personality traits that are overrepresented in steroid users (i.e., antisocial, borderline, and histrionic personality disorder) 78 and use of other drugs. 79 People who misuse anabolic steroids report more anger than nonusers, 80 as well as more fights, verbal aggression, and violence toward their significant others, 81 sometimes called "roid rage." One study suggests that the mood and behavioral effects seen during anabolic-androgenic steroid misuse may result from secondary hormonal changes. 82

Scientists have attempted to test the association between anabolic steroids and aggression by administering high steroid doses or placebo for days or weeks to human volunteers and then assessing behavioral symptoms. In one such study, researchers found that testosterone over a six week period was associated with increased aggression, as assessed by a questionnaire and computer-based model of aggressive behavior. 83 In addition, high steroid doses produced greater feelings of irritability and aggression than placebo, 84 although the effects appear to be highly variable across individuals, 19 and other studies have not shown that effect. 85 One possible explanation, according to the researchers, is that some but not all anabolic steroids increase irritability and aggression.

Psychiatric Disorders

Anabolic steroid users are more likely than nonusers to report anxiety. 34,86 Moderate to high doses of anabolic steroids are also associated with major mood disorders such as mania, hypomania, 87 and major depression. 86,87 In one study, manic symptoms were not uniform across individuals, with most showing little psychological change, whereas a few demonstrated prominent effects. 19

Other Drug Use

Anabolic steroid users are more likely to use drugs such as marijuana, prescription opioids, cocaine, 88 or heroin. 86 In a study of men admitted to treatment for opioid use disorders, 25 percent reported prior use of anabolic steroids. Some described first learning about opioids from friends at the gym, and that they first purchased opioids from the same person who had sold them the anabolic steroids. 89 In a study of anabolic steroid users dependent upon the injectable opioid analgesic nalbuphine, most reported that they began using nalbuphine to treat pain from weightlifting injuries. They also described widespread use of nalbuphine in their gyms. 90

Research also indicates that some users might turn to other drugs to alleviate some of the negative effects of anabolic steroids. For example, a study of 227 men admitted in 1999 to a private treatment center for addiction to heroin or other opioids found that 9.3 percent had previously misused anabolic steroids. Of these, most reported using opioids to counteract insomnia, irritability, depression, and withdrawal from anabolic steroids. 91

What are the risks of anabolic steroid use in teens?

Unlike most illicit drug use, misuse of anabolic steroids most commonly begins in young adulthood rather than adolescence. But steroid use in teens is of concern, especially since the hormonal systems they interact with play a critical role in brain development during these years. 92–96  In adolescent rodents, exposure to anabolic steroids increased neuronal spine densities in the hippocampus and amygdala—brain regions involved in learning and emotions (e.g., aggression), respectively. Four weeks after withdrawal, these increases in neuronal spine densities returned to normal in the amygdala, but not in the hippocampus. This suggests that pubertal steroid exposure could produce long-lasting structural changes in certain brain regions. 97

Teens who use anabolic steroids may also be at increased risk for some cognitive side effects compared with adults. For example, males who begin using anabolic steroids during the teen years show increased impulsivity and decreased attention, compared to men who began using steroids in their adult years. 98  In adolescent rats, anabolic steroid exposure is associated with electrolytic imbalances, hyperactivity, anxiety, and increased sympathetic autonomic modulation (e.g., fight or flight response) during adulthood, even when steroid use was discontinued during adolescence. 99  In addition, adolescent male hamsters given anabolic steroids show increased aggression, even after steroid use is discontinued. These aggressive effects are paralleled by changes in levels of serotonin  100,101  and androgen receptors in the rodent brain. 102

How do anabolic steroids work in the brain?

Anabolic steroids act at androgen receptors to influence cellular functioning and gene expression. In addition to regulating pathways involved in the development of male characteristics, 103  activation of androgen receptors also produces rapid increases in calcium levels within skeletal muscle, heart, and brain cells. 104  Calcium plays important roles in neuronal signaling.

Research with human cells demonstrates that anabolic steroids also interact with certain types of GABA A  receptors, which could mediate the increased anxiety reported by steroid users. 105,106  In addition, animal studies show that anabolic steroids increase serotonin levels in brain regions involved in mood  107  and dopamine levels in reward-related brain regions. 107,108  Chronic use of anabolic steroids has also been shown to cause dysfunction of these reward pathways in animals. Specifically, rats given twice daily nandrolone injections for four weeks showed loss of sweet preference (a sign of reward dysfunction) that was accompanied by reductions of dopamine, serotonin, and noradrenaline in the nucleus accumbens, a reward-related brain region. 109

Are anabolic steroids addictive?

An undetermined percentage of steroid users may develop a steroid use disorder. Substance use disorders are defined by continued use despite adverse consequences; for steroid users, these may include physical or psychological problems such as breast growth (in men), sexual dysfunction, high blood pressure, excessive fats in the blood, heart disease, mood swings, severe irritability, or aggressiveness. Anabolic steroid users also may give up other important activities for fear that they will miss workouts, violate their dietary restrictions, or be prevented from using steroids. Steroid users also typically spend large amounts of time and money obtaining the drugs, and they may try to reduce or stop anabolic steroid use without success—possibly due to depression, anxiety about losing muscle mass, or and other unpleasant effects of withdrawal. 110

Withdrawal from steroids occurs when an individual develops dependence. A review of the research suggests that about 32 percent of people who misuse anabolic steroids become dependent. 23  Symptoms of dependence can include tolerance, which is needing to take more steroids to achieve the same effects. Another indicator of dependence is withdrawal once anabolic steroid use stops. 110  Withdrawal symptoms can include fatigue, restlessness, loss of appetite, insomnia, reduced sex drive, and steroid cravings. 111  The most dangerous of the withdrawal symptoms is depression, because it sometimes leads to suicide attempts. 112

How are anabolic steroids tested in athletes?

Although non-athlete weightlifters account for the bulk of anabolic steroid misuse, occasional steroid use by professional and Olympic athletes to improve performance or cheat in competition ("doping") has done the most to raise awareness of steroid misuse. The World Anti-Doping Agency (WADA) was founded in 1999 to consistently apply anti-doping policies across sports organizations and governments around the world. Non-compliant organizations can face sanctions such as event cancellation, loss of WADA funding, or ineligibility to host events. 114

This pie chart shows which sports had the greatest number of anti-doping rules violations in 2015. The categories, "other" "bodybuilding" "athletics" and "weightlifting" were at the top of the list.

Refinements in drug testing have improved the ability to detect anti-doping violations, resulting in increased numbers of reported violations over recent years. For example, the discovery of long-term steroid metabolites has lengthened the drug detection window, making it more difficult for athletes to pass drug tests by simply discontinuing steroid use just prior to an event. In addition, more sensitive technologies have allowed detection of lower metabolite thresholds. 115

Although testing procedures are now in place to deter steroid use among professional and Olympic athletes, new designer drugs constantly become available that can escape detection and put athletes willing to cheat one step ahead of testing efforts. 116–118 To detect early use of designer steroids and provide more accurate baseline standards for each athlete, testing laboratories store data from each drug testing sample. These samples are then used as reference points for future testing, thereby eliminating the possibility that a person tests positive simply because he or she has naturally elevated levels of testosterone when compared to the general population. 119 Long-term use of designer steroids suppresses levels endogenous steroids in urine samples, which could be the first indication that an athlete is taking a designer steroid. 117

Drug Testing and Nutritional Supplements

Athletes taking over-the-counter nutritional supplements may believe that such products are safe. However, nutritional supplements are not subjected to the same pre-approval requirements and quality tests as FDA-approved medications. 120 For example, some supplements advertised to promote weight loss have been found to contain banned stimulants such as ephedrine  121 or clenbuterol. 122 Other research shows that supplements sometimes contain prohormones or anabolic steroids. 123 In a study looking at 634 nutritional supplements from 13 different countries, 15 percent included some type of prohormone not listed on the label. 115 Another study showed that some non-labeled prohibited substances could be detected by drug tests up to 144 hours later. 124

Nutritional supplements sometimes contain banned substances that are not indicated in their labels. 115,124 The FDA notes that consumers should be wary if a product meets any of these criteria:

  • products claiming to be alternatives to FDA-approved drugs or to have effects similar to prescription drugs
  • products claiming to be a legal alternative to anabolic steroids
  • products that are marketed primarily in a foreign language or those that are marketed through mass e-mails
  • sexual enhancement products promising rapid effects such as working in minutes to hours, or long-lasting effects such as 24 hours to 72 hours
  • products that provide warnings about testing positive in performance enhancement drug tests  125

According to WADA’s codes, athletes are responsible for any prohibited substance found in their samples, regardless of whether ingestion was intentional or unintentional. However, sanctions may be reduced or avoided if the athlete can demonstrate that the substance was ingested through no significant fault or negligence on his/her part, or in some circumstances where the athlete did not intend to enhance performance. 126

What can be done to prevent steroid misuse?

Research suggests that high school athletes are less likely to use steroids if their peers and parents disapprove, indicating that peers and parents can be strong partners in prevention efforts. 127

However, research shows that simply teaching students about steroids' adverse effects does not convince adolescents that they will be adversely affected, nor does such instruction discourage young people from taking steroids in the future. Presenting both the risks and benefits of anabolic steroid use is more effective in convincing adolescents about steroids' negative effects, apparently because the students find a balanced approach more credible. 128

Research also indicates that some adolescents misuse steroids as part of a pattern of high-risk behaviors such as drinking and driving, carrying a gun, driving a motorcycle without a helmet, and using other illicit drugs. This suggests that a prevention program should focus on comprehensive high-risk behavior screening and counseling among teens who use anabolic steroids. 129

NIDA-Funded Prevention Research Helps Reduce Steroid Misuse

Studies show that one year after completion of the program, compared with a control group, ATLAS-trained football student athletes in 15 high schools had:

  • less use of anabolic steroids and less intention to misuse them in the future
  • less misuse of alcohol, amphetamines, and narcotics
  • less misuse of "athletic enhancing" supplements
  • less likelihood of engaging in hazardous behaviors such as drinking and driving
  • better knowledge about anabolic steroid, alcohol, and marijuana effects; better knowledge of alternatives to steroid misuse; greater confidence in athletic abilities; and improved nutritional behaviors  130

What treatments are effective for anabolic steroid misuse?

People who use steroids often do not seek treatment for their use, with one study reporting that 56 percent of users had never told their physician about their use. 133  This could be because users feel their physician lacks knowledge about anabolic steroids. 133  In addition, many internet sites devoted to anabolic steroids and other APEDs challenge the professionalism of health care providers and offer their own medically questionable advice on the use of APEDs. 134  This makes it important for health care providers to be educated on the signs and symptoms of steroid use in their patients. 111

Current views recommend that treatment for steroid use address the underlying causes of the steroid use. This can include:

  • psychological therapies (and possibly medications) for muscle dysmorphia
  • endocrine therapies to restore function in those suffering from hypogonadism and to alleviate symptoms of depression
  • antidepressants for those whose depression does not respond to endocrine therapies
  • pharmacological and psychosocial treatments for patients who are also dependent on opioids, which appear to also be effective in alleviating signs of anabolic steroid dependence  135

Find More Resources on Anabolic Steroids and Other Appearance and Performance Enhancing Drugs

  • Learn more about steroids and their legal status from the U.S. Drug Enforcement Administration
  • Review patient resources on anabolic steroids from MedlinePlus .
  • Explore publications about steroids from the U.S. Substance Abuse and Mental Health Services Administration.
  • Kanayama G, Pope HG. History and epidemiology of anabolic androgens in athletes and non-athletes.  Mol Cell Endocrinol . March 2017. doi:10.1016/j.mce.2017.02.039
  • Rashid H, Ormerod S, Day E. Anabolic androgenic steroids: what the psychiatrist needs to know.  Adv Psychiatr Treat . 2007;13(3):203-211.
  • Lipsett MB, Korenman SG. Androgen Metabolism.  JAMA . 1964;190(8):757-762. doi:10.1001/jama.1964.03070210063011
  • Shahidi NT. A review of the chemistry, biological action, and clinical applications of anabolic-androgenic steroids.  Clin Ther . 2001;23(9):1355-1390.
  • Testosterone Information.  https://www.fda.gov/drugs/drugsafety/ postmarketdrugsafetyinformationforpatientsandproviders/ucm161874.htm . Published March 3, 2015. Accessed May 26, 2017.
  • Rao PK, Boulet SL, Mehta A, et al. Trends in Testosterone Replacement Therapy Use from 2003 to 2013 among Reproductive-Age Men in the United States.  J Urol . 2017;197(4):1121-1126. doi:10.1016/j.juro.2016.10.063.
  • Brennan R, Wells JSG, Van Hout MC. The injecting use of image and performance-enhancing drugs (IPED) in the general population: a systematic review.  Health Soc Care Community . January 2016. doi:10.1111/hsc.12326.
  • Hildebrandt T, Langenbucher JW, Carr SJ, Sanjuan P. Modeling population heterogeneity in appearance- and performance-enhancing drug (APED) use: applications of mixture modeling in 400 regular APED users.  J Abnorm Psychol . 2007;116(4):717-733. doi:10.1037/0021-843X.116.4.71.
  • Gennaro MC, Abrigo C. Caffeine and theobromine in coffee, tea and cola-beverages.  Fresenius J Anal Chem . 1992;343(6):523-525.
  • Ephedra. NCCIH.  https://nccih.nih.gov/health/ephedra . Published November 9, 2011. Accessed November 6, 2017.
  • Office of Dietary Supplements. Ephedra.  https://ods.od.nih.gov/Health_Information/Ephedra.aspx . Published n.d. Accessed December 13, 2017.
  • American Thyroid Association. Thyroid and Weight FAQ. June 2012.  http://www.thyroid.org/thyroid-and-weight/ . Accessed November 6, 2017.
  • 108th Congress FS.  Regulation of Dietary Supplements: Hearing Before the Committee on Commerce, Science and Transportation .; 2003.  https://www.gpo.gov/fdsys/pkg/CHRG-108shrg20196/pdf/CHRG-108shrg20196.pdf . Accessed February 7, 2018.
  • 108th Congress.  Anabolic Steroid Control Act of 2004 . Vol S.2195.; 2004.  https://www.congress.gov/bill/108th-congress/senate-bill/2195/all-info . Accessed April 28, 2017.
  • Freeman ER, Bloom DA, McGuire EJ. A brief history of testosterone.  J Urol . 2001;165(2):371-373. doi:10.1097/00005392-200102000-00004.
  • Altschule MD, Tillotson KJ. The use of testosterone in the treatment of depressions.  N Engl J Med . 1948;239(27):1036-1038. doi:10.1056/NEJM194812302392704.
  • Wade N. Anabolic Steroids: Doctors Denounce Them, but Athletes Aren’t Listening.  Science . 1972;176(4042):1399-1403. doi:10.1126/science.176.4042.1399.
  • Buckley WE, Yesalis CE, Friedl KE, Anderson WA, Streit AL, Wright JE. Estimated prevalence of anabolic steroid use among male high school seniors.  JAMA . 1988;260(23):3441-3445.
  • Pope HG, Kouri EM, Hudson JI. Effects of supraphysiologic doses of testosterone on mood and aggression in normal men: a randomized controlled trial.  Arch Gen Psychiatry . 2000;57(2):133-140; discussion 155-156.
  • Significant Dates in U.S. Food and Drug Law History. 2014.  https://www.fda.gov/about-fda/fda-history/milestones-us-food-and-drug-law . Accessed June 4, 2016.
  • A Dangerous and Illegal Way to Seek Athletic Dominance and Better Appearance - A Guide for Understanding the Dangers of Anabolic Steroids. March 2004. https://www.deadiversion.usdoj.gov/pubs/brochures/steroids/public/. Accessed April 25, 2017.
  • Kanayama G, Boynes M, Hudson JI, Field AE, Pope HG. Anabolic steroid abuse among teenage girls: an illusory problem?  Drug Alcohol Depend . 2007;88(2-3):156-162. doi:10.1016/j.drugalcdep.2006.10.013.
  • Pope HG, Kanayama G, Athey A, Ryan E, Hudson JI, Baggish A. The lifetime prevalence of anabolic-androgenic steroid use and dependence in Americans: current best estimates.  Am J Addict Am Acad Psychiatr Alcohol Addict . 2014;23(4):371-377. doi:10.1111/j.1521-0391.2013.12118.x.
  • Irving LM, Wall M, Neumark-Sztainer D, Story M. Steroid use among adolescents: findings from Project EAT.  J Adolesc Health Off Publ Soc Adolesc Med . 2002;30(4):243-252.
  • Pope HG, Khalsa JH, Bhasin S. Body Image Disorders and Abuse of Anabolic-Androgenic Steroids Among Men.  JAMA . 2017;317(1):23-24. doi:10.1001/jama.2016.17441.
  • Ip EJ, Barnett MJ, Tenerowicz MJ, Perry PJ. The Anabolic 500 survey: characteristics of male users versus nonusers of anabolic-androgenic steroids for strength training.  Pharmacotherapy . 2011;31(8):757-766. doi:10.1592/phco.31.8.757.
  • Gruber AJ, Pope HG. Compulsive weight lifting and anabolic drug abuse among women rape victims.  Compr Psychiatry . 1999;40(4):273-277.
  • Wright S, Grogan S, Hunter G. Motivations for Anabolic Steroid use Among Bodybuilders.  J Health Psychol . 2000;5(4):566-571. doi:10.1177/135910530000500413.
  • American Academy of Pediatrics. Adolescents and anabolic steroids: a subject review. American Academy of Pediatrics. Committee on Sports Medicine and Fitness.  Pediatrics . 1997;99(6):904-908.
  • Bahrke MS, Wright JE, Strauss RH, Catlin DH. Psychological moods and subjectively perceived behavioral and somatic changes accompanying anabolic-androgenic steroid use.  Am J Sports Med . 1992;20(6):717-724.
  • Beiner JM, Jokl P, Cholewicki J, Panjabi MM. The effect of anabolic steroids and corticosteroids on healing of muscle contusion injury.  Am J Sports Med . 1999;27(1):2-9.
  • Ferry A, Noirez P, Page CL, Salah IB, Daegelen D, Rieu M. Effects of anabolic/androgenic steroids on regenerating skeletal muscles in the rat.  Acta Physiol Scand . 1999;166(2):105-110. doi:10.1046/j.1365-201x.1999.00549.x.
  • Ferry A, Vignaud A, Noirez P, Bertucci W. Respective effects of anabolic/androgenic steroids and physical exercise on isometric contractile properties of regenerating skeletal muscles in the rat.  Arch Physiol Biochem . 2000;108(3):257-261. doi:10.1076/1381345520000710831ZFT257.
  • Eklöf A-C, Thurelius A-M, Garle M, Rane A, Sjöqvist F. The anti-doping hot-line, a means to capture the abuse of doping agents in the Swedish society and a new service function in clinical pharmacology.  Eur J Clin Pharmacol . 2003;59(8-9):571-577. doi:10.1007/s00228-003-0633-z.
  • Medline Plus. Testosterone Topical. June 2016.  https://www.nlm.nih.gov/medlineplus/druginfo/meds/a605020.html#why .
  • Trenton AJ, Currier GW. Behavioural manifestations of anabolic steroid use.  CNS Drugs . 2005;19(7):571-595.
  • Evans NA. Gym and tonic: a profile of 100 male steroid users.  Br J Sports Med . 1997;31(1):54-58.
  • Wilson JD. Androgen abuse by athletes.  Endocr Rev . 1988;9(2):181-199. doi:10.1210/edrv-9-2-181.
  • Frati P, Busardò FP, Cipolloni L, Dominicis ED, Fineschi V. Anabolic Androgenic Steroid (AAS) related deaths: autoptic, histopathological and toxicological findings.  Curr Neuropharmacol . 2015;13(1):146-159. doi:10.2174/1570159X13666141210225414.
  • Bronson FH, Matherne CM. Exposure to anabolic-androgenic steroids shortens life span of male mice.  Med Sci Sports Exerc . 1997;29(5):615-619.
  • Urhausen A, Albers T, Kindermann W. Are the cardiac effects of anabolic steroid abuse in strength athletes reversible?  Heart Br Card Soc . 2004;90(5):496-501.
  • Kaskutas LA. Alcoholics anonymous effectiveness: faith meets science.  J Addict Dis . 2009;28(2):145-157. doi:10.1080/10550880902772464.
  • Vanberg P, Atar D. Androgenic anabolic steroid abuse and the cardiovascular system.  Handb Exp Pharmacol . 2010;(195):411-457. doi:10.1007/978-3-540-79088-4_18.
  • Baggish AL, Weiner RB, Kanayama G, et al. Cardiovascular Toxicity of Illicit Anabolic-Androgenic Steroid Use.  Circulation . 2017;135(21):1991-2002. doi:10.1161/CIRCULATIONAHA.116.026945.
  • El Scheich T, Weber A-A, Klee D, Schweiger D, Mayatepek E, Karenfort M. Adolescent ischemic stroke associated with anabolic steroid and cannabis abuse.  J Pediatr Endocrinol Metab JPEM . 2013;26(1-2):161-165. doi:10.1515/jpem-2012-0057.
  • Santamarina RD, Besocke AG, Romano LM, Ioli PL, Gonorazky SE. Ischemic stroke related to anabolic abuse.  Clin Neuropharmacol . 2008;31(2):80-85. doi:10.1097/WNF.0b013e3180ed4485.
  • Palatini P, Giada F, Garavelli G, et al. Cardiovascular effects of anabolic steroids in weight-trained subjects.  J Clin Pharmacol . 1996;36(12):1132-1140.
  • Bhasin S, Woodhouse L, Casaburi R, et al. Testosterone dose-response relationships in healthy young men.  Am J Physiol Endocrinol Metab . 2001;281(6):E1172-1181.
  • Linton MF, Yancey PG, Davies SS, Jerome WG (Jay), Linton EF, Vickers KC. The Role of Lipids and Lipoproteins in Atherosclerosis. In: De Groot LJ, Chrousos G, Dungan K, et al., eds.  Endotext . South Dartmouth (MA): MDText.com, Inc.; 2000.  https://www.ncbi.nlm.nih.gov/books/NBK343489/ . Accessed April 21, 2017.
  • Robles-Diaz M, Gonzalez-Jimenez A, Medina-Caliz I, et al. Distinct phenotype of hepatotoxicity associated with illicit use of anabolic androgenic steroids.  Aliment Pharmacol Ther . 2015;41(1):116-125. doi:10.1111/apt.13023.
  • Schwingel PA, Cotrim HP, Santos CR dos, et al. Recreational Anabolic-Androgenic Steroid Use Associated With Liver Injuries Among Brazilian Young Men.  Subst Use Misuse . 2015;50(11):1490-1498. doi:10.3109/10826084.2015.1018550.
  • Kosaka A, Takahashi H, Yajima Y, et al. Hepatocellular carcinoma associated with anabolic steroid therapy: report of a case and review of the Japanese literature.  J Gastroenterol . 1996;31(3):450-454.
  • Socas L, Zumbado M, Pérez-Luzardo O, et al. Hepatocellular adenomas associated with anabolic androgenic steroid abuse in bodybuilders: a report of two cases and a review of the literature.  Br J Sports Med . 2005;39(5):e27. doi:10.1136/bjsm.2004.013599.
  • Wakabayashi T, Onda H, Tada T, Iijima M, Itoh Y. High incidence of peliosis hepatis in autopsy cases of aplastic anemia with special reference to anabolic steroid therapy.  Acta Pathol Jpn . 1984;34(5):1079-1086.
  • Hansma P, Diaz FJ, Njiwaji C. Fatal Liver Cyst Rupture Due to Anabolic Steroid Use: A Case Presentation.  Am J Forensic Med Pathol . 2016;37(1):21-22. doi:10.1097/PAF.0000000000000218.
  • Bonetti A, Tirelli F, Catapano A, et al. Side effects of anabolic androgenic steroids abuse.  Int J Sports Med . 2008;29(8):679-687. doi:10.1055/s-2007-965808.
  • Liu PY, Swerdloff RS, Christenson PD, Handelsman DJ, Wang C, Hormonal Male Contraception Summit Group. Rate, extent, and modifiers of spermatogenic recovery after hormonal male contraception: an integrated analysis.  Lancet Lond Engl . 2006;367(9520):1412-1420. doi:10.1016/S0140-6736(06)68614-5.
  • Torres-Calleja J, González-Unzaga M, DeCelis-Carrillo R, Calzada-Sánchez L, Pedrón N. Effect of androgenic anabolic steroids on sperm quality and serum hormone levels in adult male bodybuilders.  Life Sci . 2001;68(15):1769-1774.
  • Calzada L, Torres-Calleja J, Martinez JM, Pedrón N. Measurement of androgen and estrogen receptors in breast tissue from subjects with anabolic steroid-dependent gynecomastia.  Life Sci . 2001;69(13):1465-1469.
  • Christou MA, Christou PA, Markozannes G, Tsatsoulis A, Mastorakos G, Tigas S. Effects of Anabolic Androgenic Steroids on the Reproductive System of Athletes and Recreational Users: A Systematic Review and Meta-Analysis.  Sports Med Auckl NZ . March 2017. doi:10.1007/s40279-017-0709-z.
  • Schürmeyer T, Knuth UA, Belkien L, Nieschlag E. Reversible azoospermia induced by the anabolic steroid 19-nortestosterone.  Lancet Lond Engl . 1984;1(8374):417-420.
  • Orlandi MA, Venegoni E, Pagani C. Gynecomastia in two young men with histories of prolonged use of anabolic androgenic steroids.  J Ultrasound . 2010;13(2):46-48. doi:10.1016/j.jus.2010.07.006.
  • Chimento A, Sirianni R, Zolea F, et al. Nandrolone and stanozolol induce Leydig cell tumor proliferation through an estrogen-dependent mechanism involving IGF-I system.  J Cell Physiol . 2012;227(5):2079-2088. doi:10.1002/jcp.22936.
  • Baker J. A report on alterations to the speaking and singing voices of four women following hormonal therapy with virilizing agents.  J Voice Off J Voice Found . 1999;13(4):496-507.
  • Scott MJ, Scott AM. Effects of anabolic-androgenic steroids on the pilosebaceous unit.  Cutis . 1992;50(2):113-116.
  • Nieschlag E, Vorona E. MECHANISMS IN ENDOCRINOLOGY: Medical consequences of doping with anabolic androgenic steroids: effects on reproductive functions.  Eur J Endocrinol Eur Fed Endocr Soc . 2015;173(2):R47-58. doi:10.1530/EJE-15-0080.
  • Zemel BS, Katz SH. The contribution of adrenal and gonadal androgens to the growth in height of adolescent males.  Am J Phys Anthropol . 1986;71(4):459-466. doi:10.1002/ajpa.1330710409.
  • Bierich JR. Effects and side effects of anabolic steroids in children.  Acta Endocrinol Suppl (Copenh) . 1961;39(Suppl 63):89-110.
  • Seynnes OR, Kamandulis S, Kairaitis R, et al. Effect of androgenic-anabolic steroids and heavy strength training on patellar tendon morphological and mechanical properties.  J Appl Physiol Bethesda Md 1985 . 2013;115(1):84-89. doi:10.1152/japplphysiol.01417.2012.
  • Kraus SL, Emmert S, Schön MP, Haenssle HA. The dark side of beauty: acne fulminans induced by anabolic steroids in a male bodybuilder.  Arch Dermatol . 2012;148(10):1210-1212. doi:10.1001/archdermatol.2012.855.
  • Melnik B, Jansen T, Grabbe S. Abuse of anabolic-androgenic steroids and bodybuilding acne: an underestimated health problem.  J Dtsch Dermatol Ges J Ger Soc Dermatol JDDG . 2007;5(2):110-117. doi:10.1111/j.1610-0387.2007.06176.x.
  • Voelcker V, Sticherling M, Bauerschmitz J. Severe ulcerated “bodybuilding acne” caused by anabolic steroid use and exacerbated by isotretinoin.  Int Wound J . 2010;7(3):199-201. doi:10.1111/j.1742-481X.2010.00676.x.
  • Rich JD, Dickinson BP, Flanigan TP, Valone SE. Abscess related to anabolic-androgenic steroid injection.  Med Sci Sports Exerc . 1999;31(2):207-209.
  • Cabb E, Baltar S, Powers DW, Mohan K, Martinez A, Pitts E. The Diagnosis and Manifestations of Liver Injury Secondary to Off-Label Androgenic Anabolic Steroid Use.  Case Rep Gastroenterol . 2016;10(2):499-505. doi:10.1159/000448883.
  • Yoshida EM, Erb SR, Scudamore CH, Owen DA. Severe cholestasis and jaundice secondary to an esterified testosterone, a non-C17 alkylated anabolic steroid.  J Clin Gastroenterol . 1994;18(3):268-270.
  • Ip EJ, Yadao MA, Shah BM, Lau B. Infectious disease, injection practices, and risky sexual behavior among anabolic steroid users.  AIDS Care . 2016;28(3):294-299. doi:10.1080/09540121.2015.1090539.
  • Hughes TK, Fulep E, Juelich T, Smith EM, Stanton GJ. Modulation of immune responses by anabolic androgenic steroids.  Int J Immunopharmacol . 1995;17(11):857-863.
  • Perry PJ, Kutscher EC, Lund BC, Yates WR, Holman TL, Demers L. Measures of aggression and mood changes in male weightlifters with and without androgenic anabolic steroid use.  J Forensic Sci . 2003;48(3):646-651.
  • Lundholm L, Frisell T, Lichtenstein P, Långström N. Anabolic androgenic steroids and violent offending: confounding by polysubstance abuse among 10,365 general population men.  Addict Abingdon Engl . 2015;110(1):100-108. doi:10.1111/add.12715.
  • Burnett KF, Kleiman ME. Psychological characteristics of adolescent steroid users.  Adolescence . 1994;29(113):81-89.
  • Choi PY, Pope HG. Violence toward women and illicit androgenic-anabolic steroid use.  Ann Clin Psychiatry Off J Am Acad Clin Psychiatr . 1994;6(1):21-25.
  • Daly RC, Su T-P, Schmidt PJ, Pagliaro M, Pickar D, Rubinow DR. Neuroendocrine and behavioral effects of high-dose anabolic steroid administration in male normal volunteers.  Psychoneuroendocrinology . 2003;28(3):317-331.
  • Kouri EM, Lukas SE, Pope HG, Oliva PS. Increased aggressive responding in male volunteers following the administration of gradually increasing doses of testosterone cypionate.  Drug Alcohol Depend . 1995;40(1):73-79.
  • Bahrke MS, Yesalis CE, Wright JE. Psychological and behavioural effects of endogenous testosterone and anabolic-androgenic steroids. An update.  Sports Med Auckl NZ . 1996;22(6):367-390.
  • Tricker R, Casaburi R, Storer TW, et al. The effects of supraphysiological doses of testosterone on angry behavior in healthy eugonadal men--a clinical research center study.  J Clin Endocrinol Metab . 1996;81(10):3754-3758. doi:10.1210/jcem.81.10.8855834.
  • Ip EJ, Lu DH, Barnett MJ, Tenerowicz MJ, Vo JC, Perry PJ. Psychological and physical impact of anabolic-androgenic steroid dependence.  Pharmacotherapy . 2012;32(10):910-919. doi:10.1002/j.1875-9114.2012.01123.
  • Pope HG, Katz DL. Psychiatric and medical effects of anabolic-androgenic steroid use. A controlled study of 160 athletes.  Arch Gen Psychiatry . 1994;51(5):375-382.
  • Kanayama G, Pope HG. Illicit use of androgens and other hormones: recent advances.  Curr Opin Endocrinol Diabetes Obes . 2012;19(3):211-219. doi:10.1097/MED.0b013e3283524008.
  • Kanayama G, Cohane GH, Weiss RD, Pope HG. Past anabolic-androgenic steroid use among men admitted for substance abuse treatment: an underrecognized problem?  J Clin Psychiatry . 2003;64(2):156-160.
  • Wines JD, Gruber AJ, Pope HG, Lukas SE. Nalbuphine hydrochloride dependence in anabolic steroid users.  Am J Addict . 1999;8(2):161-164.
  • Arvary D, Pope HG. Anabolic-androgenic steroids as a gateway to opioid dependence.  N Engl J Med . 2000;342(20):1532. doi:10.1056/NEJM200005183422018.
  • Keenan BS, Richards GE, Ponder SW, Dallas JS, Nagamani M, Smith ER. Androgen-stimulated pubertal growth: the effects of testosterone and dihydrotestosterone on growth hormone and insulin-like growth factor-I in the treatment of short stature and delayed puberty.  J Clin Endocrinol Metab . 1993;76(4):996-1001. doi:10.1210/jcem.76.4.8473416.
  • Morris JA, Jordan CL, Breedlove SM. Sexual differentiation of the vertebrate nervous system.  Nat Neurosci . 2004;7(10):1034-1039. doi:10.1038/nn1325.
  • Romeo RD, Richardson HN, Sisk CL. Puberty and the maturation of the male brain and sexual behavior: recasting a behavioral potential.  Neurosci Biobehav Rev . 2002;26(3):381-391.
  • Schulz KM, Molenda-Figueira HA, Sisk CL. Back to the future: The organizational-activational hypothesis adapted to puberty and adolescence.  Horm Behav . 2009;55(5):597-604. doi:10.1016/j.yhbeh.2009.03.010.
  • Zehr JL, Nichols LR, Schulz KM, Sisk CL. Adolescent development of neuron structure in dentate gyrus granule cells of male Syrian hamsters.  Dev Neurobiol . 2008;68(14):1517-1526. doi:10.1002/dneu.20675.
  • Cunningham RL, Claiborne BJ, McGinnis MY. Pubertal exposure to anabolic androgenic steroids increases spine densities on neurons in the limbic system of male rats.  Neuroscience . 2007;150(3):609-615. doi:10.1016/j.neuroscience.2007.09.038.
  • Hildebrandt T, Langenbucher JW, Flores A, Harty S, Berlin HA, Berlin H. The influence of age of onset and acute anabolic steroid exposure on cognitive performance, impulsivity, and aggression in men.  Psychol Addict Behav J Soc Psychol Addict Behav . 2014;28(4):1096-1104. doi:10.1037/a0036482.
  • Olivares EL, Silveira ALB, Fonseca FV, et al. Administration of an anabolic steroid during the adolescent phase changes the behavior, cardiac autonomic balance and fluid intake in male adult rats.  Physiol Behav . 2014;126:15-24. doi:10.1016/j.physbeh.2013.12.006.
  • Grimes JM, Melloni RH. Prolonged alterations in the serotonin neural system following the cessation of adolescent anabolic-androgenic steroid exposure in hamsters (Mesocricetus auratus).  Behav Neurosci . 2006;120(6):1242-1251. doi:10.1037/0735-7044.120.6.1242.
  • Ricci LA, Rasakham K, Grimes JM, Melloni RH. Serotonin-1A receptor activity and expression modulate adolescent anabolic/androgenic steroid-induced aggression in hamsters.  Pharmacol Biochem Behav . 2006;85(1):1-11. doi:10.1016/j.pbb.2006.06.022.
  • Menard CS, Harlan RE. Up-regulation of androgen receptor immunoreactivity in the rat brain by androgenic-anabolic steroids.  Brain Res . 1993;622(1-2):226-236.
  • Matsumoto T, Sakari M, Okada M, et al. The androgen receptor in health and disease.  Annu Rev Physiol . 2013;75:201-224. doi:10.1146/annurev-physiol-030212-183656.
  • Vicencio JM, Estrada M, Galvis D, et al. Anabolic androgenic steroids and intracellular calcium signaling: a mini review on mechanisms and physiological implications.  Mini Rev Med Chem . 2011;11(5):390-398.
  • Yang P, Jones BL, Henderson LP. Mechanisms of anabolic androgenic steroid modulation of alpha(1)beta(3)gamma(2L) GABA(A) receptors.  Neuropharmacology . 2002;43(4):619-633.
  • Yang P, Jones BL, Henderson LP. Role of the alpha subunit in the modulation of GABA(A) receptors by anabolic androgenic steroids.  Neuropharmacology . 2005;49(3):300-316. doi:10.1016/j.neuropharm.2005.03.017.
  • Kindlundh AM, Lindblom J, Bergström L, Wikberg JE, Nyberg F. The anabolic-androgenic steroid nandrolone decanoate affects the density of dopamine receptors in the male rat brain.  Eur J Neurosci . 2001;13(2):291-296.
  • Thiblin I, Finn A, Ross SB, Stenfors C. Increased dopaminergic and 5-hydroxytryptaminergic activities in male rat brain following long-term treatment with anabolic androgenic steroids.  Br J Pharmacol . 1999;126(6):1301-1306. doi:10.1038/sj.bjp.0702412.
  • Zotti M, Tucci P, Colaianna M, et al. Chronic nandrolone administration induces dysfunction of the reward pathway in rats.  Steroids . 2014;79:7-13.
  • Kanayama G, Brower KJ, Wood RI, Hudson JI, Pope HG. Issues for DSM-V: clarifying the diagnostic criteria for anabolic-androgenic steroid dependence.  Am J Psychiatry . 2009;166(6):642-645. doi:10.1176/appi.ajp.2009.08111699.
  • Brower KJ, Blow FC, Young JP, Hill EM. Symptoms and correlates of anabolic-androgenic steroid dependence.  Br J Addict . 1991;86(6):759-768.
  • Malone DA, Dimeff RJ, Lombardo JA, Sample RH. Psychiatric effects and psychoactive substance use in anabolic-androgenic steroid users.  Clin J Sport Med Off J Can Acad Sport Med . 1995;5(1):25-31.
  • 2015 Anti-Doping Rule Violations (ADRVs) Report .; 2017.  https://www.wada-ama.org/sites/default/files/resources/files/2015_adrvs_report_web_release_0.pdf .
  • Compliance Monitoring.  https://www.wada-ama.org/en/what-we-do/compliance-monitoring/compliance-monitoring-program . Published 2017. Accessed May 18, 2017.
  • Geyer H, Schänzer W, Thevis M. Anabolic agents: recent strategies for their detection and protection from inadvertent doping.  Br J Sports Med . 2014;48(10):820-826. doi:10.1136/bjsports-2014-093526.
  • Catlin DH, Sekera MH, Ahrens BD, Starcevic B, Chang Y-C, Hatton CK. Tetrahydrogestrinone: discovery, synthesis, and detection in urine.  Rapid Commun Mass Spectrom RCM . 2004;18(12):1245-1049. doi:10.1002/rcm.1495.
  • Catlin DH, Ahrens BD, Kucherova Y. Detection of norbolethone, an anabolic steroid never marketed, in athletes’ urine.  Rapid Commun Mass Spectrom RCM . 2002;16(13):1273-1275. doi:10.1002/rcm.722.
  • Sekera MH, Ahrens BD, Chang Y-C, Starcevic B, Georgakopoulos C, Catlin DH. Another designer steroid: discovery, synthesis, and detection of “madol” in urine.  Rapid Commun Mass Spectrom RCM . 2005;19(6):781-784. doi:10.1002/rcm.1858.
  • Mareck U, Geyer H, Opfermann G, Thevis M, Schänzer W. Factors influencing the steroid profile in doping control analysis.  J Mass Spectrom JMS . 2008;43(7):877-891. doi:10.1002/jms.1457.
  • Yonamine M, Garcia PR, de Moraes Moreau RL. Non-intentional doping in sports.  Sports Med Auckl NZ . 2004;34(11):697-704.
  • Jung J, Hermanns-Clausen M, Weinmann W. Anorectic sibutramine detected in a Chinese herbal drug for weight loss.  Forensic Sci Int . 2006;161(2-3):221-222. doi:10.1016/j.forsciint.2006.02.052.
  • Parr MK, Koehler K, Geyer H, Guddat S, Schänzer W. Clenbuterol marketed as dietary supplement.  Biomed Chromatogr BMC . 2008;22(3):298-300. doi:10.1002/bmc.928.
  • Geyer H, Braun H, Burke LM, Stear SJ, Castell LM. A-Z of nutritional supplements: dietary supplements, sports nutrition foods and ergogenic aids for health and performance--Part 22.  Br J Sports Med . 2011;45(9):752-754. doi:10.1136/bjsports-2011-090180.
  • De Cock KJ, Delbeke FT, Van Eenoo P, Desmet N, Roels K, De Backer P. Detection and determination of anabolic steroids in nutritional supplements.  J Pharm Biomed Anal . 2001;25(5-6):843-852.
  • Tainted Products Marketed as Dietary Supplements.  https://www.accessdata.fda.gov/scripts/sda/sdNavigation.cfm?sd=tainted_supplements_cder . Accessed January 9, 2018.
  • World Anti-Doping Code. 2009.  https://www.wada-ama.org/sites/default/files/resources/files/wada_anti-doping_code_2009_en_0.pdf . Accessed May 23, 2017.
  • Elliot D, Goldberg L. Intervention and prevention of steroid use in adolescents.  Am J Sports Med . 1996;24(6 Suppl):S46-47.
  • Goldberg L, Bents R, Bosworth E, Trevisan L, Elliot DL. Anabolic steroid education and adolescents: do scare tactics work?  Pediatrics . 1991;87(3):283-286.
  • Middleman AB, Faulkner AH, Woods ER, Emans SJ, DuRant RH. High-risk behaviors among high school students in Massachusetts who use anabolic steroids.  Pediatrics . 1995;96(2 Pt 1):268-272.
  • Goldberg L, MacKinnon DP, Elliot DL, Moe EL, Clarke G, Cheong J. The adolescents training and learning to avoid steroids program: preventing drug use and promoting health behaviors.  Arch Pediatr Adolesc Med . 2000;154(4):332-338.
  • Elliot DL, Goldberg L, Moe EL, Defrancesco CA, Durham MB, Hix-Small H. Preventing substance use and disordered eating: initial outcomes of the ATHENA (athletes targeting healthy exercise and nutrition alternatives) program.  Arch Pediatr Adolesc Med . 2004;158(11):1043-1049. doi:10.1001/archpedi.158.11.1043.
  • Elliot DL, Goldberg L, Moe EL, et al. Long-term Outcomes of the ATHENA (Athletes Targeting Healthy Exercise & Nutrition Alternatives) Program for Female High School Athletes.  J Alcohol Drug Educ . 2008;52(2):73-92.
  • Pope HG, Kanayama G, Ionescu-Pioggia M, Hudson JI. Anabolic steroid users’ attitudes towards physicians.  Addict Abingdon Engl . 2004;99(9):1189-1194. doi:10.1111/j.1360-0443.2004.00781.x.
  • Brennan BP, Kanayama G, Pope HG. Performance-enhancing drugs on the web: a growing public-health issue.  Am J Addict Am Acad Psychiatr Alcohol Addict . 2013;22(2):158-161. doi:10.1111/j.1521-0391.2013.00311.x.
  • Kanayama G, Brower KJ, Wood RI, Hudson JI, Pope HG. Treatment of anabolic-androgenic steroid dependence: Emerging evidence and its implications.  Drug Alcohol Depend . 2010;109(1-3):6-13. doi:10.1016/j.drugalcdep.2010.01.011

research articles about steroids

  • Subscribe to journal Subscribe
  • Get new issue alerts Get alerts

Secondary Logo

Journal logo.

Colleague's E-mail is Invalid

Your message has been successfully sent to your colleague.

Save my selection

Anabolic-Androgenic Steroid Use in Sports, Health, and Society

BHASIN, SHALENDER; HATFIELD, DISA L.; HOFFMAN, JAY R.; KRAEMER, WILLIAM J.; LABOTZ, MICHELE; PHILLIPS, STUART M.; RATAMESS, NICHOLAS A.

1 Department of Medicine, Brigham and Women’s Hospital, Boston, MA

2 Department of Kinesiology, University of Rhode Island, Kingston, RI

3 Department of Physical Therapy, Ariel University, Ariel, Israel

4 Department of Human Sciences, The Ohio State University, Columbus, OH

5 InterMed, P.A., South Portland, ME

6 Department of Pediatrics, Tufts University School of Medicine, Boston, MA

7 Department of Kinesiology, McMaster University, Hamilton, ON

8 Department of Health and Exercise Science, The College of New Jersey, Ewing, NJ

Address for correspondence: Stuart M. Phillips, Ph.D., F.A.C.S.M., Department of Kinesiology, McMaster University Ivor Wynne Centre 1280 Main St, West Hamilton, Ontario, Canada L8S 4K1; E-mail: [email protected] .

Supplemental digital content is available for this article. Direct URL citations appear in the printed text and are provided in the HTML and PDF versions of this article on the journal’s Web site ( www.acsm-msse.org ).

This consensus statement is an update of the 1987 American College of Sports Medicine (ACSM) position stand on the use of anabolic-androgenic steroids (AAS). Substantial data have been collected since the previous position stand, and AAS use patterns have changed significantly. The ACSM acknowledges that lawful and ethical therapeutic use of AAS is now an accepted mainstream treatment for several clinical disorders; however, there is increased recognition that AAS are commonly used illicitly to enhance performance and appearance in several segments of the population, including competitive athletes. The illicit use of AAS by competitive athletes is contrary to the rules and ethics of many sport governing bodies. Thus, the ACSM deplores the illicit use of AAS for athletic and recreational purposes. This consensus statement provides a brief history of AAS use, an update on the science of how we now understand AAS to be working metabolically/biochemically, potential side effects, the prevalence of use among athletes, and the use of AAS in clinical scenarios.

 This consensus statement is an update of the previous position stand from the American College of Sports Medicine (ACSM), published in 1987 ( 1 ). Since then, a substantial amount of scientific data on anabolic-androgenic steroids (AAS) has emerged and the circumstances of AAS use has evolved in the athletic, recreational, and clinical communities. The objective of this consensus statement is to provide readers with a brief summary of the current evidence and extend the recommendations provided in the 1987 document ( 1 ). Key topics discussed are the brief history of AAS, epidemiology, methods, and patterns of AAS use, androgen physiology and ergogenic effects, side effects of AAS, and clinical uses of AAS (see Box 1). The writing group used the rating system of the National Heart Lung and Blood Institute ( Table 1 ) and a consensus approach to synthesize the available evidence from clinical trials and case reports, narrative and systematic reviews, and meta-analyses ( 3 ). The recommendations represent the consensus of the writing panel, the ACSM, and incorporate guidance from other professional organizations with expertise in the area.

Evidence Category Sources of Evidence Definition
A RCT (rich body of data) Evidence is from endpoints of well-designed RCT (or trials that depart only minimally from randomization) that provide a consistent pattern of findings in the population for which the recommendation is made. Category A therefore requires substantial numbers of studies involving substantial numbers of participants.
B RCT (limited body of data) Evidence is from endpoints of intervention studies that include only a limited number of RCT, or subgroup analysis of RCT, or meta-analysis of RCT. In general, Category B pertains when few randomized trials exist, they are small in size, and the trial results are somewhat inconsistent, or the trials were undertaken in a population that differs from the target population of the recommendation.
C Non-RCT and observational studies Evidence is from outcomes of uncontrolled or nonrandomized trials or from observational studies.
D Panel consensus judgment Expert judgment is based on the panel’s synthesis of evidence from experimental research described in the literature and/or derived from the consensus of panel members based on clinical experience or knowledge that does not meet the above-listed criteria. This category is used only in cases where the provision of some guidance was deemed valuable but an adequately compelling clinical literature addressing the subject of the recommendation was deemed insufficient to justify placement in one of the other categories (A through C)

BOX 1. ACSM Consensus Statements and Recommendations Summary.

Consensus Statements and Recommendations

  • 1. The administration of AAS in a dose-dependent manner significantly increases muscle strength, lean body mass, endurance, and power. The effects are primarily seen when AAS use is accompanied by a progressive training program. Evidence Category A .
  • 2. Historically, AAS use was primarily seen in competitive athletes and aspiring bodybuilders and powerlifters. Recreational AAS use appears to have surpassed athletic AAS use indicated by survey prevalence estimates demonstrating that recreational trainees are the leading consumers of AAS. The ACSM deplores the illicit use of AAS for recreational purposes. Evidence Category C .
  • 3. AAS are classified as schedule III drugs, banned by several sport governing bodies, and are illegal to use for athletic purposes. The ACSM deplores the illicit use of AAS for recreational use and performance enhancement in athletes. Evidence Category D .
  • 4. Coaches, trainers, and medical staffs should monitor and be cognizant of visible signs of AAS use and abuse. These include (but are not limited to): a substantial increase in muscle mass, strength, and power in a relatively short period of time (or the reverse which could denote AAS withdrawal); acne that is resistant to medical treatment; development of unexplainable rash, gynecomastia, increased body hair, and prominent increases in surface vascularity; changes in temperament, mood, and aggressive behavior (severe depression or suicidality could indicate AAS withdrawal); facial masculinization and fluid retention; and muscle mass that appears disproportionate to body structure or pubertal status in young athletes. In addition, the presence of AAS-related materials (books, articles, websites, dealer information, needles, vials) on the individual could reflect intent and may warrant further dialogue from the coaching, trainer, and medical staffs. Medical staff should be aware of regulations and documentation requirements regarding use of AAS for athletes with medical indications for their use. Evidence Category C .
  • 5. Use and abuse of AAS is associated with several notable adverse effects in men and women including (but not limited to) suppression of the hypothalamic-pituitary-gonadal axis, psychological changes, immunosuppression, and unhealthy cardiovascular, hematological, reproductive, hepatic, renal, integumentary, musculoskeletal, and metabolic effects. Several adverse effects may be reversible upon discontinuation but some could pose health risks beyond the duration of AAS use. Evidence Category B .
  • 6. Use of AAS in prepubertal and peripubertal children may lead to early virilization, premature growth plate closure, and reduced stature. Evidence Category C .
  • 7. Coaches, trainers, and medical staffs should be cognizant of the reasons for AAS use and abuse and deter use when possible. Prevention programs based on education may assist; and providing the individual with scientific nutrition and training advice is a recommended strategy to mitigate the temptation of AAS use. Evidence Category D .
  • 8. Androgen replacement therapy is approved for the medical treatment of several clinical diseases and abnormalities. The ACSM acknowledges the lawful and ethical use of AAS for clinical purposes and supports the physicians’ ability to provide androgen therapy to patients when deemed medically necessary. The reader is referred to guidelines established by the Endocrine Society ( 4 ). Evidence Category C .

INTRODUCTION

Anabolic-androgenic steroids are drugs chemically and pharmacologically related to testosterone (T) that promote muscle growth and are not estrogens, progestins, or corticosteroids. An androgen is any natural or synthetic steroid hormone capable of promoting the development of male primary and secondary sexual characteristics via binding to androgen receptors at the tissue level. The term anabolic describes a hormone or other substance capable of enhancing the growth of somatic tissue, such as skeletal muscle and bone. In a sport-related setting, this is typically used to describe the enhancement of skeletal muscle. Table 2 presents nomenclature associated with AAS. In the United States, AAS are classified as Schedule III controlled substances ( 5 ). Although AAS have legitimate medicinal use, nontherapeutic use among athletes and recreationally active young men and women is performed to improve strength, power, increase muscle mass, and improve appearance. Athletic and recreational (i.e., noncompetitive) use of AAS has been widespread over the last 50 yr, creating considerable interest by the scientific and medical communities, as well as sport governing bodies, in examining the potential medical, legal, and ethical issues surrounding the use of these substances. All major national and international sports organizations have banned the illicit use of AAS by athletes.

Testosterone Hormone with strong anabolic and androgenic effects. produced by the testes in males. lesser quantities are produced by ovaries in women and by the adrenal glands in both sexes. the hypothalamus and pituitary regulate testosterone production in humans.
Testosterone derivative Chemically altered testosterone resulting in changes in solubility, pharmacokinetics, and/or clinical effects.
Testosterone esters Testosterone derivative with an ester group bound to testosterone to enhance oil solubility. This slows testosterone absorption and increases duration of effect, and allows for depot injections of testosterone
SARM Drugs designed to optimize anabolic tissue growth, while minimizing androgenic side effects. No current clinical applications, but research suggests potential therapeutic benefit in cancer, prostatic hyperplasia, and hypogonadism
Designer anabolic-androgenic steroids Synthetic steroids fabricated with intent to evade drug testing, or current laws prohibiting nonprescribed use
Testosterone enhancers/boosters Variety of substances purported to increase testosterone levels or effects, usually by increasing endogenous testosterone production or decreasing metabolism
HCG An analog to LH. Stimulates Leydig cells in the testes. Increases testosterone levels and sperm production
Clomiphene citrate (Clomid) Estrogen receptor modulator. Increases LH production. Has been shown relieve hypogonadal symptoms and maintain testosterone levels in men with symptomatic hypogonadism for up to 3 yr
Kisspeptins Peptide that appears to be important for onset of puberty and regulation of sperm production. Current evidence not definitive in regards to effects on androgen production in humans.

HISTORICAL PERSPECTIVES

Anabolic-androgenic steroids use has been examined extensively in various chapters, books, meta-analyses, and reviews ( 5–12 ). The effects of testicular extracts and castration on animals and humans have been a source of fascination for thousands of years ( 13,14 ). Suggestions that the consumption of testis tissue could improve impotence were noted ~140 BC ( 13 ). The mid 1700s to late 1800s marked a time where interest in testicular endocrinology increased ( 14 ). Table 3 depicts a brief historical timeline of some key events in AAS use in athletes. Testosterone was synthesized and biochemically described in the late 1920s and 1930s, and a host of different synthetic variations have been developed since ( 5,15,16 ). Testosterone or AAS use by athletes began in the 1940s and 1950s, and increased considerably thereafter, culminating in high usage during the 1968 Olympic Games ( 5,6 ). It has been speculated that the first appearance of AAS use among female athletes dates back to the late 1950s/early 1960s in Soviet track and field athletes ( 17 ).

Year Event
1889 Brown-Sequard suggest increases in muscle strength and endurance can occur following the injection of testicular extracts over a span of 2 wk
1896 Zoth and Pregel first suggested androgen use in athletes and investigated effects of testicular extracts on muscular strength and athletic performance
1927 Fred Koch and Lemuel McGee (University of Chicago) isolated an impure but potent form of T by pulverizing several tons of bull testicles
1928 The IAAF were first to ban doping agents
1935 Testosterone was isolated and the first two papers on testosterone synthesis were published—oral and injectable preparations were available to the medical community shortly thereafter—Kochakian reported T stimulates anabolism and suggested therapies could be useful for several disorders
Early 1940s Case studies suggested that human subjects were given testosterone in Germany and undocumented reports suggested AAS was administered to German soldiers during World War II
1942 Samuels, Henschel, and Keys published “Influence of methyl testosterone on muscular work and creatine metabolism in normal young men” in the Journal of Clinical Endocrinology and Metabolism—an early study investigating AAS in men
1945 De Kruif published “The Male Hormone” and suggested interest in athletes using testosterone to see the beneficial effects
Late 1940s West Coast bodybuilders began experimenting with T preparations
1952 Legendary strength athlete and coach Bob Hoffman of York Barbell speculates that Soviet athletes were using hormones during the Olympic Games
1954 Dr. John Ziegler was told by Soviet coaches that Soviet Weightlifters were using testosterone—he returned to United States and experimented on himself along with some weightlifters
1958 Ciba Pharmaceutical Company manufactured Dianabol—soon thereafter Dr. Ziegler began administering the drug to some of the York Barbell Weightlifting team
1963 Reports of NFL players first using AAS
1964 Androgen use increased greatly primarily in strength and power sports
1965 Oral Turinabol was synthesized by a German state-owned pharmaceutical company
A few scientists gained interest in studying AAS and performance in men as 2 studies were published between 1965 and 1969
1967 The IOC established a medical commission and developed a list of prohibited substances and methods
1968 Large increases in AAS use was seen including stacking and doses exceeding 2–5 times therapeutic doses—estimated that at least 1/3 of US track & field team and most of the German team used drugs in the 1968 Olympics
1969 The editor of Track & Field News (John Hendershott) called AAS the “Breakfast of Champions”
1973–1974 First testing procedures for androgens proposed radioimmunoassay and gas chromatography and mass spectrometry (RIA, GC-MS) and used in 1974 at the Commonwealth Games in Auckland New Zealand where 9 of 55 samples testing positive for androgens
1976 Drug testing instituted at the Olympic Games in Montreal—only 8 of 275 tests were deemed positive despite the majority of athletes admitting to using AAS in training—athletes began shifting to T from AAS as a result of drug testing
1976 The ACSM National Conference included a symposium and roundtable meeting on AAS in sports—two polarized groups evolved: those who thought AAS were “fool’s gold” or “myth” versus those who understood the ergogenic potential of the drugs—the prevailing medical opinion was that AAS were ineffective until the 1980s (possibly to dissuade use in part) which lead to mistrust between athletes and the medical community leading athletes to the black market for drugs and information
1977 The ACSM publishes the “Position statement on the use and abuse of anabolic-androgenic steroids in sports”—concluded that “… there is no conclusive evidence that extremely large doses of anabolic-androgenic steroids either aid or hinder performance …”
1980 Dr. Manfred Donike developed a method for testing testosterone—the 6:1 T:E ratio
1981 1st edition of the (by Dan Duchaine [nicknamed “The Steroid Guru”]) was published
1982–1983 Methods to circumvent T:E ratio (i.e., use of hCG, clomiphene, epitestosterone, and timing of T administration) were used to enable doping without detection
1984 ACSM National Conference included symposium on “Drug Use in Sports” with >12 scientific presentations with several focusing on AAS in athletes
1987 The ACSM publishes an updated position stand “The use of anabolic-androgenic steroids in sports”—revised position to AAS in the presence of an adequate diet and training can contribute to increases in lean body weight and muscular strength
1988 Testing for masking agents and diuretics begins—US government passes the Anti-Drug Abuse Act which made distribution or possession of AAS for nonmedical reasons a federal offense
1990 US government passes Anabolic Steroid Control Act—inserted 27 AAS and related drugs as Class III drugs where simple possession could result in incarceration
1994 US Congress passes Dietary Supplement Health and Education Act to protect consumers from certain substances
2001 The WADA was formed
2002 The advent of designer AAS began
2004 The Anabolic Steroid Control Act is revised to include 26 new compounds including prohormones
2005 T:E ratio lowered to 4:1 by WADA for a positive doping test

The sophistication of AAS use by athletes in the late 1960s was characterized by a host of different “stacking routines” (i.e., the consumption of two or more drugs in an attempt to improve the response) using various oral and injectable AAS preparations ( 5 ). Initially, many physicians did not believe AAS improved performance, and the International Olympic Committee (IOC) did not include AAS on the banned substance list. The ACSM adopted this position in their first AAS position stand in 1977 but later corrected in the 1987 publication ( 1 ). Although the 1970s marked a time where AAS use was known mostly among competitive athletes, the 1980s marked a time where AAS use spread well beyond athletics to gyms, health clubs, and public awareness of AAS use increased. The Anti-Drug Abuse Act (1988), Anabolic Steroid Control Act (1990, 2004), and Dietary Supplement Health and Education Act (1994) were enacted, in part, to stem the growing use of AAS. Only a few studies (~17) on AAS use and strength/hypertrophy increases were conducted before the 1980s, and these cumulatively showed minimal effects in untrained men, but significant responses in trained men, despite doses less than that used by many athletes ( 6,7,10 ). The sophisticated protocols and array of drugs used recreationally and by athletes remained a “black box” from a research perspective.

Of current concern is the ease by which AAS users may obtain AAS via the Internet and the proliferation of men’s health clinics. In addition to the use of AAS by competitive athletes, a growing segment of AAS users are nonathletes. Management of men with damaged hypothalamic-pituitary-gonadal regulatory pathways became a new area of medicine resulting in indiscriminate AAS use ( 18 ). Interest in AAS persists as research identifies new information regarding the performance and health aspects of the drugs and through stories of purported use in the sports world. The World Anti-Doping Agency (WADA) has developed new antidoping measures, including blood sampling, guidelines for international information gathering and sharing and revamping their “Athlete Biological Passport” guidelines. While AAS use in sports continues, increases in AAS use in the general population appear to have outpaced athletic use in the last decade ( 19 ).

EPIDEMIOLOGY OF AAS USE

Peer-reviewed studies examining the frequency of illicit AAS use have declined in the past decade despite concern over the growing AAS epidemic in the United States. These studies often rely on self-reports and are fraught with sampling bias, small sample sizes, possible confusion regarding supplement and AAS use, and suboptimal ascertainment ( 5 ). Many AAS users are secretive, with one survey finding that 56% of respondents would not disclose their physicians’ use ( 20 ). Athletes may be unwilling to discuss their use with researchers even when anonymity and confidentially are guaranteed for fear it may jeopardize their career; thus, leading to differences in what athletes reported on surveys versus their actual activities ( 21 ).

In 2014, the National Institute on Drug Abuse estimated that 1.3 million Americans were AAS users, while the Endocrine Society estimated between 2.9 and 4.0 million Americans have used AAS at some point in their lives ( 18,22,23 ). Other reports showed that the number of users might be as high as 4 million men in the United States, with ~100,000 new AAS users annually ( 6,23,24 ). The age of onset of use begins later than most drugs, with only 6% of users starting before 18 ( 23 ).

Although the general public and medical communities attribute AAS use primarily to competitive athletes ( 6 ), research does not support this misperception. Muscle dysmorphia (“megarexia”) is a dominant risk factor for illicit AAS use and indicates that AAS use is often used in pursuit of a more muscular appearance rather than for enhanced athletic performance ( 25 ). Recreationally active individuals age 15 to 24 yr are more likely to use AAS than athletes participating in organized sport ( 26 ). However, reports on the prevalence of illicit AAS use in athlete and nonathlete populations are widely variable. Anabolic-androgenic steroids have been reported in 9% to 67% of elite athletes, while reports of AAS use among gym attendees ranged from 3.5% to 80% ( 27 ). In all areas, men report higher prevalence than women, although the prevalence in women is increasing ( 28 ). Studies in girls have shown prevalence rates between 0.4% and 1.0% in adolescents, ~1.2% in collegiate athletes, and ~10.3% in elite athletes ( 27 ). Others have reported AAS use in young athletes ranging between 0.6% and 6.6% in teenage boys, 0.0% to 3.3% in teenage girls, and between 0.8% and 9.1% for collegiate male athletes ( 29–32 ). Peer-reviewed studies report the highest prevalence of use in weightlifters, powerlifters and bodybuilders, with rates ranging from 33.3% to 79.5% ( 31,33 ).

Several studies have examined sport and activity participation among self-reported AAS users. A survey study of >500 male AAS users (mean age of 29) showed ~70% were recreational exercisers versus 12% competitive bodybuilders, 8% competitive weightlifters, and 9% competitive athletes in other sports ( 34 ). Participation in high school sports was not associated with an increased risk of AAS use ( 34 ). A survey of 12 female AAS users indicated that 33% of the women were recreational users, while 67% participated in competitive bodybuilding and weightlifting. These women used a polypharmacy approach, but their weekly dose was lower than male AAS users ( 35 ). Female users were less likely to stack, more likely to pyramid and less likely to inject AAS than male users ( 35 ).

Rates of AAS use in athletes are sometimes inferred from rates of positive doping tests. However, this data has some inherent limitations, including ongoing updates to banned substances lists, variable drug testing methodologies, and variable lists of targeted substances tested by organizations that do not follow WADA protocols. It has been estimated that drug testing alone may underestimate drug use in elite athletes by 8-fold ( 21 ). The Anti-Doping Administration and Management System maintained by WADA now allows any sports body to share drug testing information. While AAS use in particular divisions, such as men’s vs women’s and underage athletes is still difficult to obtain, the testing databases now include much larger numbers of athletes than in the past. Anabolic agents constitute 87% of atypical findings reported by WADA and 46% of all adverse analytical findings (International Amateur Athletics Federation) ( 36,37 ). Stanazolol and nandrolone have the highest number of AAF at 20% and 14%, respectively, while an “unidentified anabolic agent” (e.g., “designer” AAS) was the third most common at 11% ( 36 ).

The true nature of AAS use and abuse in athletes and recreationally trained individuals is difficult to discern and is often underestimated. In addition to surveys and doping results, other sources of information on AAS use include investigated journalism and government hearings. Unfortunately, all of these methods have significant methodological issues that reduce their estimation accuracy ( 17 ). Journalists have interviewed current and former athletes, coaches, team physicians, and trainers whose estimate of AAS use in sports is much higher than survey reports. There has been an inconsistency between the number of individuals demonstrating signs of AAS use and the statistical prevalence generated via surveys. Drug testing is often limited by circumventing positive tests and has done little to quantify “real-life” use or dissuade AAS use at high levels of competition. Obtaining accurate measures of AAS use in athletes is difficult given the challenges of reducing bias; testing issues, and sincerity needed during interviews and survey completion, for example, fear of accountability, fear of loss of potential income or suspension, or fear of being perceived as a cheater or athlete who needed drugs to be successful.

Attempts have been made to identify the type of individual prone to using AAS ( 38–40 ). Hildebrandt et al. ( 39 ) reported 4 clusters of users from highest to lowest risk, each with different levels of motivation for AAS use: 1) polypharmacy (i.e., use of multiple drugs) approach with high risk (~11%); 2) fat burning (~17%); 3) muscle building (~21%); and 4) low-risk use designed to reduce fat and build muscle (~52%). Others have reported a four-level typology: 1) expert type (exemplifies controlled risk-taking, is knowledgeable about AAS and fascinated with effects on the human body, is scientific and may be focused on muscularity); 2) athlete type (interested in performance enhancement and is competitive); 3) well-being type (interested in looking and feeling good with low risk-taking); and 4) YOLO “You Only Live Once” type (is haphazard using risky behavior, quick improvements, impressing others and peer recognition is important) ( 38,40 ). Despite the typology, athletes’ motivation to use AAS is multi-faceted and influenced by many factors ( Table 4 ).

↑ muscle mass, strength, power, endurance, speed, fitness levels, energy, blood volume, BMD Encouragement from friends, family, coaches to do so
↑ sex drive ↑ recovery between workouts and competitions, pain tolerance
↑ athletic success, chances of winning, individual performance, strive to maximize potential Fear their opponents are using and they must use AAS in order to maintain competitive balance
↑ social recognition, peer acceptance, self-confidence, aggression, mental intensity and alertness Lack of fear in getting caught or being held accountable by sport governing bodies
↑ financial gain, chance at scholarship, notoriety ↓ body fat and injury risk, ↑ weight loss
Improve appearance, self-esteem, mood, personal protection ↓ aging effects, likelihood of muscle dysmorphia

Several extensive, national studies indicate an overall downward trend in lifetime AAS use among adolescents since peaking in the early 2000s ( 42 ). Monitoring the Future (MTF) is administered annually to a sample of 8th, 10th, and 12th grade students ( 43 ). The MTF reported peak prevalence rates for lifetime AAS use in 2000 to 2002 of 3% to 4% compared with 2018 data in Table 5 (i.e., ~1%–3%). The Youth Risk Behavior Survey (YRBS) is administered annually to a sample of high school students and reports an overall prevalence of 2.9% in 2017 (See Table 6 ), after peaking in 2001 at 5% ( 44 ). Although the YRBS is widely cited, concern has been raised that the term “steroid” is vague and potentially conflated with corticosteroids or steroid-like dietary supplements ( 45 ). Surveys that delineate the type of steroid show usage rates that are markedly lower than those seen in the YRBS data ( 45 ). Although AAS use rates in adolescents are low, ~1 in 8 AAS users initiates their use before age 18 ( 23 ). Several correlates of increased AAS use risk in this group include fitness-related activity ( 46,47 ); weight-related concerns (perceptions of very underweight or overweight status) ( 48,49 ); sexual preference and gender identity ( 25,44 ); and race and ethnicity ( 43,44 ). Some view current AAS use as an epidemic given the emergence of AAS availability through internet/mail order and “backroom” laboratories ( 18,50 ).

8th Grade 10th Grade 12th Grade
Overall 1.1% 1.2% 1.6%
Male 1.0% 1.3% 2.2%
Female 1.1% 0.9% 0.9%
White 1.0% 1.1% 1.4%
African American 1.2% 1.3% 2.9%
Hispanic 1.1% 1.0% 1.3%
Overall Females Males
Ever used steroids 2.9% 2.4% 3.3%
By race/ethnicity
 Black 2.2% 1.8% 2.7%
 White 3.6% 2.6% 4.6%
 Hispanic 3.5% 3.1% 3.8%
By sexual contact
 Opposite sex only 3.9% 2.6% 4.9%
 Same sex or both sexes 8.0% 7.2% 10.1%
 No sexual contact 0.7% 1.0% 0.5%

METHODS/PATTERNS OF AAS USE

Patterns of AAS use in athletes and resistance-trained populations vary greatly and depend upon: AAS type, self-administration routes, dosages, cycling patterns and durations, and ancillary drugs. A “polypharmacy approach” is commonly used where supraphysiologic doses of injectable and oral AAS are stacked and pyramided progressively in cycles, while ancillary drugs are consumed to minimize side effects, promote other areas of health and fitness, and/or enhance T levels during off-cycles, or periods in between cycles ( Table 7 ). Figure 1 depicts survey results from two studies on usage patterns for >2400 predominately male AAS users ( 34,41 ). These studies indicated that 99.2% of users reported using injectable AAS or a combination of oral and injectable AAS, and >40% used ancillary drugs, such as antiestrogens ( 41 ). Ip et al. ( 34 ) reported that 79% of AAS users “stacked” drugs, 18% used the “pyramid” approach (i.e., where drug intake is progressively increased, plateaus, and then is decreased or tapered until the end of the cycle), and only 9% thought physicians and pharmacists were knowledgeable about AAS. Interestingly, AAS users spent an average of 268 ± 472 h researching AAS prior to use ( 34 ).

AAS
Anabol 4–19 (norclostebol acetate) Myagen (bolasterone)
Anadrol (oxymetholone) Parabolan (trenbolone hexahydrobenzylcarbonate)
Anavar (oxandrolone)
Cheque drops (mibolerone) Primobolan (methenolone)
Dianabol (methandrostenolone) Primobolan depot (methenolone enanthate)
Deca durabolin (nandrolone decanoate) Proviron (mesterolone)
Durabolin (nandrolone phenylproprionate) Testosterone (androderm, AndroGel, Striant, testoderm)
Dynabol (nandrolone cypionate)
Dynabolan (nandrolone undecanoate) Testosterone blend (Sustanon, Omnadren, Equitest, Sten, Testoviron)
Equipoise (boldenone undecanoate)
Finajet (trenbolone acetate) Testosterone cypionate (Depo-Testosterone)
Genabol (norbolethone) TE (Delatestryl)
Halotestin (fluoxymesterone) Testosterone proprionate (Oreton)
Madol (desoxymethyltestosterone) Testosterone suspension (Andronaq)
Masteron (drostanolone) Tetrahydrogestrinone
Metandren (methyltestosterone) Trenabol (trenbolone enanthate)
Metribolone (methyltrienolone) Turinabol (chlorodehydromethyltestosterone)
Miotolan (furazabol) Winstrol (stanozolol)
Banned Prohormone/OTC Steroids
1-Testosterone Epi-DHT
4-Hydroxytestosterone 19-Norandrostenediol
Boldione 19-Norandrostenedione
Androstenediol, 1-,4-Androstenediol Halodrol
Androstenedione, 1-, 5-Androstenedione Superdrol
1-Androsterone Methylhydroxynandrolone
Androstanolone Prostanozol
Epiandrosterone, 1-Epiandrosterone
Andarine (S4) RAD-140 (testolone)
Ligandrol (LGD-4033) YK-11
Ostarine (enobosarm)
Antiestrogens
Arimidex (anastrozole) Faslodex (fulvestrant)
Aromasin (exemestane) Femara (letrozole)
Clomid (clomiphene citrate) Fertodur (cyclofenil)
Cytadren (aminoglutethimide) Lentaron (formestane)
Evista (raloxifene) Nolvadex (tamoxifen citrate)
Fareston (toremifene citrate) Teslac (testolactone)
Ancillary Drugs
Accutane (isotretinoin) Lasix (furosemide)
Cardarine Cytomel (liothyronine sodium)
Abuterol Synthroid (levothyroxine sodium)
Clenbuterol Human growth hormone (somatotropin, protropin, nutropin, humatrope, genotropin, norditropin)
Ephedrine
HCG
Catapres GHRH secratogues (CJC-1295, Mod GRF 1–29, Egrifta [tesamorelin acetate], Geref [sermorelin acetate])
Aldactone (spironolactone)
Dyrenium (triamterene)
Hydrodiuril (hydrochlorthiazide) Growth hormone releasing peptides (Lenomorelin [Ghrelin]. GHRP-1 to  GHRP-6, hexarelin [examorelin], ipamorelin,  ibutamoren mesylate [MK-677])
Epitestosterone
Probenecid (masking agent)
Synthol (site enhancer)
GH fragments (HGH fragment 176–191, AOD-9604)
IGF-1 and variants (Increlex [mecasermin])
Mechano growth factor
Insulin

F1

ANDROGEN PHYSIOLOGY

Testosterone is the principal androgen and has both androgenic (masculinizing) and anabolic (tissue building) effects. Testosterone is synthesized from cholesterol via the Δ-4 or Δ-5 pathways through the sequential action of several enzymes ( Fig. 2 ). In men, >95% of T is synthesized in the Leydig cells of the testes (with smaller adrenal contributions) under control of the hypothalamic-anterior pituitary-gonadal axis where gonadotropin-releasing hormone stimulates the release of luteinizing hormone (LH). Healthy men produce ~4 to 9 mg of T per day (10–35 nmol·L −1 ) whereas women have approximately 0.5 to 2.3 nmol·L −1 of circulating T in the blood ( 5 ). Gonadotropin-releasing hormone function is under the control of hypothalamic neuropeptides, such as kisspeptins, neurokinin-B, dynorphin-A, and phoenixins ( 51 ). In women, androgens are produced primarily by the ovaries and adrenal glands ( 52 ). Skeletal muscle produces small amounts of androgens ( 53 ). Testosterone circulates in the blood bound to sex hormone-binding globulin (44%–60%), albumin, orosomucoid, and cortisol-binding globulin. Testosterone and other 19-carbon androgens can be converted to 5α-dihydrotestosterone (DHT) by the action of steroid 5α-reductase or converted to estradiol or estrone by the aromatase enzyme. The liver inactivates T, and the resultant metabolites are excreted in the urine.

F2

Androgens perform many ergogenic, anabolic, and anticatabolic functions in skeletal muscle and neuronal tissue, leading to increased muscle strength, power, endurance, and hypertrophy in a dose-dependent manner ( 54 ). A meta-analysis concluded that short-term AAS use increases muscle strength substantially more than placebo and that strength gains and muscle hypertrophy are greater in trained individuals than in nontrained individuals ( 55 ). Gains in body mass and lean body mass (LBM) of ~5% to 20% from AAS use have been reported ( 56 ). Figure 3 depicts some physiological ramifications of androgens that could affect physical performance. However, the findings of controlled clinical trials of T and other AAS may differ from the practical experience of athletes due to the inclusion of mostly untrained subjects in controlled clinical trials; the use of lower doses of T or AAS in clinical trials than those used by many athletes; the use of multiple AAS in stacks with other drugs over long periods; and differences in nutritional patterns, training programs, and study design ( 5,27 ).

F3

Exogenous androgens are often administered orally or parenterally but are also available in cream, nasal spray, buccal, subcutaneous pellets, patches, and gel. Orally administered T is absorbed well but is degraded rapidly. The esterification of the 17-beta-hydroxyl group (e.g., T enanthate, cypionate, decanoate, undecanoate, propionate) makes the androgen more hydrophobic, causing a slow release from the muscle into circulation, increasing the duration of action. When administered intramuscularly, the androgen ester is slowly absorbed into the circulation, where it is then rapidly de-esterified by esterase enzymes to T. Intrinsic potency, bioavailability, and rate of clearance from the circulation are determinants of the biological activity. Other oral and injectable AAS are T, DHT, or 19-nortestosterone derivatives (e.g., methyltestosterone, methandrostenolone, fluoxymesterone, nandrolone decanoate, oxandrolone, trenbolone, stanozolol, and other designer-AAS).

An important and relevant question is how long the effects of a dose of AAS would last in an athlete? That is, how long would potential strength gains or gains in muscle mass persist? The answer to the question is undoubtedly complex and dependent on the AAS being used and their potency (see Fig. 1 ), the history of AAS in the athlete ( 57 ), the athlete’s training age, sex ( 58,59 ), and potentially the developmental stage of the athlete relative to puberty and adulthood (i.e., 18 yr of age). The literature in this area is, unsurprisingly, sparse, but some studies suggest that the effects of AAS persist for weeks after taking the steroids, but at ~12 wk after taking AAS that the effects, at least insofar as strength and muscle mass are concerned, are largely absent ( 55,60 ). For example, Giorgi et al. ( 61 ) showed that testosterone enanthate (TE) (3.5 mg·kg −1 ) administration for 12 wk during training resulted in greater increases in strength, muscle girth, and muscle thickness than a group given a placebo. However, after 12 wk without TE administration, but while still training, there was a reversion of strength and muscle in the TE group to levels no different from the placebo group. In contrast, others have observed preservation of AAS-induced gains in strength and LBM that persist after AAS usage has ceased, at least in the short-term ( 62 ).

Persistent and long-term (at least 5 yr) AAS use in a mixed sample of strength (strongman and powerlifters) and aesthetic sport (bodybuilding) athletes has been reported, in comparison to non-AAS, to result in persistent (i.e., in comparison to a matched group) elevations in LBM, muscle fiber area, capillary density, myonuclei density, and strength that were dose-dependent ( 57 ). The observation that long-term AAS use results in increased myonuclei density ( 57 ) suggesting that a much longer ‘muscle memory’ is perhaps possible in AAS users, particularly those who use AAS early in life. Evidence for such a mechanism comes from preclinical models ( 10 ), where young mice were exposed to AAS and subsequently increased their myonuclear content, resulting in a substantial hypertrophic advantage later in life. The authors of this work ( 63 ) even went so far as to suggest, “… the benefits of even episodic drug [AAS] abuse might be long lasting, if not permanent, in athletes. Our data suggest that the World Anti-Doping Code calling for only 2 yr of ineligibility after… [a doping violation for AAS] use… should be reconsidered.” Support for whether an AAS-induced increase in myonuclear number in humans is lacking; however, if present, then AAS-induced increases in myonuclei are theoretically advantageous to an athlete even if strength and lean mass advantages have been lost.

Residual effects of endogenous testosterone exposure in testosterone-suppressed transgender females are areas of active study and debate. These effects vary greatly depending upon the developmental stage of treatment initiation and will be much less when treatment is initiated before pubertal onset. There is a dichotomy when looking at measures of prepubertal athletic performance. Studies evaluating age-group athletic records report no significant differences in top age-group performances between boys and girls younger than 10 to 12 yr old ( 64–66 ). However, some studies evaluating more specific measures of strength and aerobic capacity reveal an 8% to 10% advantage in prepubertal biologic males relative to females, even after normalizing for body size ( 67,68 ). These performance differences may be residual effects from higher testosterone levels during early infancy (e.g., “mini-puberty”) and/or nonandrogenic genetic factors. Currently, there are no data on the durability of these performance differences in transgender females who start gender-affirming treatment before puberty.

Postpubertal testosterone suppression has variable impacts on performance-related parameters. Within 3 months of starting hormone suppression, hematocrit decreases by 4% to within normal values for cisgender females ( 69 ). A recent systematic review also evaluated evidence to date regarding treatment-related reductions in muscle size, strength, and LBM ( 70 ), summarized in Table 8 . Although the changes documented in Table 8 , along with an increase in fat mass, may contribute to significant reductions in athletic performance, the current lack of data in active or athletic populations makes the magnitude of these changes difficult to assess.

Cisgender Males (Reference) Cisgender Females (Relative to Cisgender Males) Transgender Females (Pretreatment, Relative to Cisgender Males) Reductions in Transgender Females with T Supression (12 mo Posttreatment)
LBM 100 70% 94%–92% −1% to 5.5%*
Muscle CSA 100 94%–88% −1.5% to 12%
Strength 100 64% (handgrip) 90%–86% (handgrip) 1.5% to –7% (handgrip)

ANDROGEN SIGNALING

Androgen signaling at the tissue level occurs primarily genomically through the classical androgen receptor (AR) with multiple levels of integration with other anabolic/catabolic pathways ( 71 ). Testosterone, DHT, and other AAS bind to cytoplasmic AR ( 72 ). Androgen receptor activity is altered at various sites; phosphorylation may augment androgen/AR transcriptional action (in the presence or absence of androgens) ( 73 ). Androgen receptor signaling is activated primarily by ligand binding, but under some circumstances through ligand-independent mechanisms (e.g., insulin like-growth factor-1 [IGF-1] induced mitogen-activated protein kinase-ERK1/2, p38 and c-Jun N-terminal kinase phosphorylation) ( 74 ) that may sensitize it to anabolic signals in the presence of low androgens ( 75 ). The AR is up-regulated following resistance training and short-term androgen administration ( 54 ).

Upon androgen binding to the ligand-binding domain (LBD) of the AR, the liganded AR undergoes phosphorylation, dimerization, and conformational changes, recruits coregulators, and translocates into the nucleus, where it regulates the transcription of androgen response elements (ARE) of the androgen-responsive genes ( 76 ). Androgen binding activates and stabilizes the AR, which is selectively induced by T, DHT, and AAS ( 77 ). Greater stability is seen with DHT than T ( 78 ). Binding affinity for the AR varies between androgens. Nandrolone and metenolone have a higher binding affinity than T, while stanozolol, methandienone, and fluoxymesterone have a lower binding affinity than T; and oxymetholone has a minimal binding affinity ( 79 ). Androgen binding to the AR completes the pocket that serves as a recruiting surface for co-activators ( 80 ). Some co-activators include BAF57 and 60a, SRC1 and 3, and ARA50 and 74. The activity of these co-regulators and the role of T in ribosome biogenesis may be important in mediating the anabolic effects of AAS on skeletal muscle.

Androgen/AR binding activates signaling through the Wnt-β-catenin pathway. The presence of T (in a dose-dependent manner) increases AR-β-catenin interaction and transcriptional capacity ( 81 ). Androgens promote myogenesis via multiple pathways. Satellite cells and myoblasts express AR and androgen binding, increasing satellite cell activation, proliferation, mobilization, differentiation, and incorporation into skeletal muscle ( 82 ). Androgens increase myogenesis via increased Notch signaling of satellite cells ( 83 ) and increased expression of IGF-1 ( 84 ). Androgen binding to AR on pluripotent mesenchymal cells increases their commitment to myogenesis and inhibits adipogenic differentiation via β-catenin signaling ( 85,86 ). Testosterone upregulates follistatin expression (which blocks signaling through the TGFβ-SMAD 2/3) and increases myogenic differentiation ( 82,84,86–88 ). Androgens may be anticatabolic by decreasing glucocorticoid receptor (GR) expression, interfering with cortisol binding, or the AR-T complex may compete with the cortisol-GR complex for cis -element binding sites on DNA ( 88–91 ).

Nongenomic AR signaling is rapid, with short latency periods acting independently of nuclear receptors ( 92 ). Nongenomic effects are thought to be mediated by direct binding to a target molecule, through intracellular AR activation (i.e., Src kinase), through a transmembrane AR receptor, or via changes in membrane fluidity ( 92 ). Nongenomic signaling involving G-protein 2nd messenger system and may either increase intracellular calcium concentrations via PI3K, phospholipase C, and IP 3 signaling ( 93 ), stimulate the activation of mitogen-activated protein kinase signaling ( 94 ), and mammalian target of rapamycin pathway signaling ( 95 ). Cross-talk between IGF-1 signaling and nongenomic AR signaling appears critical to mediating some anabolic effects ( 96 ). Nongenomic signaling occurs rapidly within seconds to minutes, much faster than classic genomic signaling, which takes hours and requires the constant presence of androgens to maintain intracellular signaling.

SIDE EFFECTS ASSOCIATED WITH ANDROGEN USE AND ABUSE

Investigations examining the safety of androgen use in various populations have been largely inadequate as there is tremendous variability in androgen dosages and patterns of use, including stacking of multiple AAS and concurrent use of accessory drugs ( 5 ). Figure 4 depicts the variety of adverse physiological and psychological effects associated with AAS use. These include relatively rare effects and those that are commonly expected, particularly with long-term AAS abuse ( 30 ).

F4

A survey of 500 AAS users (99% male) who had extensive experience (8 wk to 25 yr with 95% having >1–3 yr of AAS use) with high doses showed that 23% to 64% of respondents reported minor side effects (e.g., testicular atrophy, acne, fluid retention, insomnia, sexual dysfunction, gynecomastia) ( 97 ). Other common effects of AAS use include deleterious changes in cardiovascular (CV) risk factors: decreased plasma high-density lipoprotein (HDL) cholesterol ( 98 ), changes in clotting factors ( 99 ), and mood or psychiatric disturbances ( 79 ). Suppression of the hypothalamic-pituitary-testicular axis and spermatogenesis may result in infertility, while elevations in liver enzymes may reflect liver dysfunction ( 100–102 ). In one study, competitive athletes who used AAS during their competitive careers were more likely to die prematurely than athletes who did not ( 103 ). The use of nonsterile needles and needle sharing practices for intramuscular injections increase the risk for infection, muscle abscess, sepsis, and communicable diseases, such as human immunodeficiency virus (HIV) and hepatitis B and C ( 5 ).

Although CV effects are commonly reported with AAS use, based on an extensive review, the FDA concluded that “... the studies have significant limitations that weaken their evidentiary value for confirming a causal relationship between testosterone and adverse cardiovascular outcomes ” ( 104 ). Part of the difficulty in studying the effects of AAS on CV health is that the impacts of androgens on CV function vary with dose, method of administration, and aromatization potential ( 5 ). Parenteral administration of physiologic T replacement doses are associated with CV function and vary with dose, method of administration, and aromatization potential ( 5 ) with small decreases in plasma HDL, with little or no effect on total cholesterol, low-density lipoprotein (LDL) or triglycerides ( 105–107 ). However, supraphysiologic T doses are associated with significant reductions in HDL ( 108,109 ). Orally administered 17-alpha-alkylated, nonaromatizing AAS produce greater reductions in HDL and increases in LDL than when AAS are administered parenterally ( 110 ). Angell et al. ( 111 ) reported that self-administering AAS (median daily dose = 228 mg) for >2 yr was associated with smaller longitudinal LV strain, right ventricular (RV) ejection fraction, and altered diastolic function compared with nonusers. Others showed impaired RV free wall strain and strain rate associations with AAS abuse in competitive bodybuilders ( 112 ). D’Andrea et al. ( 113 ) showed associations between AAS use (~31 wk; weekly dose = 525 mg) and left atrial impairment (a marker of diastolic burden) in elite bodybuilders compared with nonusers. An increase in left ventricular (LV) mass occurs during resistance training ( 114–116 ); however, potential additional effects from AAS use in humans are unclear. In rats, only high T doses (up to 20 mg per kg body mass) induced cardiac hypertrophy with an impaired contractile process ( 117 ).

Deceased men who had used AAS showed greater cardiac mass than nonusers ( 118 ). Multivariate analysis indicated that increases in heart size were explained by increased body mass and by AAS use. Risk for adverse cardiac events associated with LV mass is supported by case reports detailing sudden death among power athletes who self-administered AAS ( 100,119–122 ). Case reports are largely anecdotal, and a causal relationship between AAS use and risk of sudden death has not been established. Strength/power athletes self-administering AAS have short QT intervals but increased QT dispersion compared with endurance athletes with similar LV mass who have long QT intervals but do not have increased QT dispersion ( 123 ). The interval from the peak to the end of the ECG T wave (Tp-e), Tp-e/QT ratio, and Tp-e/QTc ratio increases in AAS users, suggesting a link between AAS and ventricular arrhythmias, which may increase the risk for sudden death ( 124 ).

Increases in liver enzymes, cholestatic jaundice, hepatic neoplasms, and peliosis hepatis are associated with the use of oral, 17-alpha alkylated AAS ( 102,125,126 ), but not with parenterally administered T or its esters ( 127 ). The association between liver toxicity and AAS use is based on increases in AST and ALT. These enzymes are not liver-specific and are often elevated from muscle damage after resistance exercise ( 101,128 ); thus, possibly overstating the risk of hepatic dysfunction ( 128,129 ).

Endogenous LH and follicle stimulating hormone secretion are suppressed during AAS use, with subsequent effects on testicular T secretion and sperm count ( 130,131 ). Depending on the dose and duration of AAS use, endogenous T, LH, and follicle stimulating hormone may take weeks to months to return to homeostatic levels ( 132 ), and the long-term effects are not well understood. High-dose androgen administration in men is associated with breast tenderness and enlargement, for example, gynecomastia ( 5,133 ), thought to result from peripheral conversion of androgens to estrogens in men administering aromatizable AAS ( 134 ). The prevalence of gynecomastia is unknown, but prevalence rates as high as 54% were reported in AAS users ( 5 ). The use of nonsterile needles and needle-sharing practices for intramuscular injections increases the risk for infection, muscle abscess, sepsis, and communicable diseases, such as HIV and hepatitis B and C ( 5 ).

There is no evidence that T causes prostate cancer, but testosterone replacement therapy (TRT) is associated with a small increase in prostate specific antigen levels in older men with low T, which increases the risk of urological referral for prostate biopsy ( 5 ). Because many older men harbor subclinical prostate cancer, a prostate biopsy may lead to subclinical low-grade prostate cancer detection. Notably, however, TRT increases the risk of prostate biopsy.

The psychological effects of AAS use have garnered much publicity, especially on issues of aggression and suicide. However, the evidence is inconclusive due to the lack of sensitivity of the research instruments used to measure aggressive behavior, large variability in RT programs, preexisting personality or psychiatric disorders, and prevalence of multiple high-risk behaviors and use of other substances, such as alcohol, psychoactive drugs, and dietary supplements ( 5 ). Interestingly, physiologic T replacement in hypogonadal men may improve mood and attenuate negative aspects of mood ( 4 ). Morrison et al. ( 135 ) reported that the aggression and anxiety-provoking influences of androgens in animals are likely a developmental phenomenon and that adult exposure may be anxiolytic over the long term. However, underlying psychological dysfunction may cause a greater susceptibility to AAS use, and high doses of AAS may provoke a “rage” reaction in some individuals with preexisting psychopathology ( 136,137 ). Self-administration of AAS may increase the risk for mood disorders, such as mania, hypomania and depression ( 136,138 ). Resting T concentrations are related to posttraumatic stress (PTSD), in which higher T is associated with a lower risk for PTSD ( 139 ). Further, long-term use of AAS in former weightlifters was associated with poor cognitive function and negative changes in brain morphology ( 140,141 ). Approximately 30% of illicit AAS users will develop AAS dependence, and there is some overlap between AAS dependence and the mechanisms and risk for opioid dependence ( 142,143 ). Sudden discontinuation of exogenous AAS use in those who are dependent or have suppressed endogenous production may result in severe depression and suicidality ( 142,143 ). A multidisciplinary and medically supervised treatment program is indicated for individuals with AAS dependence.

Women self-administering AAS may undergo masculinization and experience hirsutism, deepening of the voice, enlargement of the clitoris, widening of the upper torso, decreased breast size, menstrual irregularities, and male pattern baldness ( 144 ). Some of these adverse effects may not be reversible ( 5 ).

Many of the side effects in adults may be seen in adolescents, but information on use in children is scant. Exogenous AAS exposure in preadolescence triggers pubertal onset and may result in early epiphyseal maturation and closure, leading to loss of ultimate height potential ( 40 ). Although mild acne is common during adolescence ( 40 ), AAS use may result in severe nodular acne, particularly on the back and shoulders, which is often resistant to treatment.

CLINICAL USES OF ANDROGEN THERAPY

Although athletes and recreational trainees have reported obtaining AAS from physicians for illicit purposes ( 26,33,50 ), several clinically approved uses of T exist. Of concern are potential illicit use stemming from a clinical prescription of T given the increased number of antiaging and wellness clinics. The sale of therapeutic T preparations in the United States quadrupled between 2001 and 2011 ( 145 ), and an estimated >2.3 million men received physician-prescribed T therapy as of 2013 ( 146 ). In military treatment facilities, the number of androgen prescriptions increased > twofold (23% per year) from 2007 to 2011, mainly in 35- to-44-yr-old men ( 147 ). Currently, therapeutic T is mostly used to treat primary (i.e., testicular failure) and secondary (i.e., reduced LH) hypogonadism ( 148 ). Androgen therapy has numerous clinical uses outlined in Table 9 ( 145,146 ). A substantial fraction of young men receiving T prescriptions are former AAS users trying to restore endogenous T production ( 149–151 ). The Endocrine Society Clinical Practice Guideline ( 148 ) details decision making regarding androgen therapy and the reader is referred to their specific guidelines on the diagnosis, treatment, and monitoring of hypogonadism in men ( 134 ).

 Male hypogonadism
 Primary
 Examples: Testicular trauma/torsion/irradiation, cryptorchidism, orchiectomy,  Klinefelter syndrome, chromosome abnormalities, LH and follicle stimulating  hormone receptor gene mutations, androgen synthesis disorders, myotonic  dystrophy, hypothyroidism
 Secondary
 Examples: Irradiation/tumor of hypothalamus or pituitary, drugs/medications  (opioids, marijuana, glucocorticoids, AAS), alcoholism, sleep deprivation,  surgery, trauma, eating disorder/relative energy deficiency, Kallman syndrome,  Prader-Willi syndrome
 Mixed primary and secondary
 Examples: diabetes, obesity, HIV infection, chronic obstructive pulmonary disease,  chronic kidney disease, liver disease, aging, cancer
 Hypoactive sexual desire disorder in postmenopausal females
 Constitutional delay of growth and puberty
 Gender-affirming treatment for transgender males
Contraindications:
 Cancer: prostate, breast, skin
 High prostate specific antigen
 Erythrocytosis/polycythemia
 Sleep apnea
 Venous thromboembolism
 CV disease
 Fertility problems

Testosterone replacement therapy has been shown to improve sexual activity ( 152–155 ), vertebral and femoral bone mineral density (BMD) and microarchitecture ( 156,157 ), hemoglobin content ( 158,159 ), LBM, maximal voluntary strength and physical function ( 160–164 ), and reduces body fat and BMI ( 162,165,166 ). There have also been reports of TRT reducing neuroinflammation and depressive symptoms ( 167–169 ), reducing blood pressure and improving lipid profiles ( 166 ), and neuronal regeneration ( 154,156,170–177 ), and may not change or improve cognitive function in older men ( 174,178,179 ). There is a low frequency of adverse events associated with TRT ( 2,148,153,180–190 ). However, all TRT should be accompanied by a structured monitoring plan ( 148 ). The Endocrine Society recommends evaluating symptoms, adverse events, lower urinary tract symptoms, and measurements of T levels, hematocrit, and prostate specific antigen at baseline, 3 to 6 months after starting treatment, and annually thereafter ( 148 ).

Testosterone and free T levels decline with advancing age after peaking in the second and third decades of life ( 191–194 ), leading to increased risk of sexual dysfunction; decreased muscle mass and strength, BMD, mobility; increased falls and fractures, late-life low grade persistent depressive disorder (dysthymia), and CV mortality ( 148,195 ). Low T is associated with an increased risk of diabetes, metabolic syndrome, and increased carotid artery intima-media thickness ( 196,197 ). Whether older men with age-related T decline should receive TRT remains a matter of debate. The Endocrine Society Guideline for TRT of hypogonadal men recommends against routinely prescribing T to all men, 65 yr or older, with low T levels ( 148 ). Decisions regarding TRT should be individualized after discussing potential risks and benefits in men with both symptoms suggestive of consistent T deficiency and burden of symptoms (e.g., low libido, unexplained anemia, osteoporosis) and presence of other co-morbid conditions that increase the risk of T treatment ( 148 ). The shared decision making should weigh the patient’s and clinician’s values. In male children, physiologic doses of T are used for brief periods to initiate pubertal development in those with constitutional delay of growth and puberty. Testosterone is needed permanently for children with congenital or acquired hypogonadism.

Recent interest has focused on the role of T in athletic performance in transgender and sexual developmentally distinct athletes. Individuals transitioning to females may require a therapeutic-use exemption for spironolactone, which is often used to block the androgen receptor and lower overall testosterone levels. Currently, trans female athletes subject to WADA testing must document subthreshold T levels for at least 12 months before being allowed to compete as a female. The IOC sets this threshold at <10 nM, and World Athletics (formerly the International Amateur Athletics Federation) at <5 nM. Interested readers can obtain a much deeper discussion of this topic in several reviews ( 198–200 ).

CONCLUSIONS

Anabolic-androgenic steroids include a wide spectrum of compounds that exert their effects through various mechanisms. Anabolic-androgenic steroid use is advantageous in athletic performance predominantly through enhancements in strength, power, increases in muscle mass, reduced recovery time, and other factors. Major competitive sporting bodies ban the use of AAS; however, the predominant area of AAS usage has now expanded into clinical scenarios, persons undergoing sexual reassignment, and by those interested in AAS for purely aesthetic enhancement. Thus, it is not only athletes who are using AAS to gain performance advantages but also other individuals for various reasons. Use for AAS to enhance athletic performance is banned, and coaches, trainers, and medical staff should monitor for signs of use. The use/abuse of AAS has several notable side effects with various consequences that are, in some cases, reversible. Coaches, parents, trainers, and medical staff need to understand why athletes might use AAS and provide educational programming in a preventive capacity. The position of the ACSM is that the illicit use of AAS for athletic and recreational purposes is, in many cases, illegal, unethical and also poses a substantial health risk. Nonetheless, TRT is used in treating various conditions, and clinicians may elect to use this therapy when medically necessary. The ACSM acknowledges the lawful and ethical use of AAS for clinical purposes and supports the physicians’ ability to provide androgen therapy to patients when deemed medically necessary.

This article is published as an official pronouncement of the American College of Sports Medicine and is an update of the 1987 ACSM position stand on the use of anabolic-androgenic steroids. Click here https://links.lww.com/MSS/C362 to download a slide deck that summarizes this ACSM pronouncement on anabolic-androgenic steroid use. This pronouncement was reviewed for the American College of Sports Medicine by members-at-large and the Pronouncements Committee.

Care has been taken to confirm the accuracy of the information present and to describe generally accepted practices. However, the authors, editors, and publisher are not responsible for errors or omissions or for any consequences from the application of the information in this publication and make no warranty, expressed or implied, with respect to the currency, completeness, or accuracy of the contents of the publication. The application of this information in a particular situation remains the professional responsibility of the practitioner; the clinical treatments described and recommended may not be considered absolute and universal recommendations.

  • Cited Here |
  • Google Scholar

TESTOSTERONE; HYPERTROPHY; SKELETAL MUSCLE; ANDROGEN; STRENGTH; PERFORMANCE

Supplemental Digital Content

  • MSS_2022_10_31_PHILLIPS_20-01062_SDC1.pptx; [PowerPoint] (4.24 MB)
  • + Favorites
  • View in Gallery

Readers Of this Article Also Read

Nutrition and athletic performance, the female athlete triad, exercise and physical activity for older adults, quantity and quality of exercise for developing and maintaining..., exercise/physical activity in individuals with type 2 diabetes: a consensus....

U.S. flag

An official website of the United States government

The .gov means it’s official. Federal government websites often end in .gov or .mil. Before sharing sensitive information, make sure you’re on a federal government site.

The site is secure. The https:// ensures that you are connecting to the official website and that any information you provide is encrypted and transmitted securely.

  • Publications
  • Account settings
  • My Bibliography
  • Collections
  • Citation manager

Save citation to file

Email citation, add to collections.

  • Create a new collection
  • Add to an existing collection

Add to My Bibliography

Your saved search, create a file for external citation management software, your rss feed.

  • Search in PubMed
  • Search in NLM Catalog
  • Add to Search

The science of steroids

Affiliations.

  • 1 Department of Biochemistry and Molecular Biology, Monash University, Melbourne, Vic, 3800, Australia; Division of Endocrinology & Metabolism, Hudson Institute, Monash Medical Centre, Clayton, Vic, Australia. Electronic address: [email protected].
  • 2 Department of Biochemistry and Molecular Biology, Monash University, Melbourne, Vic, 3800, Australia.
  • 3 The Richie Centre, Hudson Institute, Monash Medical Centre, Clayton, Vic, Australia; Department of Obstetrics & Gynaecology, Monash Medical Centre, Clayton, Vic, Australia.
  • PMID: 31147162
  • DOI: 10.1016/j.siny.2019.05.005

Steroids are complex lipophilic molecules that have many actions in the body to regulate cellular, tissue and organ functions across the life-span. Steroid hormones such as cortisol, aldosterone, estradiol and testosterone are synthesised from cholesterol in specialised endocrine cells in the adrenal gland, ovary and testis, and released into the circulation when required. Steroid hormones move freely into cells to activate intracellular nuclear receptors that function as multi-domain ligand-dependent transcriptional regulators in the cell nucleus. Activated nuclear receptors modify expression of hundreds to thousands of specific target genes in the genome. Steroid hormone actions in the fetus include developmental roles in the respiratory system, brain, and cardiovascular system. The synthetic glucocorticoid steroid betamethasone is used antenatally to reduce the complications of preterm birth. Development of novel selective partial glucocorticoid receptor agonists may provide improved therapies to treat the respiratory complications of preterm birth and spare the deleterious effects of postnatal glucocorticoids in other organs.

Keywords: Betamethasone; Fetal lung development; Glucocorticoid receptor; Glucocorticoids; Steroids.

© 2019 Elsevier Ltd. All rights reserved.

PubMed Disclaimer

Similar articles

  • An orphan nuclear receptor activated by pregnanes defines a novel steroid signaling pathway. Kliewer SA, Moore JT, Wade L, Staudinger JL, Watson MA, Jones SA, McKee DD, Oliver BB, Willson TM, Zetterström RH, Perlmann T, Lehmann JM. Kliewer SA, et al. Cell. 1998 Jan 9;92(1):73-82. doi: 10.1016/s0092-8674(00)80900-9. Cell. 1998. PMID: 9489701
  • Steroid signal transduction activated at the cell membrane: from plants to animals. Marcinkowska E, Wiedłocha A. Marcinkowska E, et al. Acta Biochim Pol. 2002;49(3):735-45. Acta Biochim Pol. 2002. PMID: 12422243 Review.
  • Nuclear receptor coactivator function in reproductive physiology and behavior. Molenda HA, Kilts CP, Allen RL, Tetel MJ. Molenda HA, et al. Biol Reprod. 2003 Nov;69(5):1449-57. doi: 10.1095/biolreprod.103.019364. Epub 2003 Jul 9. Biol Reprod. 2003. PMID: 12855594 Free PMC article. Review.
  • Hormonal influences during fetal lung development. Ballard PL. Ballard PL. Ciba Found Symp. 1980;78:251-74. doi: 10.1002/9780470720615.ch14. Ciba Found Symp. 1980. PMID: 6907082
  • Nuclear receptors outside the nucleus: extranuclear signalling by steroid receptors. Levin ER, Hammes SR. Levin ER, et al. Nat Rev Mol Cell Biol. 2016 Dec;17(12):783-797. doi: 10.1038/nrm.2016.122. Epub 2016 Oct 12. Nat Rev Mol Cell Biol. 2016. PMID: 27729652 Free PMC article. Review.
  • Steroid profiling in human primary teeth via liquid chromatography-tandem mass spectrometry for long-term retrospective steroid measurement. Wu RS, Hamden JE, Salehzadeh M, Li MX, Poudel A, Schmidt KL, Kobor MS, Soma KK. Wu RS, et al. PLoS One. 2024 Aug 28;19(8):e0309478. doi: 10.1371/journal.pone.0309478. eCollection 2024. PLoS One. 2024. PMID: 39197060 Free PMC article.
  • Differently increased volumes of multiple brain areas in Npc1 mutant mice following various drug treatments. Antipova V, Heimes D, Seidel K, Schulz J, Schmitt O, Holzmann C, Rolfs A, Bidmon HJ, González de San Román Martín E, Huesgen PF, Amunts K, Keiler J, Hammer N, Witt M, Wree A. Antipova V, et al. Front Neuroanat. 2024 Jul 16;18:1430790. doi: 10.3389/fnana.2024.1430790. eCollection 2024. Front Neuroanat. 2024. PMID: 39081805 Free PMC article.
  • Antithrombotic pharmacodynamics and metabolomics study in raw and processed products of Whitmania pigra Whitman. Kui H, Lei Y, Jia C, Xin Q, Tursun R, Zhong M, Liu C, Yuan R. Kui H, et al. Heliyon. 2024 Mar 19;10(7):e27828. doi: 10.1016/j.heliyon.2024.e27828. eCollection 2024 Apr 15. Heliyon. 2024. PMID: 38596067 Free PMC article.
  • The ethnomedicine, phytochemistry, and pharmacological properties of the genus Bersama: current review and future perspectives. Nigussie G, Ashenef S, Meresa A. Nigussie G, et al. Front Pharmacol. 2024 Mar 21;15:1366427. doi: 10.3389/fphar.2024.1366427. eCollection 2024. Front Pharmacol. 2024. PMID: 38576479 Free PMC article. Review.
  • Effects of Different Heat Treatments on Yak Milk Proteins on Intestinal Microbiota and Metabolism. Shu S, Jing R, Li L, Wang W, Zhang J, Luo Z, Shan Y, Liu Z. Shu S, et al. Foods. 2024 Jan 6;13(2):192. doi: 10.3390/foods13020192. Foods. 2024. PMID: 38254494 Free PMC article.

Publication types

  • Search in MeSH

Related information

  • Cited in Books

LinkOut - more resources

Full text sources.

  • Elsevier Science
  • Citation Manager

NCBI Literature Resources

MeSH PMC Bookshelf Disclaimer

The PubMed wordmark and PubMed logo are registered trademarks of the U.S. Department of Health and Human Services (HHS). Unauthorized use of these marks is strictly prohibited.

  • Open access
  • Published: 30 December 2019

Treatments for people who use anabolic androgenic steroids: a scoping review

  • Geoff Bates   ORCID: orcid.org/0000-0001-6932-2372 1 ,
  • Marie-Claire Van Hout 1 ,
  • Joseph Tay Wee Teck 2 &
  • Jim McVeigh 3  

Harm Reduction Journal volume  16 , Article number:  75 ( 2019 ) Cite this article

16k Accesses

36 Citations

15 Altmetric

Metrics details

A growing body of evidence suggests that anabolic androgenic steroids (AAS) are used globally by a diverse population with varying motivations. Evidence has increased greatly in recent years to support understanding of this form of substance use and the associated health harms, but there remains little evidence regarding interventions to support cessation and treat the consequences of use. In this scoping review, we identify and describe what is known about interventions that aim to support and achieve cessation of AAS, and treat and prevent associated health problems.

A comprehensive search strategy was developed in four bibliographic databases, supported by an iterative citation searching process to identify eligible studies. Studies of any psychological or medical treatment interventions delivered in response to non-prescribed use of AAS or an associated harm in any setting were eligible.

In total, 109 eligible studies were identified, which included case reports representing a diverse range of disciplines and sources. Studies predominantly focussed on treatments for harms associated with AAS use, with scant evidence on interventions to support cessation of AAS use or responding to dependence. The types of conditions requiring treatment included psychiatric, neuroendocrine, hepatic, kidney, cardiovascular, musculoskeletal and infectious. There was limited evidence of engagement with users or delivery of psychosocial interventions as part of treatment for any condition, and of harm reduction interventions initiated alongside, or following, treatment. Findings were limited throughout by the case report study designs and limited information was provided.

This scoping review indicates that while a range of case reports describe treatments provided to AAS users, there is scarce evidence on treating dependence, managing withdrawal, or initiating behaviour change in users in any settings. Evidence is urgently required to support the development of effective services for users and of evidence-based guidance and interventions to respond to users in a range of healthcare settings. More consistent reporting in articles of whether engagement or assessment relating to AAS was initiated, and publication within broader health- or drug-related journals, will support development of the evidence base.

Introduction

Human enhancement drug use differs from other forms of drug use by virtue of the motivation or purpose of their use. Typically, they are not consumed either for a treatment of an illness or injury nor for instant gratification through their psychoactive properties. Instead, their function is an attempt to change an individual’s appearance or improve a skill, ability or activity [ 1 , 2 ]. Characterised by man’s endeavour to gain an advantage over his competitor, their usage is by no means a new phenomenon, featured in social, ritual and sporting contexts throughout recorded history. Attempts to classify enhancement drugs have resulted in the six broad categories of drugs to increase lean muscle mass, to suppress appetite or reduce weight, to change the appearance of the hair or skin, to increase sexual desire or enhance performance, to improve cognitive function and to enhance mood or social interaction. Over the past 30 years, there has been growing media, policy and academic interest in this form of drug use, in particular the classification of drugs used to enhance musculature size and strength. Most notable within this category are the anabolic androgenic steroids (AAS) and their associated drugs [ 3 , 4 , 5 , 6 ]. Also included in this classification are a range of other hormones [ 7 , 8 , 9 , 10 , 11 ] including human growth hormone [ 12 , 13 ] and insulin [ 7 , 14 ].

While AAS doping remains a concern for sport, both at elite and recreational levels [ 15 , 16 , 17 ], the wider societal impact is now apparent [ 4 , 18 , 19 ]. Although prevalence estimates of clandestine behaviours such as AAS are notoriously difficult, a growing body of evidence has indicated that while well established in North America, northern Europe and Australia, there are concerns across the globe [ 6 , 19 ].

In recent years, research has provided a more nuanced understanding of AAS use in relation to the diverse characteristics and motivations of users [ 20 , 21 , 22 , 23 , 24 , 25 , 26 , 27 ], together with knowledge of the variety and patterns of drug use from both academic studies [ 28 , 29 , 30 , 31 , 32 , 33 , 34 ] and other sources [ 35 ]. Extensive research and comprehensive reviews have provided details of the identified adverse health conditions experienced by users of these durgs [ 36 ], while new research has identified new and concerning health risks [ 37 , 38 ] and the potential for transmission of blood-borne viruses [ 20 , 29 , 39 , 40 , 41 , 42 , 43 ].

A body of research has discussed the risk of developing AAS dependence and it is estimated that up to 30% of AAS users may develop dependence, characterised by the simultaneous use of multiple AAS in large doses over long periods of time [ 36 , 44 ]. While AAS are not explicitly recognised in the Diagnostic and statistical manual of mental disorders (DSM 5) as one of nine classes of drugs [ 45 ], they may be considered under the tenth ‘other (or unknown) substance’ class. The DSM 5 determines the severity of a substance use disorder from mild to severe according to the presence of up to 11 criteria. It is argued that while there are differences between AAS and psychoactive drugs dependence, such as that AAS are typically used over a period of weeks and months to increase muscularity rather than to achieve a ‘high’ in the short-term, these criteria are still highly applicable to AAS dependence [ 46 ]. Criteria such as tolerance, withdrawal, use of the substance in larger amounts, unsuccessful attempts to reduce or stop using the substance, and time spent on activity related to the substance use have all been identified as features of AAS dependence [ 44 , 46 ]. A number of hypotheses to explain AAS dependence have been put forward [ 47 , 48 ] and recommendations for treating what has been described as steroid ‘abuse’ or dependence have long been proposed [ 49 , 50 , 51 ].

Recent recommendations to treat steroid dependence include a staged discontinuation, managing withdrawal symptoms, maintaining abstinence and attenuating complications of chronic use [ 51 , 52 , 53 ]. Long-term use of AAS at high doses may lead to the development of a range of withdrawal symptoms following cessation, including depression, insomnia, suicidal ideation and fatigue, which may persist for many months [ 47 , 51 , 54 ]. Withdrawal is characterised by psychiatric and neuroendocrine symptoms, with the user ultimately re-initiating AAS to alleviate or avoid their onset. Supporting discontinuation may require a multidisciplinary approach with input from health professionals such as a GP, addiction specialist, psychiatrist and endocrinologist [ 53 ]. Swedish guidelines for diagnosing and treating AAS ‘abuse’ [ 55 ] include advice around psychosocial treatments, such as cognitive behavioural therapy, counselling group therapy and motivational interviewing. These therapies address the user’s preoccupation with enhancing their muscularity, their experiences of past bullying or violence, and resulting self-esteem and confidence issues. Brower (2009) believes that these entrenched psychological issues should be addressed once acute withdrawal is resolved as part of successful treatment [ 51 ]. Muscle dysmorphia and associated drive for muscularity [ 56 , 57 , 58 ] may be risk factors for both initiating and continuing AAS use, and potentially dependence [ 52 ]. It may be necessary to identify and address such disorders through counselling or psychotherapies as part of AAS treatment to reduce likelihood of re-initiation [ 53 ].

There has been a fourfold increase in the number of English language academic papers published between 1995 and 2015 [ 59 ]. However, there remains scant evidence in relation to effective policy and practice within the topic. While we have a greater understanding of the environmental influences and risk factors for use [ 17 , 60 , 61 , 62 ], there are few robust findings to support the effective prevention of AAS use. Little progress has been made in answering the fundamental questions of how do we make AAS less attractive and how do we make these drugs less accessible to those at risk of initiating use [ 63 , 64 , 65 , 66 ].

Tensions between some AAS users and the medical community are well documented [ 26 , 67 , 68 , 69 ] and long established [ 70 ], predating anti-doping or legislative control in most countries. Although psychological harm and the potential demand for interventions to address dependence are also well recognised [ 71 , 72 , 73 , 74 , 75 ] and diagnostic tools available [ 52 , 76 ], available services are few and far between. Harm reduction programmes, in the form of needle and syringe programmes (NSP), have clearly been successful in engaging AAS users in Australia [ 42 , 43 , 77 , 78 ] and, in particular, the United Kingdom [ 5 , 30 , 79 , 80 ]. However, even where uptake of service is high, substantial numbers of AAS users do not access these services [ 26 , 68 , 80 , 81 ]. Policy guidance regarding the delivery of harm reduction services for AAS users, centred around NSP provision, is in place in the United Kingdom [ 82 , 83 ], with its importance recognised in National Drug Strategy and Treatment guidelines [ 84 , 85 ]. While these guidelines are based on well-established principles of treatment engagement and harm reduction, there is an urgent need to identify where we have evidence to support specific interventions and where the evidence gaps remain.

The overall aim of this review was to identify and describe what is known about psychosocial and medical interventions that aim to support and achieve cessation of AAS, and treat and prevent associated health consequences. Specifically, the review aimed to identify:

What studies have examined the implementation and impact of interventions to support ASS cessation, and manage the health consequences related to cessation?

What studies have examined the implementation and impact of interventions to treat the harms or side effects associated with AAS use?

What are the implications of these findings, and what are the gaps in the evidence base that research in this area needs to address?

Methodology

The review was undertaken following Arksey and O’Malley’s guidance for scoping reviews, which informed the development of review methods and write-up of methods and findings [ 86 ].

Inclusion and exclusion criteria

Studies were eligible that included males or females with current or discontinued use of AAS alone, or AAS use alongside other substances. Use for any reason (for example, strength or sporting enhancement, aesthetic reasons) was acceptable with the exception of where AAS were prescribed or taken as part of a treatment regimen or in a controlled medical setting. Studies of any psychosocial or medical treatment interventions were eligible, including those that aimed to support individuals to discontinue AAS use or to treat the health consequences of current or past use. This included, but was not restricted to, treating AAS withdrawal, physical or psychological dependence, injuries, acute conditions, chronic conditions, side effects and blood-borne viruses. Studies that did not provide a description of the treatment given or those that did not describe any outcome following treatment at any follow-up time were excluded. Interventions that took place in any setting were eligible, including, but not restricted to, primary and secondary care, community settings such as drugs misuse services, NSPs and AAS clinics, sport and fitness environments, and prisons.

All types of study designs were considered due to the anticipated lack of high-quality controlled trials. Articles published in English were eligible with no date restrictions.

Search strategy

Initially, a comprehensive search was carried out in four bibliographic databases (Medline, PsycINFO, Sports Discus and the Social Sciences Citation Index) in January 2018. A search strategy was developed initially in Medline and adapted for the other databases. The full Medline search is provided in Additional file 2 .

The reference lists of all identified papers were screened to identify potentially eligible studies. Forward citation searches for included articles were executed in PubMed and the identified studies were assessed against the review inclusion criteria. This iterative process continued for all articles identified through these steps. Due to the nature of the evidence base, with studies likely to cover a broad range of topics and to be published in a wide variety of sources, these additional searches were expected to be important to identify relevant literature. Initially, titles and abstracts for all articles identified were reviewed against the inclusion criteria by one reviewer. A sample of 10% was independently reviewed by a second reviewer. The full texts for all articles included at this stage were retrieved and subjected to further screening against inclusion criteria.

Data extraction and synthesis

The relevant characteristics of identified studies were extracted into structured tables. This included population characteristics and details of their AAS use, the symptoms requiring treatment or reasons for seeking help, diagnosis, details of the treatment given and the outcomes of this treatment. Studies were grouped by the types of harms identified in Pope and colleagues’ review of the harms associated with AAS use [ 36 ]. A formal assessment of study quality was not undertaken, as this step is not recommended for scoping reviews [ 86 ]. However, comments on the overall nature, strengths and limitations of the evidence base are provided alongside discussion of review findings.

Identification of studies

Database searching identified 3,684 articles. Following screening of article title and abstracts against review inclusion criteria, full-text articles were accessed for 76 articles and these were again reviewed against the inclusion criteria. An additional 64 studies were identified through checking the reference lists and citations of the included articles. These were screened in the same manner. Following full-text screening, 46 articles were excluded, predominantly because no treatments were reported. The reasons for exclusion at this stage are reported in Fig. 1 .

figure 1

Flow of studies through the review

Summary of findings

In total, 109 studies met the review inclusion criteria. Summaries of the included studies are provided in Table 1 , grouped by the type of condition that required treatment. The studies were carried out in 28 countries, most prominently the USA ( n = 33) and the UK ( n = 21). One study followed a retrospective chart review design with the others case report ( n = 94) or case series ( n = 14) designs. With the lack of any controlled studies, it was difficult to draw conclusions relating to the effectiveness of any treatments provided. Additionally, there were substantial variations across studies in the depth of reporting about participants, settings, condition requiring treatment, the treatments provided and outcomes. The identified studies were published in sources representing a diverse range of disciplines.

Across the included studies, all participants were male. They included a wide range of ages, with the majority in their 20s and 30s, and represented a broad range of experience using AAS from recent initiators to long-term use. Participants’ motivations and history were not reported in a consistent manner to understand factors driving AAS use, but they were frequently described as participating in bodybuilding or weight-lifting activities. The types of conditions requiring treatment included psychiatric ( n = 12), neuroendocrine ( n = 11), hepatic ( n = 25), kidney ( n = 6), cardiovascular ( n = 26), musculoskeletal ( n = 13) and infectious ( n = 7). A further eight studies were categorised as ‘other’ disorders. In a small number of studies, participants were diagnosed with multiple conditions, but they have been grouped by the primary diagnosis.

Further details on participants’ AAS use, conditions requiring treatment, the treatments provided and outcomes are provided in Additional file 1 .

Treatment to support AAS cessation

Four studies reported abstinence-focussed interventions following a diagnosis of AAS dependence. In two cases, patients participated briefly in a drug treatment programme [ 88 , 97 ] before withdrawing. In one, the patient received medication and psychosocial interventions to manage AAS and opioid withdrawal [ 93 ] and withdrawal symptoms abated over time. Detail on the nature of these treatments was not provided. In the remaining study, the patient received medication for a short period before deciding to resume their AAS use due to withdrawal symptoms [ 98 ]. There was no evidence identified here, however, regarding psychosocial interventions that have sought to address any associated psychological disorders amongst users seeking treatment for their AAS use or any other condition. Additionally, no evidence was identified on approaches to reduce risk of relapse by developing social support systems, improving self-confidence or managing stress, all identified as potentially important factors to be addressed during AAS treatment [ 51 , 52 , 55 ].

Two studies were identified in this review where individuals who discontinued AAS use needed treatment for subsequent psychiatric symptoms including depression and suicidal ideation [ 87 , 89 ]. A further 11 studies reported treatments for neuroendocrine disorders, primarily with men who had discontinued their AAS use prior to the onset of symptoms. Administering AAS suppresses the hypothalamic–pituitary testicular axis, particularly when used in large amounts and for long periods, and inhibits production of testosterone [ 195 ]. Men who discontinue long-term AAS use are at risk of hypogonadism and while this may frequently be temporary and resolve spontaneously, it may in some cases persist for long periods after cessation, requiring medical treatment [ 51 , 196 , 197 , 198 ]. Symptoms of hypogonadism may be behind the withdrawal experiences of people with a dependence on AAS [ 51 ]. These difficult experiences have been identified as an influencing factor in users’ decisions to continue or re-instate AAS use [ 52 ]. The limited evidence here shows that positive outcomes are consistently reported in the treatment of men suffering with neuroendocrine disorders following AAS cessation.

Treatment for harms associated with AAS use

The bulk of the evidence identified related to current or former users receiving treatment for an acute or chronic condition or injury associated with their AAS use. This included psychiatric disorders ( n = 12), hepatic and kidney disorders ( n = 31), cardiovascular disorders ( n = 26), musculoskeletal disorders ( n = 13) and a range of other disorders ( n = 8). The management of such conditions in the AAS-using group is similar to that of the general population [ 53 ] and details are described in the tables in the additional material provided. There was, however, limited evidence of engagement with users regarding their AAS use as part of their more general treatment. There were examples where participants were stated to have discontinued AAS following treatment and remained abstinent at follow-up [ 133 , 157 , 159 ], but patients’ AAS status at this time was not routinely reported.

Treatment as an opportunity for engagement

In a small proportion of studies ( n = 10), it was reported that some form of intervention to bring about, or maintain change in AAS use was included as part of the treatment provided. This was most commonly instruction or advice to discontinue AAS use, with a more substantial element such as counselling only reported in three studies [ 139 , 145 , 180 ]. Where reported, such efforts were based on suppling risk information associated with AAS but not support with discontinuation, such as managing withdrawal symptoms. No form of harm reduction interventions were initiated alongside or following any treatments provided. Only one study [ 145 ] reported signposting or referral to another service for further support.

In comparison to people who use other psychoactive drugs, AAS users are less likely to suffer acute adverse effects from their substance use, or to have their occupational performance or relationships impaired and are, therefore, less reliant upon health professionals [ 44 ]. Research has consistently indicated this group to be reluctant to seek medical help or engage with health professionals [ 67 , 199 , 200 , 201 ]. Where health professionals identify AAS use in a patient and are providing treatment for an associated harm, this may, therefore, provide a rare opportunity to motivate changes in behaviour. There were examples in this review of studies that included recent initiators. For example, in 12/25 studies included here reporting hepatic disorders, patients had initiated AAS use fewer than 6 months prior to treatment. Contact with a health professional at this stage could provide a valuable opportunity to engage with the individual about their motivations and substance use before habitual use develops or becomes entrenched, or identify and treat any underlying factors. In a further 5/25 studies, long-term AAS use of over 5 years was reported, and up to 15 years. For such individuals, this contact could provide opportunity to test for disorders associated with long-term use, promote behaviour change and discuss long-term plans for discontinuation of use.

Encouraging discontinuation and delivering harm reduction with patients treated for a disorder associated with AAS

Where a patient is receiving treatment, there will be a range of factors that affect the appropriateness of delivering any form of AAS intervention or investigating any other potential harms. For example, in many of the studies identified, the individuals treated had discontinued their AAS use a substantial time prior to seeking treatment. Additionally, many were diagnosed with acute conditions, for which immediate, and in some cases substantial, treatment was required. In such cases, it is not surprising that the acute harm will be the focus of the treatment. However, where AAS use is suspected or confirmed, a number of diagnostic tests may be appropriate to identify potential physiological or psychiatric harms [ 53 ]. Recommendations for general practitioners who identify AAS use in a patient include strongly encouraging cessation and management of withdrawal symptoms in those that do discontinue, as well as information on injecting practices, promoting alternatives to AAS and informing about long-term health harms for those who continue to use [ 202 ]. Continued encouragement and monitoring of psychiatric and physiological complications is recommended for those who are not prepared to consider discontinuation [ 53 ].

An instruction not to use AAS may be effective in some cases, but for individuals who are highly motivated to use AAS in response to a desire to change their appearance or performance, it may have little impact. Experiencing harm or increasing knowledge of potential risks may not only reduce motivation to use amongst users who may accept risks as a potential consequence of use, but also one that they can manage through their practices [ 60 ]. Where it is identified that users intend to continue administering AAS following treatment, it is important that they receive appropriate harm reduction advice, such as on safe injecting, blood-borne viruses (BBVs) and AAS cycles. For example, in seven studies, treatments for infectious complications associating with injecting AAS were reported. There was no indication of relevant harm reduction work included alongside treatment, such as advice or demonstration relating to injecting or injecting techniques in any of these studies, with the exception of Rich and colleagues who reported provision of counselling on the risks of BBVs [ 180 ].

Research over the past 30 years has provided a far richer understanding of the populations of AAS users, their characteristics, behaviours and motivations. While the specific risks attached to each AAS and the probability or magnitude of harm associated with highly individualised and complex drug regimens cannot be known, we now have a far greater understanding of the potential harms caused by these drugs. However, the evidence base for interventions has not kept pace. The examples of treatment identified in this review were set within primary and secondary care facilities. No studies were identified that explored the effectiveness of any approaches to encourage cessation or treat dependence within other settings where health professionals are likely to encounter users, such as steroid clinics, drugs services or NSPs. Consequently, there is a lack of any evidence on the effectiveness of such services for bringing about behaviour change in users. Within any setting there is scarce evidence on treating AAS dependence, including initiating and maintain cessation and managing withdrawal symptoms outside of case reports of former users seeking support for neuroendocrine disorders.

The findings of this scoping review are characterised by missed opportunities. While the failure to report good practice or supplementary activity is not proof that it does not occur, without confirmation we cannot make assumptions. The extensive literature outlining the symptomatic treatment of AAS-related harms within numerous medical and surgical specialisms fails to provide evidence of intervention or referral to address the major causative factor, the patients’ AAS use. This scoping review has reported only a sample of the myriad of case reports involving the treatment of AAS-related harms. These case reports not only demonstrate the lack of evidence of intervention effectiveness to support the cessation of AAS use or reduce the associated harms, they also fail to show that actual activity occurred. As a minimum, future case reports should report if any assessment for AAS dependence were conducted. Details of advice or interventions provided to AAS users or any referral or signposting are also essential information. Referrals to primary care, endocrinologists, addiction specialists or harm reduction providers are essential building blocks in identifying care pathways and potential effective interventions. Case reports are published predominantly in clinical journals, often relating to medical or surgical specialisms. The publication of reports in broader health or public health journals or journals related to drug use, addiction or harm reduction would facilitate the inclusion of clinical experiences within a wider approach to addressing the harms associated with AAS use.

Despite the comprehensive research and literature relating to AAS dependence, there remains little evidence regarding effective interventions to support cessation of use or management of withdrawal. It is hoped that the development diagnostic tools [ 46 ], guidelines for clinical management [ 85 ] and harm reduction [ 82 ] or the commissioning of health services [ 83 ] will be accompanied by robust research and evaluation. Evaluations to date have been small scale and lack generalizability.

In addition to the need to ensure accurate and consistent reporting of activity and an upscaling of research and evaluation, there is a need to ensure that interventions are culturally appropriate to the target groups. Much of the work to date has focused on the bodybuilding communities of North America, Northern Europe and Australia. It is clear that AAS use is a global issue, with research emerging from low–middle income countries around the world in addition to industrialised high-income states. Of added significance is the diversity of individual AAS users. Interventions will need to be tailored to meet the varied characteristics and motivations of users, going beyond those looking to achieve a stylised “bodybuilding appearance” or excel at sport or even the young males attempting to bulk up. Evidence from the United Kingdom indicates that there are as many AAS users over 40 years of age as there are those under the age of 25 years [ 31 ]. It is well established that AAS use is not restricted to men and while rates amongst women are much lower [ 203 ], the complexities of treatment and care are undoubtedly much higher [ 23 , 204 , 205 ]. Prevalence of AAS use is higher amongst groups with specific characteristics such as professions where size or strength is an asset [ 206 , 207 , 208 , 209 ], amongst gay and bisexual men [ 20 , 22 , 29 , 210 , 211 ] and those using or who have previously used other drugs [ 212 ] [ 30 , 33 , 67 , 212 , 213 , 214 ]. These “sub groups” may or may not require specific interventions and may merely illustrate the complexities of human nature. The majority of AAS users will not initiate or continue AAS by virtue of membership of one of these groups but will have a range of susceptibilities and motivations for use.

Beyond these challenges, to develop effective services for users of AAS is the ongoing lack of confidence that some communities of AAS users feel towards health care professionals and primary care in particular [ 30 , 67 , 199 ] and a feeling that reliable and relevant health information can be gained elsewhere [ 215 ]. Built on the long-standing dismissive approach towards the effectiveness of anabolic steroids by elements of the health profession [ 216 , 217 ] and an ongoing ‘just say no’ stance amongst some practitioners, it is evident that establishing trust through listening to the AAS-using communities will be an essential element of intervention and service development [ 26 ].

Conclusions

This scoping review of the literature has identified treatments given to AAS users for a wide range of physiological and psychological harms. Despite the large number of articles identified, the evidence base consists of case reports of predominantly treatment of physiological harms and there is scarce evidence on treating dependence, managing withdrawal, or initiating behaviour change in users in any settings. Evidence is urgently required to support the development of effective services for users and of evidence-based guidance and interventions to respond to users in a range of healthcare settings. More consistent reporting in articles of whether engagement or assessment relating to AAS was initiated, and publication within broader health- or drug-related journals, will support development of the evidence base.

Availability of data and materials

All data generated or analysed during this study are included in this published article and its supplementary information files.

Abbreviations

  • Anabolic androgenic steroids

Blood-borne virus

Diagnostic and statistical manual of mental disorders

Needle and syringe programme

Evans-Brown M, McVeigh J, Perkins C, Bellis M. Human enhancement drugs: the emerging challenges to public health. Liverpool: North West Public Health Observatory; 2012.

Google Scholar  

McVeigh J, Evans-Brown M, Bellis MA. Human enhancement drugs and the pursuit of perfection. Adicciones. 2012;24(3):185–90.

Article   PubMed   Google Scholar  

Kanayama G, Pope HG. History and epidemiology of anabolic androgens in athletes and non-athletes. Mol Cell Endocrinol. 2018;464(C):4–13.

Article   CAS   PubMed   Google Scholar  

Kanayama G, Kaufman MJ, Pope HG. Public health impact of androgens. Curr Opin Endocrinol. 2018;25(3):218–23.

Article   CAS   Google Scholar  

McVeigh J, Begley E. Anabolic steroids in the UK: an increasing issue for public health. Drugs. 2017;24(3):278–85.

Sagoe D, Pallesen S. Androgen abuse epidemiology. Curr Opin Endocrinol Diabetes Obes. 2018;25(3):185–94.

Anderson LJ, Tamayose JM, Garcia JM. Use of growth hormone, IGF-I, and insulin for anabolic purpose: Pharmacological basis, methods of detection, and adverse effects. Mol Cell Endocrinol. 2017.

Van Hout MC, Hearne E. Netnography of female use of the synthetic growth hormone CJC-1295: pulses and potions. Subst Use Misuse. 2016;51(1):73–84.

Brennan R, Van Hout MC, Wells J. Heuristics of human enhancement risk: a little chemical help? Int J Health Prom Educ. 2013;51(4):212–27.

Article   Google Scholar  

Thevis M, Thomas A, Kohler M, Beuck S, Schanzer W. Emerging drugs: mechanism of action, mass spectrometry and doping control analysis. J Mass Spectrom. 2009;44(4):442–60.

Thevis M, Beuck S, Thomas A, Kortner B, Kohler M, Rodchenkov G, et al. Doping control analysis of emerging drugs in human plasma - identification of GW501516, S-107, JTV-519, and S-40503. Rapid Commun Mass Spectrom. 2009;23(8):1139–46.

Holt RI, Sonksen PH. Growth hormone, IGF-I and insulin and their abuse in sport. Br J Pharmacol. 2008;154(3):542–56.

Article   CAS   PubMed   PubMed Central   Google Scholar  

Evans-Brown M, McVeigh J. Injecting human growth hormone as a performance-enhancing drug-perspectives from the United Kingdom. J Substance Use. 2009;14(5):267–88.

Underwood M. "Slin is the safest and most anabolic hormone": exploring bodybuilders' use of insulin as a performance and image enhancing drug. Drug Alcohol Rev. 2018;37:S70–S.

Ulrich R, Pope HG, Cleret L, Petroczi A, Nepusz T, Schaffer J, et al. Doping in two elite athletics competitions assessed by randomized-response surveys. Sports Med. 2018;48(1):211–9.

Mottram DR, Chester N. Drugs in Sport. 7 ed: Routledge; 2018.

Backhouse SH, Griffiths C, McKenna J. Tackling doping in sport: a call to take action on the dopogenic environment. Br J Sports Med. 2017.

Auchus RJ, Brower KJ. The public health consequences of performance-enhancing substances: who bears responsibility? JAMA. 2017;318(20):1983–4.

Sagoe D, Molde H, Andreassen CS, Torsheim T, Pallesen S. The global epidemiology of anabolic-androgenic steroid use: a meta-analysis and meta-regression analysis. Ann Epidemiol. 2014;24(5):383–98.

Ip EJ, Doroudgar S, Shah-Manek B, Barnett MJ, Tenerowicz MJ, Ortanez M, et al. The CASTRO study: unsafe sexual behaviors and illicit drug use among gay and bisexual men who use anabolic steroids. Am J Addict. 2019;28:101.

PubMed   Google Scholar  

Blashill AJ, Calzo JP, Griffiths S, Murray SB. Anabolic steroid misuse among US adolescent boys: disparities by sexual orientation and race/ethnicity. Am J Public Health 2017(0):e1-e3.

Griffiths S, Murray SB, Dunn M, Blashill AJ. Anabolic steroid use among gay and bisexual men living in Australia and New Zealand: Associations with demographics, body dissatisfaction, eating disorder psychopathology, and quality of life. Drug Alcohol Dependence. 2017;181:170–6.

Börjesson A, Gårevik N, Dahl M-L, Rane A, Ekström L. Recruitment to doping and help-seeking behavior of eight female AAS users. Subst Abuse Treat Prev Policy. 11(2016, 1):11.

Zahnow R, McVeigh J, Bates G, Hope V, Kean J, Campbell J, et al. Identifying a typology of men who use anabolic androgenic steroids (AAS). Int J Drug Policy. 2018;55:105–12.

Christiansen AV, Vinther AS, Liokaftos D. Outline of a typology of men’s use of anabolic androgenic steroids in fitness and strength training environments. Drugs. 2016;24(3):295–305.

Underwood M. The unintended consequences of the current approach to blood borne virus prevention amongst people who inject image and performance enhancing drugs: a commentary based on enhanced bodybuilder perspectives. Int J Drug Policy. 2019;67:19–23.

Ip EJ, Trinh K, Tenerowicz MJ, Pal J, Lindfelt TA, Perry PJ. Characteristics and behaviors of older male anabolic steroid users. J Pharm Pract. 2015;28(5):450–6.

Teck JTW, McCann M. Tracking internet interest in anabolic-androgenic steroids using Google Trends. Int J Drug Policy. 2018;51:52–5.

Ip EJ, Yadao MA, Shah BM, Doroudgar S, Perry PJ, Tenerowicz MJ, et al. Polypharmacy, infectious diseases, sexual behavior, and psychophysical health among anabolic steroid-using homosexual and heterosexual gym patrons in San Francisco's Castro District. Subst Use Misuse. 2017;52(7):959–68.

Hope VD, McVeigh J, Marongiu A, Evans-Brown M, Smith J, Kimergard A, et al. Prevalence of, and risk factors for, HIV, hepatitis B and C infections among men who inject image and performance enhancing drugs: a cross-sectional study. BMJ Open. 2013;3(9):e003207–e.

Article   PubMed   PubMed Central   Google Scholar  

Begley E, McVeigh J, Hope V, Bates G, Glass R, Campbell J, et al. Image and Performance Enhancing Drugs: 2016 National Survey Results. Liverpool: Liverpool John Moores University; 2017.

Dodge T, Hoagland MF. The use of anabolic androgenic steroids and polypharmacy: A review of the literature. Drug Alcohol Depend. 2011.

Sagoe D, McVeigh J, Bjornebekk A, Essilfie MS, Andreassen CS, Pallesen S. Polypharmacy among anabolic-androgenic steroid users: a descriptive metasynthesis. Subst Abuse Treat Pr. 2015;10(ARTN 12).

Jennings CJ, Patten E, Kennedy MC, Kelly C. Examining the profile and perspectives of individuals attending harm reduction services who are users of performance and image enhancing drugs. Merchants Quay Ireland: Dublin; 2014.

Llewellyn W. Anabolics. 11 ed. Jupiter, FL: Molecular Nutrition; 2017.

Pope HG, Wood RI, Rogol A, Nyberg F, Bowers L, Bhasin S. Adverse health consequences of performance-enhancing drugs: an Endocrine Society scientific statement. Endocr Rev. 2014;35(3):341–75.

Westlye LT, Kaufmann T, Alnaes D, Hullstein IR, Bjornebekk A. Brain connectivity aberrations in anabolic-androgenic steroid users. Neuroimage Clin. 2017;13:62–9.

Bjornebekk A, Walhovd KB, Jorstad ML, Due-Tonnessen P, Hullstein IR, Fjell AM. Structural brain imaging of long-term anabolic-androgenic steroid users and nonusing weightlifters. Biol Psychiatry. 2017;82(4):294–302.

Article   PubMed   CAS   Google Scholar  

Hope VD, McVeigh J, Smith J, Glass R, Njoroge J, Tanner C, et al. Low levels of hepatitis C diagnosis and testing uptake among people who inject image and performance enhancing drugs in England and Wales, 2012-15. Drug Alcohol Depend. 2017;179:83–6.

Hope VD, Harris R, McVeigh J, Cullen KJ, Smith J, Parry JV, et al. Risk of HIV and hepatitis B and C over time among men who inject image and performance enhancing drugs in England and Wales: Results from cross-sectional prevalence surveys, 1992-2013. Jaids-J Acq Imm Def. 2016;71(3):331–7.

Ip EJ, Yadao MA, Shah BM, Lau B. Infectious disease, injection practices, and risky sexual behavior among anabolic steroid users. Aids Care. 2016;28(3):294–9.

Rowe R, Berger I, Yaseen B, Copeland J. Risk and blood-borne virus testing among men who inject image and performance enhancing drugs, Sydney, Australia. Drug Alcohol Rev. 2017;36(5):658–66.

Iversen J, Hope VD, McVeigh J. Access to needle and syringe programs by people who inject image and performance enhancing drugs. Int J Drug Policy. 2016;31:199–200.

Kanayama G, Brower K, Wood R, Hudson J, Pope H. Anabolic-androgenic steroid dependence: an emerging disorder. Addiction. 2009;104(12):1966–78.

Association AP. Diagnostic and statistical manual of mental disorders (DSM-5®): American Psychiatric Pub; 2013.

Book   Google Scholar  

Kanayama G, Brower KJ, Wood RI, Hudson JI, Pope HG. Issues for DSM-V: clarifying the diagnostic criteria for anabolic-androgenic steroid dependence. Am J Psychiatr. 2009;166(6):642–4.

Kashkin KB, Kleber HD. Hooked on hormones?: An anabolic steroid addiction hypothesis. JAMA. 1989;262(22):3166–70.

Brower KJ. Rehabilitation for anabolic-androgenic steroid dependence. Clin Sport Med. 1989;1:171–81.

Giannini AJ, Miller N, Kocjan DK. Treating steroid abuse: a psychiatric perspective. Clin Pediatr. 1991;30(9):538–42.

Corcoran JP, Longo ED. Psychological treatment of anabolic-androgenic steroid-dependent individuals. J Subst Abuse Treat. 1992;9(3):229–35.

Brower KJ. Anabolic steroid abuse and dependence in clinical practice. Phys Sportsmed. 2009;37(4):131–40.

Kanayama G, Brower KJ, Wood RI, Hudson JI, Pope HG Jr. Treatment of anabolic-androgenic steroid dependence: Emerging evidence and its implications. Drug Alcohol Depend. 2010;109(1-3):6–13.

Casavant M, Griffith J. Anabolic steroid use disorder New Jersey: BMJ Americas Office; 2017 [Available from: https://bestpractice.bmj.com/topics/en-gb/987 ]. Accessed Aug 2018.

Malone DA, Dimeff RJ, Lombardo JA, Sample RH. Psychiatric effects and psychoactive substance use in anabolic-androgenic steroid users. Clin J Sport Med. 1995;5(1):25–31.

Arver S, Borjesson A, Bottiger Y, Edin A, Garevic N, Lundmark J, et al. Swedish clinical guidelines on: The abuse of anabolic androgenic steroids and other hormonal drugs. Stockholm: Karolinska University Hospital; 2013.

Rohman L. The relationship between anabolic androgenic steroids and muscle dysmorhpia: a review. Eating Disord. 2009;17(3):187–99.

Kanayama G, Barry S, Hudson J, Pope H. Body image and attitudes toward male roles in anabolic-androgenic steroid users. Am J Psychiatr. 2006;163(4):697–703.

Pope H, Gruber AJ, Choi P, Olivardia R, Phillips KA. Muscle dysmorphia: an underrecognized form of body dysmorphic disorder. Psychosomatics. 1997;38(6):548–57.

McVeigh J. The public health implications of anabolic steroid use in the United Kingdom. A shot in the Dark: Steroids, IPEDs - the Hidden Harm Conference; 26th April 2018; Colchester2018.

Hanley Santos G, Coomber R. The risk environment of anabolic-androgenic steroid users in the UK: examining motivations, practices and accounts of use. Int J Drug Policy. 2017;40:35–43.

Bates G, Tod D, Leavey C, McVeigh J. An evidence-based socioecological framework to understand men’s use of anabolic androgenic steroids and inform interventions in this area. Drugs. 2018:1–9.

Pope HG, Kanayama G, Hudson JI. Risk factors for illicit anabolic-androgenic steroid use in male weightlifters: a cross-sectional cohort study. Biol Psychiatr. 2012;71(3):254–61.

de Ronde W. Preventing anabolic steroid abuse: a long way to go. J Intern Med. 2018.

Bates G, Begley E, Tod D, Jones L, Leavey C, McVeigh J. A systematic review investigating the behaviour change strategies in interventions to prevent misuse of anabolic steroids. J Health Psychol. 2017. https://doi.org/10.1177/1359105317737607 .

Backhouse S, Collins C, Defoort Y, McNamee M, Parkinson A, Sauer M. Study on doping prevention: a map of legal, regulatory and prevention practice provisions in EU 28. Luxembourg: Publications Office of the European Union; 2014. Report No.: 9279435426.

NICE. Drug misuse prevention: targeted interventions (NG64). London: NICE; 2017.

Zahnow R, McVeigh J, Ferris J, Winstock A. Adverse effects, health service engagement, and service satisfaction among anabolic androgenic steroid users. Contemp Drug Prob. 2017;44(1):69–83.

Kimergard A, McVeigh J. Variability and dilemmas in harm reduction for anabolic steroid users in the UK: a multi-area interview study. Harm Reduct J. 2014;11(ARTN 19).

Dunn M, Henshaw R, McKay F. Understanding health service use and needs of performance and image enhancing drug users in regional Queensland. Drug Alcohol Rev. 2014;33:24.

Wade CH. Anabolic steroids: doctors denounce them, but athletes aren't listening. Science. 1972;176:1401–3.

Brower KJ, Blow FC, Beresford TP, Fuelling C. Anabolic-androgenic steroid dependence. J Clin Psychiatry. 1989;50(1):31–3.

CAS   PubMed   Google Scholar  

Copeland J, Peters R, Dillon P. A study of 100 anabolic-androgenic steroid users. Med J Aust. 1998;168(6):311–2.

Midgley SJ, Heather N, Davies JB. Dependence-producing potential of anabolic-androgenic steroids. Addict Res. 1999;7(6):539–50.

Pope HG Jr, Kanayama G, Athey A, Ryan E, Hudson JI, Baggish A. The lifetime prevalence of anabolic-androgenic steroid use and dependence in Americans: current best estimates. Am J Addict. 2014;23(4):371–7.

Ip EJ, Lu DH, Barnett MJ, Tenerowicz MJ, Vo JC, Perry PJ. Psychological and physical impact of anabolic-androgenic steroid dependence. Pharmacotherapy. 2012;32(10):910–9.

Pope HG Jr, Kean J, Nash A, Kanayama G, Samuel DB, Bickel WK, et al. A diagnostic interview module for anabolic-androgenic steroid dependence: preliminary evidence of reliability and validity. Exp Clin Psychopharmacol. 2010;18(3):203–2013.

Van de Ven K, Maher L, Wand H, Memedovic S, Jackson E, Iversen J. Health risk and health seeking behaviours among people who inject performance and image enhancing drugs who access needle syringe programs in Australia. Drug Alcohol Rev. 2018:doi: https://doi.org/10.1111/dar.12831 .

Jacka B, Peacock A, Degenhardt L, Bruno R, Clare P, Kemp R, et al. Trends in PIEDs use among male clients of needle-syringe programs in Queensland, Australia; 2007-2015. Int J Drug Policy. 2017;46:74–8.

Kimergard A, McVeigh J. Environments, risk and health harms: a qualitative investigation into the illicit use of anabolic steroids among people using harm reduction services in the UK. BMJ OPEN. 2014;4(6).

McVeigh J, Beynon C, Bellis MA. New challenges for agency based syringe exchange schemes: analysis of 11 years of data (1991–2001) in Merseyside and Cheshire, United Kingdom. Int J Drug Policy. 2003;14(5-6):399–405.

Glass R, Hope VD, Njoroge J, Edmundson C, Smith J, McVeigh J, et al. Secondary distribution of injecting equipment obtained from needle and syringe programmes by people injecting image and performance enhancing drugs: England and Wales, 2012-15. Drug Alcohol Depend. 2018;195:40–4.

NICE. Needle and syringe programmes NICE public health guidance. NICE: National Institute for Health and Care Excellence; 2014.

Public Health England. Providing effective services for people who use image and performance enhancing drugs. London: PHE Publications; 2015.

HM Government. 2017 Drug Strategy. London 2017.

Department of Health. Drug misuse and dependence: UK guidelines on clinical management. London: Department of Health; 2017.

Arksey H, O'Malley L. Scoping studies: towards a methodological framework. Int J Soc Res Methodol. 2005;8(1):19–32.

Malone DA, Dimeff RJ. The use of fluoxetine in depression associated with anabolic steroid withdrawal: a case series. J Clin Psychiatry. 1992;53(4):130–2.

Hays LR, Littleton S, Stillner V. Anabolic steroid dependence. Am J Psychiatr. 1990;147(1):122.

Allnutt S, Chaimowitz G. Anabolic steroid withdrawal depression: a case report. The Canadian Journal of Psychiatry / La Revue canadienne de psychiatrie. 1994;39(5):317–8.

Papazisis G, Kouvelas D, Mastrogianni A, Karastergiou A. Anabolic androgenic steroid abuse and mood disorder: a case report. Int J Neuropsychopharmacol. 2007;10(2):291–3.

Gahr M, Kolle MA, Baumgarten E, Freudenmann RW. Mania related to mesterolone in a previously mentally healthy person. J Clinical Psychopharmacol. 2012;32(5):734–5.

Rashid W. Testosterone abuse and affective disorders. J Subst Abuse Treatment. 2000;18(2):179–84.

Ranjan R, Parmar A, Pattanayak RD, Dhawan A. Dependence on anabolic-androgenic steroids: a case report and brief review. Delhi Psychiatry J. 2014;17(2):481–4.

Duffy RM, Kelly BD. Steroids, psychosis and poly-substance abuse. Irish J Psychol Med. 2015;32(2):227–30.

Franey DG, Espiridion ED. Anabolic steroid-induced mania. Cureus. 2018;10(8):e3163–e.

PubMed   PubMed Central   Google Scholar  

Stanley A, Ward M. Anabolic steroids--the drugs that give and take away manhood. A case with an unusual physical sign. Med Sci Law. 1994;34(1):82–3.

Brower KJ, Blow FC, Beresford TP, Fuelling C. Anabolic-androgenic steroid dependence. J Clin Psychiatry. 1989;50:31.

Tennant F, Black DL, Voy RO. Anabolic steroid dependence with opioid-type features. N Engl J Med. 1988;319(9):578.

van Breda E, Keizer HA, Kuipers H, Wolffenbuttel BHR. Androgenic anabolic steroid use and severe hypothalamic-pituitary dysfunction: a case study. Int J Sports Med. 2003;24(3):195–6.

Tan RS, Vasudevan D. Use of clomiphene citrate to reverse premature andropause secondary to steroid abuse. Fertility Sterility. 2003;79(1):203–5.

Menon DK. Successful treatment of anabolic steroid-induced azoospermia with human chorionic gonadotropin and human menopausal gonadotropin. Fertil Steril. 2003;79(Suppl 3):1659–61.

Gill GV. Anabolic steroid induced hypogonadism treated with human chorionic gonadotropin. Postgrad Med J. 1998;74(867):45–6.

Gazvani MR, Buckett W, Luckas MJ, Aird IA, Hipkin LJ, Lewis-Jones DI. Conservative management of azoospermia following steroid abuse. Hum Reprod. 1997;12(8):1706–8.

Bickelman C, Ferries L, Eaton RP. Impotence related to anabolic steroid use in a body builder. Response to clomiphene citrate. Western J Med. 1995;162(2):158–60.

CAS   Google Scholar  

Turek PJ, Williams RH, Gilbaugh JH III, Lipshultz LI. The reversibility of anabolic steroid-induced azoospermia. J Urol. 1995;153(5):1628–30.

Cohen JJ, Honig S. Anabolic steroid-associated infertility: a potentially treatable and reversible cause of male infertility. Fertility Sterility. 2005;84:S223.

Pirola I, Cappelli C, Delbarba A, Scalvini T, Agosti B, Assanelli D, et al. Anabolic steroids purchased on the Internet as a cause of prolonged hypogonadotropic hypogonadism. Fertility Sterility. 2010;94(6):2331.e1–3.

Street C, Scally MC. Pharmaceutical intervention of anabolic steroid induced hypogonadism - our success at restoration of the HPG axis. Med Sci Sports Exercise. 2000;32(5S).

Jarow JP, Lipshultz LI. Anabolic steroid-induced hypogonadotropic hypogonadism. Am J Sports Med. 1990;18(4):429–31.

Martin NM, Abu Dayyeh BK, Chung RT. Anabolic steroid abuse causing recurrent hepatic adenomas and hemorrhage. World J Gastroenterol. 2008;14(28):4573–5.

Sánchez-Osorio M, Duarte-Rojo A, Martínez-Benítez B, Torre A, Uribe M. Anabolic-androgenic steroids and liver injury. Liver Int. 2008;28(2):278–82.

Chahla E, Hammami MB, Befeler AS. Hepatotoxicity associated with anabolic androgenic steroids present in over-the-counter supplements: a case series. International Journal of Applied. 2014;4(3).

Awai HI, Yu EL, Ellis LS, Schwimmer JB. Liver toxicity of anabolic androgenic steroid use in an adolescent with nonalcoholic fatty liver disease. J Pediatr Gastroenterol Nutr. 2014;59(3):e32–e3.

Krishnan PV, Feng Z-Z, Gordon SC. Prolonged intrahepatic cholestasis and renal failure secondary to anabolic androgenic steroid-enriched dietary supplements. J Clin Gastroenterol. 2009;43(7):672–5.

Socas L, Zumbado M, Perez-Luzardo O, Ramos A, Perez C, Hernandez JR, et al. Hepatocellular adenomas associated with anabolic androgenic steroid abuse in bodybuilders: a report of two cases and a review of the literature. Br J Sports Med. 2005;39(5):e27.

Patil JJ, O'Donohoe B, Loyden CF, Shanahan D. Near-fatal spontaneous hepatic rupture associated with anabolic androgenic steroid use: a case report. Br J Sports Med. 2007;41(7):462–3.

Solbach P, Potthoff A, Raatschen HJ, Soudah B, Lehmann U, Schneider A, et al. Testosterone-receptor positive hepatocellular carcinoma in a 29-year old bodybuilder with a history of anabolic androgenic steroid abuse: a case report. BMC Gastroenterol. 2015;15:60.

Gorayski P, Thompson CH, Subhash HS, Thomas AC. Hepatocellular carcinoma associated with recreational anabolic steroid use. Br J Sports Med. 2008;42(1):74–5 discussion 5.

El Khoury C, Sabbouh T, Farhat H, Ferzli A. Severe cholestasis and bile cast nephropathy induced by anabolic steroids successfully treated with plasma exchange. Case Reports Med. 2017;2017.

Li C, Adhikari BK, Gao L, Zhang S, Liu Q, Wang Y, et al. Performance-enhancing drugs abuse caused cardiomyopathy and acute hepatic injury in a young bodybuilder. Am J Mens Health. 2018;12(5):1700–4.

Stepien PM, Reczko K, Wieczorek A, Zarebska-Michaluk D, Pabjan P, Krol T, et al. Severe intrahepatic cholestasis and liver failure after stanozolol usage - case report and review of the literature. Clin Exper Hepatol. 2015;1(1):30–3.

Elsharkawy AM, McPherson S, Masson S, Burt AD, Dawson RT, Hudson M. Cholestasis secondary to anabolic steroid use in young men. BMJ. 2012;344:e468.

Ampuero J, Garcia ES, Lorenzo MM, Calle R, Ferrero P, Gomez MR. Stanozolol-induced bland cholestasis. Gastroenterol Hepatol. 2014;37(2):71–2.

Singh V, Rudraraju M, Carey EJ, Byrne TJ, Vargas HE, Williams JE, et al. Severe hepatotoxicity caused by a methasteron-containing performance-enhancing supplement. J Clin Gastroenterol. 2008;43(3):287.

Boks M, Tiebosch AT, van der Waaij LA. A jaundiced bodybuilder Cholestatic hepatitis as side effect of injectable anabolic–androgenic steroids AU - Boks. Marije N J Sport Sci. 2017;35(22):2262–4.

Winwood PJ, Robertson DA, Wright R. Bleeding oesophageal varices associated with anabolic steroid use in an athlete. Postgrad Med J. 1990;66(780):864–5.

Ding NS, De Cruz P, Lim L, Thompson A, Desmond P. Androgenic-anabolic steroid drug-induced liver injury. Intern Med J. 2013;43(2):215–6.

Nasr J, Ahmad J. Severe cholestasis and renal failure associated with the use of the designer steroid Superdrol (methasteron): a case report and literature review. Dig Dis Sci. 2009;54(5):1144–6.

Marcacuzco Quinto AA, Manrique Municio A, Loinaz Segurola C, Jimenez Romero LC. Spontaneous hepatic rupture associated with the use of anabolic steroids. Cirugia Espanola. 2014;92(8):570–2.

Hymel BM, Victor DW, Alvarez L, Shores NJ, Balart LA. Mastabol induced acute cholestasis: a case report. World J Hepatol. 2013;5(3):133–6.

Flores A, Nustas R, Nguyen HL, Rahimi RS. Severe cholestasis and bile acid nephropathy from anabolic steroids successfully treated with plasmapheresis. ACG Case Rep J. 2016;3(2):133–5.

Diaz FC, Saez-Gonzalez E, Benlloch S, Alvarez-Sotomayor D, Conde I, Polo B, et al. Albumin dialysis with MARS for the treatment of anabolic steroid-induced cholestasis. Ann Hepatol. 2016;15(6):939–43.

Pais-Costa SR, Lima OA, Soares AF. Giant hepatic adenoma associated with anabolic-androgenic steroid abuse: case report. Arquivos brasileiros de cirurgia digestiva : ABCD = Brazilian archives of digestive surgery. 2012;25(3):180-2.

Hardt A, Stippel D, Odenthal M, Hölscher AH, Dienes H-P, Drebber U. Development of hepatocellular carcinoma associated with anabolic androgenic steroid abuse in a young bodybuilder: a case report. Case Reports in Pathology. 2012;2012:195607.

Kesler T, Sandhu RS, Krishnamoorthy S. Hepatology: hepatocellular carcinoma in a young man secondary to androgenic anabolic steroid abuse. J Gastroenterol Hepatol. 2014;29(11):1852.

Merino Garcia E, Borrego Utiel FJ, Martinez Arcos MA, Borrego Hinojosa J, Perez Del Barrio MP. Kidney damage due to the use of anabolic androgenic steroides and practice of bodybuilding. Nefrologia. 2018;38(1):101–3.

Tarashande Foumani A, Elyasi F. Oxymetholone-induced acute renal failure: a case report. Caspian J Intern Med. 2018;9(4):410–2.

Daher EF, Silva Junior GB, Queiroz AL, Ramos LM, Santos SQ, Barreto DM, et al. Acute kidney injury due to anabolic steroid and vitamin supplement abuse: report of two cases and a literature review. Int Urol Nephrol. 2009;41(3):717–23.

Colburn S, Childers WK, Chacon A, Swailes A, Ahmed FM, Sahi R. The cost of seeking an edge: recurrent renal infarction in setting of recreational use of anabolic steroids. Anna Med Surgery (2012). 2017;14:25–8.

Samaha AA, Nasser-Eddine W, Shatila E, Haddad JJ, Wazne J, Eid AH. Multi-organ damage induced by anabolic steroid supplements: a case report and literature review. J Med Case Rep. 2008;2:340.

Shimada Y, Yoritaka A, Tanaka Y, Miyamoto N, Ueno Y, Hattori N, et al. Cerebral infarction in a young man using high-dose anabolic steroids. J Stroke Cerebrovasc Dis. 2012;21(8):906.e9–11.

Sveinsson O, Herrman L. Cortical venous thrombosis following exogenous androgen use for bodybuilding. BMJ Case Rep. 2013;2013.

Garg P, Davis G, Wilson JI, Sivananthan M. Intravascular ultrasound and angiographic demonstration of left main stem thrombus-high-risk presentation in a young adult with anabolic steroid abuse. Am Heart Hosp J. 2010;8(2):E125–E7.

Shamloul RM, Aborayah AF, Hashad A, Abd-Allah F. Anabolic steroids abuse-induced cardiomyopathy and ischaemic stroke in a young male patient. BMJ Case Rep. 2014;2014.

Luc JGY, Buchholz H, Kim DH, MacArthur RGG. Left ventricular assist device for ventricular recovery of anabolic steroid-induced cardiomyopathy. J Surg Case Rep. 2018;2018(8):rjy221–rjy.

Santamarina RD, Besocke AG, Romano LM, Ioli PL, Gonorazky SE. Ischemic stroke related to anabolic abuse. Clin Neuropharmacol. 2008;31(2):80–5.

Edvardsson B. Hypertensive encephalopathy associated with anabolic–androgenic steroids used for bodybuilding. Acta Neurologica Belgica. 2015;115(3):457–8.

Sonmez E, Turkdogan KA, Yilmaz C, Kucukbuzcu S, Ozkan A, Sogutt O. Chronic anabolic androgenic steroid usage associated with acute coronary syndrome in bodybuilder. Turkish J Emerg Med. 2016;16(1):35–7.

Falkenberg M, Karlsson J, Ortenwall P. Peripheral arterial thrombosis in two young men using anabolic steroids. Eur J Vasc Endovasc. 1997;13(2):223–6.

Laroche GP. Steroid anabolic drugs and arterial complications in an athlete--a case history. Angiology. 1990;41(11):964–9.

Youssef MYZ, Alqallaf A, Abdella N. Anabolic androgenic steroid-induced cardiomyopathy, stroke and peripheral vascular disease. BMJ Case Reports. 2011;2011. https://doi.org/10.1136/bcr.12.2010.3650 .

Bispo M, Valente A, Maldonado R, Palma R, Glória H, Nóbrega J, et al. Anabolic steroid-induced cardiomyopathy underlying acute liver failure in a young bodybuilder. World J Gastroenterol. 2009;15(23):2920.

Goldstein DR, Dobbs T, Krull B, Plumb VJ. Clenbuterol and anabolic steroids: a previously unreported cause of myocardial infarction with normal coronary arteriograms. South Med J. 1998;91(8):780–4.

Güneþ Y, Erbaþ C, Okuyan E, Babalýk E, Gürmen T. Myocardial infarction with intracoronary thrombus induced by anabolic steroids. Anatol J Cardiol. 2004;4(4):357–8.

Christou GA, Christou KA, Nikas DN, Goudevenos JA. Acute myocardial infarction in a young bodybuilder taking anabolic androgenic steroids: a case report and critical review of the literature. Eur J Prev Cardiol. 2016;23(16):1785–96.

Stergiopoulos K, Brennan JJ, Mathews R, Setaro JF, Kort S. Anabolic steroids, acute myocardial infarction and polycythemia: a case report and review of the literature. Vasc Health Risk Manag. 2008;4(6):1475–80.

Santos RP, Pereira A, Guedes H, Lourenço C, Azevedo J, Pinto P. Anabolic drugs and myocardial infarction - a clinical case report. Arquivos brasileiros de cardiologia. 2015;105(3):316–9.

CAS   PubMed   PubMed Central   Google Scholar  

Ýlhan E, Demirci D, Güvenç TS, Çalýk AN. Acute myocardial infarction and renal infarction in a bodybuilder using anabolic steroids. Turk Kardiyol Dern Ars. 2010;38(4):275–8.

Huie MJ. An acute myocardial infarction occurring in an anabolic steroid user. Med Sci Sports Exerc. 1994;26(4):408–13.

Ferenchick GS, Adelman S. Myocardial infarction associated with anabolic steroid use in a previously healthy 37-year-old weight lifter. Am Heart J. 1992;124(2):507–8.

Lau DH, Stiles MK, John B, Shashidhar, Young GD, Sanders P. Atrial fibrillation and anabolic steroid abuse. Int J Cardiol. 2007;117(2):e86–7.

Mewis C, Spyridopoulos I, Kuhlkamp V, Seipel L. Manifestation of severe coronary heart disease after anabolic drug abuse. Clin Cardiol. 1996;19(2):153–5.

Ment J, Ludman PF. Coronary thrombus in a 23 year old anabolic steroid user. Heart. 2002;88(4):342.

Ahlgrim C, Guglin M. Anabolics and Cardiomyopathy in a bodybuilder: case report and literature review. J Card Fail. 2009;15(6):496–500.

Joseph J, Naqvi S, Sturm E. Reversible anabolic androgenic steroid-induced cardiomyopathy. Cardiovasc Disord Med. 2017;2(3):1–3.

Nieminen MS, Ramo MP, Viitasalo M, Heikkila P, Karjalainen J, Mantysaari M, et al. Serious cardiovascular side effects of large doses of anabolic steroids in weight lifters. Eur Heart J. 1996;17(10):1576–83.

Stannard JP, Bucknell AL. Rupture of the triceps tendon associated with steroid injections. Am J Sports Med. 1993;21(3):482–5.

Farkash U, Shabshin N, Pritsch M. Rhabdomyolysis of the deltoid muscle in a bodybuilder using anabolic-androgenic steroids: a case report. J Athletic Train (National Athletic Trainers' Association). 2009;44(1):98–100.

Fenelon C, Dalton DM, Galbraith JG, Masterson EL. Synchronous quadriceps tendon rupture and unilateral ACL tear in a weightlifter, associated with anabolic steroid use. BMJ Case Rep. 2016;2016.

Bagherifard A, Jabalameli M, Rezazadeh J, Ghaffari S, Tabrizian P. Simultaneous bilateral quadriceps tendon rupture following a low - energy trauma in a male body builder with the history of anabolic - androgenic steroids consumption. Shafa Orthopedic J. 2018;5(2).

Liow RY, Tavares S. Bilateral rupture of the quadriceps tendon associated with anabolic steroids. British J Sports Med. 1995;29(2):77–9.

KramhØFt M, Solgaard S. Spontaneous rupture of the extensor pollicis longus tendon after anabolic steroids. J Hand Surg. 1986;11(1):87.

Tapaninen T, P V. Simultaneous bilateral rupture of the quadriceps tendon associated with anabolic steroids - a case report. Annals of Clinical Case Reports. 2016;1(1220).

David HG, Green JT, Grant AJ, Wilson CA. Simultaneous bilateral quadriceps rupture: a complication of anabolic steroid abuse. J Bone Joint Surgery British Volume. 1995;77(1):159–60.

Visuri T, Lindholm H. Bilateral distal biceps tendon avulsions with use of anabolic steroids. Med Sci Sports Exerc. 1994;26(8):941–4.

Freeman BJ, Rooker GD. Spontaneous rupture of the anterior cruciate ligament after anabolic steroids. Br J Sports Med. 1995;29(4):274–5.

Adamson R, Rambaran C, D'Cruz DP. Anabolic steroid-induced rhabdomyolysis. Hospital Med (London, England: 1998). 2005;66(6):362.

Erturan G, Davies N, Williams H, Deo S. Bilateral simultaneous traumatic upper arm compartment syndromes associated with anabolic steroids. J Emerg Med. 2013;44(1):89–91.

Leopardi P, Vico G, Rosa D, Cigala F, Maffulli N. Reconstruction of a chronic quadriceps tendon tear in a body builder. Knee Surg Sports Traumatol Arthrosc. 2006;14(10):1007–11.

Rich JD, Dickinson BP, Flanigan TP, Valone SE. Abscess related to anabolic-androgenic steroid injection. Med Sci Sports Exerc. 1999;31(2):207–9.

Friedman O, Arad E, Ben AO. Body builder's nightmare: black market steroid injection gone wrong: a case report. Plastic Reconstructive Surgery Global Open. 2016;4(9):e1040.

Shiber JR. Pyomyositis due to anabolic steroid injection. J Emerg Med. 2013;44(1):e69–70.

Evans NA. Local complications of self administered anabolic steroid injections. Br J Sports Med. 1997;31(4):349–50.

Tüzel E. Spontaneous corpus cavernosum abscess in a healthy man using long-term androgenic anabolic steroids. World J Mens Health. 2015;33(1):36–8.

Grant S, Dearing J, Ghosh S, Collier A, Bal AM. Necrotizing myositis of the deltoid following intramuscular injection of anabolic steroid. Int J Infect Dis. 2010;14(9):e823–4.

Marquis C, Maffulli N. Anabolic steroid related abscess - a risk worth taking? Injury Extra. 2006;37:451–4.

Ray S, Masood A, Pickles J, Moumoulidis I. Severe laryngitis following chronic anabolic steroid abuse. J Laryngol Otol. 2008;122(3):230–2.

Maini AAN, Maxwell-Scott H, Marks DJB. Severe alkalosis and hypokalemia with stanozolol misuse. Am J Emerg Med. 2014;32(2):196.e3–4.

Labib M, Haddon A. The Adverse effects of anabolic steroids on serum lipids. Ann Clin Biochem. 1996;33(3):263–4.

Cooper I, Reeve N, Doherty W. Delayed diagnosis of a cerebrovascular accident associated with anabolic steroid use. BMJ Case Rep. 2011;2011.

Unai S, Miessau J, Karbowski P, Baram M, Cavarocchi NC, Hirose H. Caution for anabolic androgenic steroid use: a case report of multiple organ dysfunction syndrome. Respiratory Care. 2013;58(12):e159–63.

Geraci MJ, Cole M, Davis P. New onset diabetes associated with bovine growth hormone and testosterone abuse in a young body builder. Hum Exp Toxicol. 2011;30(12):2007–12.

Alaraj AM, Chamoun RB, Dahdaleh NS, Haddad GF, Comair YG. Spontaneous subdural haematoma in anabolic steroids dependent weight lifters: reports of two cases and review of literature. Acta Neurochir (Wien). 2005;147(1):85–7 discussion 7-8.

Moor JW, Khan MI. Growth hormone abuse and bodybuilding as aetiological factors in the development of bilateral internal laryngocoeles. a case report. Eur Arch Otorhinolaryngol. 2005;262(7):570–2.

Tan RS, Scally MC. Anabolic steroid-induced hypogonadism – towards a unified hypothesis of anabolic steroid action. Med Hypotheses. 2009;72(6):723–8.

Kanayama G, Hudson JI, Pope HG. Long-term psychiatric and medical consequences of anabolic–androgenic steroid abuse: a looming public health concern? Drug Alcohol Depend. 2008;98(1):1–12.

Kanayama G, Hudson JI, DeLuca J, Isaacs S, Baggish A, Weiner R, et al. Prolonged hypogonadism in males following withdrawal from anabolic-androgenic steroids: an under-recognized problem. Addiction (Abingdon, England). 2015;110(5):823–31.

de Souza GL, Hallak J. Anabolic steroids and male infertility: a comprehensive review. BJU Int. 2011;108(11):1860–5.

Pope HG, Kanayama G, Ionescu-Pioggia M, Hudson JI. Anabolic steroid users' attitudes towards physicians. Addiction. 2004;99(9):1189–94.

Hope VD, McVeigh J, Marongiu A, Evans-Brown M, Smith J, Kimergard A, et al. Injection site infections and injuries in men who inject image- and performance-enhancing drugs: prevalence, risks factors, and healthcare seeking. Epidemiol Infection. 2015;143(1):132–40.

Bates G, McVeigh J. Image and performance enhancing drugs - 2015 survey results. Liverpool: Centre for Public Health; 2016.

Brooks JHM, Ahmad I, Easton G. 10-minute consultation anabolic steroid use. Bmj-Brit Med J. 2016;355.

Kanayama G, Boynes M, Hudson JI, Field AE, Pope HG. Anabolic steroid abuse among teenage girls: an illusory problem? Drug Alcohol Depend. 2007;88(2-3):156–62.

Korkia P, Lenehan P, McVeigh J. Non-medical use of androgens among women. J Perform Enhanc Drugs. 1996;1(2):71.

Ip EJ, Barnett MJ, Tenerowicz MJ, Kim JA, Wei H, Perry PJ. Women and anabolic steroids: an analysis of a dozen users. Clin J Sport Med. 2010;20(6):475–81.

Hoberman J. Dopers in uniform: Police officers' use of anabolic steroids in the United States. In: Møller V, Waddington I, Hoberman JM, Møller V, Waddington I, Hoberman JM, editors. Routledge handbook of drugs and sport. New York, NY, US: Routledge/Taylor & Francis Group; 2015. p. 439–52.

Chapter   Google Scholar  

Turvey BE, Crowder S. Anabolic steroid abuse in public safety personnel: a forensic manual. San Diego: Academic Press; 2015.

Humphrey KR, Decker KP, Goldberg L, G. PH, Green GA. Anabolic steroid use and abuse by police officers: policy & prevention. The Police Chief. 2008;LXXV.

Waterhouse J. Rise in soldiers testing positive for steroid use. Newsbeat [Internet]. 2014 25/3/2019.

Bolding G, Sherr L, Maguire M, Elford J. HIV risk behaviours among gay men who use anabolic steroids. Addiction (Abingdon, England). 1999;94(12):1829–35.

Bolding G, Sherr L, Elford J. Use of anabolic steroids and associated health risks among gay men attending London gyms. Addiction. 2002;97(2):195–203.

Cornford CS, Kean J, Nash A. Anabolic-androgenic steroids and heroin use: a qualitative study exploring the connection. Int J Drug Policy. 2014;25(5):928–30.

Kanayama G, Pope HG, Hudson JI. Associations of anabolic-androgenic steroid use with other behavioral disorders: an analysis using directed acyclic graphs. Psychol Med. 2018:1–8.

Salinas M, Floodgate W, Ralphs R. Polydrug use and polydrug markets amongst image and performance enhancing drug users: Implications for harm reduction interventions and drug policy. Int J Drug Policy. 2019;67:43–51.

Kimergard A. A qualitative study of anabolic steroid use amongst gym users in the United Kingdom: motives, beliefs and experiences. J Subst Use. 2015;20(4):288–94.

Taylor WN. Anabolic steroids and the athlete. 2 ed. Jefferson, NC: MacFarland and Company Inc. Publishers; 2001.

Pampel FC. Drugs and Sport. New York: Facts on File; 2007.

Download references

Acknowledgements

Not applicable.

No funding was received to support this review.

Author information

Authors and affiliations.

Public Health Institute, Liverpool John Moores University, Liverpool, England

Geoff Bates & Marie-Claire Van Hout

MRC/CSO SPHSU, University of Glasgow, Glasgow, Scotland

Joseph Tay Wee Teck

Department of Sociology, Manchester Metropolitan University, Manchester, England

Jim McVeigh

You can also search for this author in PubMed   Google Scholar

Contributions

GB managed the review and lead protocol development, evidence search, article screening, data extraction and data synthesis. GB drafted the article methodology and result sections. MCVH provided methodological and topic expertise and helped to shape the review through supporting the development of the protocol and search strategy. MCVH screened a proportion of articles and commented on findings and written drafts. JT provided medical expertise supporting the development of the review and presentation of data, checked data extraction, and commented on written drafts. JMV provided topic expertise and helped to shape the review through supporting the development of the protocol, data extraction, data synthesis and presentation of findings. JMV drafted the article introduction and discussion sections. All authors read and approved the final manuscript.

Corresponding author

Correspondence to Geoff Bates .

Ethics declarations

Ethics approval and consent to participate, consent for publication, competing interests.

The authors declare that they have no competing interests.

Additional information

Publisher’s note.

Springer Nature remains neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Supplementary information

Additional file 1..

Data extraction tables. The data extraction tables contain the full data extracted from the 109 articles included in the review. This includes participant information, condition requiring treatment, the treatment provided and the outcomes of treatment.

Additional file 2.

Search strategy. The full search strategy used in Medline is provided.

Rights and permissions

Open Access This article is distributed under the terms of the Creative Commons Attribution 4.0 International License ( http://creativecommons.org/licenses/by/4.0/ ), which permits unrestricted use, distribution, and reproduction in any medium, provided you give appropriate credit to the original author(s) and the source, provide a link to the Creative Commons license, and indicate if changes were made. The Creative Commons Public Domain Dedication waiver ( http://creativecommons.org/publicdomain/zero/1.0/ ) applies to the data made available in this article, unless otherwise stated.

Reprints and permissions

About this article

Cite this article.

Bates, G., Van Hout, MC., Teck, J.T.W. et al. Treatments for people who use anabolic androgenic steroids: a scoping review. Harm Reduct J 16 , 75 (2019). https://doi.org/10.1186/s12954-019-0343-1

Download citation

Received : 02 May 2019

Accepted : 21 November 2019

Published : 30 December 2019

DOI : https://doi.org/10.1186/s12954-019-0343-1

Share this article

Anyone you share the following link with will be able to read this content:

Sorry, a shareable link is not currently available for this article.

Provided by the Springer Nature SharedIt content-sharing initiative

  • Drug treatment
  • Health care
  • Behaviour change

Harm Reduction Journal

ISSN: 1477-7517

research articles about steroids

  • Open access
  • Published: 13 March 2024

Exploring the prevalence of anabolic steroid use among men and women resistance training practitioners after the COVID-19 pandemic

  • Rastegar Hoseini   ORCID: orcid.org/0000-0001-8685-2471 1 &
  • Zahra Hoseini   ORCID: orcid.org/0000-0002-7933-2221 1  

BMC Public Health volume  24 , Article number:  798 ( 2024 ) Cite this article

3238 Accesses

8 Altmetric

Metrics details

The COVID-19 pandemic has had a significant impact on individual health and fitness routines globally. Resistance training, in particular, has become increasingly popular among men and women looking to maintain or improve their physical fitness during the pandemic. However, using Anabolic Steroids (AS) for performance enhancement in resistance training has known adverse effects. Thus, this study aimed to explore the prevalence of AS use among men and women resistance training practitioners after the COVID-19 pandemic.

A cross-sectional survey was conducted among 3,603 resistance training practitioners (1,855 men and 1,748 women) in various geographical locations impacted by COVID-19. The participants were asked to complete self-administered questionnaires, which included questions regarding demographic information, training habits, and current or prior usage of AS. The data were analyzed using SPSS statistical software and the chi-square method, with a significance level of ( P  < 0.05).

A total of 3603 men and women resistance training practitioners completed the survey. In the study, 53.05% of men and 41.99% of women used anabolic and androgenic steroids. Of those men who used steroids, 29.47% used Testosterone, while 31.20% of women used Winstrol. Additionally, 50.30% of men used steroids via injection, while 49.05% of women used them orally. According to the study, 49.99% of the participants had 6 to 12 months of experience with resistance training, and 64.25% of them underwent three training sessions per week. The analysis using the χ2 test did not reveal any significant difference between men and women in terms of duration of bodybuilding, frequency per week, and engagement in other activities.

This study shows that a significant proportion of men and women resistance training practitioners used AS, particularly among young adults with limited training experience. Thus, there is a need for targeted education and awareness campaigns to address the hazards of AS use and promote healthy training habits during the COVID-19 pandemic.

Peer Review reports

Introduction

The coronavirus pandemic has caused significant disruption to the daily activities of individuals across the world [ 1 ]. One of the areas of life that has been significantly affected is physical exercise [ 2 ]. With the closure of gyms and other sports facilities and restrictions on outdoor exercise, resistance training practitioners have been forced to adapt to new methods of training to maintain their fitness levels [ 3 ]. This disruption to training habits may have had an impact on Anabolic Steroid (AS) usage among men and women resistance training practitioners [ 4 ]. AS are synthetic substances designed to mimic the effects of natural testosterone in the body [ 5 ]. These substances have numerous applications, including medical treatment for hormonal imbalances and muscle-wasting diseases. However, their abuse in the fitness industry, particularly among bodybuilders and other resistance training practitioners, has become widespread, primarily due to their performance-enhancing effects.

Several previous studies have explored the prevalence of AS consumption in various populations, including athletes, bodybuilders, and fitness enthusiasts [ 6 , 7 ]. These studies have demonstrated that the use of these substances is not limited to men and is consumed among women as well. According to a meta-analysis studying a wide range of samples, such as students, university students, resistance training practitioners, and the general public, the global prevalence of AS consumption was estimated at approximately 3.3% [ 8 ]. About Iran, a prevalence of 0.3% AS consumption among the adult population [ 6 , 9 ]. This percentage increased to 36.66% in 2020, which was primarily investigated among men aged between 18 and 34 years [ 10 ]. Furthermore, the prevalence of AS consumption varies widely across different countries, with rates of up to 25% reported in some cases [ 11 , 12 ]. Recent studies have raised concerns that the Coronavirus Disease 2019 (COVID-19) COVID-19 pandemic might have led to an increased prevalence of AS consumption in the resistance training community [ 4 , 13 ]. The pandemic response measures have forced many people to adjust to new, home-based training methods, with limited availability of gym and fitness facilities. This shift in training patterns may have led to increased use or abuse of AS among resistance training practitioners. However, the novelty of this study lies in its focus on resistance training practitioners, the examination of anabolic steroid use after the COVID-19 pandemic, and the inclusion of both men and women in the study population. To date, no research has explored the prevalence of AS consumption among resistance trainers after the COVID-19 pandemic in the Iranian population. Despite the growing popularity of resistance training during the pandemic, there is limited research specifically examining the prevalence of AS use among individuals engaging in resistance training. This research gap is crucial as it allows us to understand the extent of AS use and its associated risks within this specific population. By addressing this research gap, valuable insights can be provided into the prevalence of AS use and its potential implications for the health and well-being of resistance training practitioners. This study aims to address the research gap by exploring the prevalence of AS use among men and women resistance training practitioners after the COVID-19 pandemic. The novelty of this study lies in its focus on a specific population and its potential to provide valuable insights into AS use within the resistance training community. By linking the research gap to the goal of the article, the existing literature aims to be contributed to and awareness of the hazards of AS use while fostering healthy training habits during the COVID-19 pandemic aims to be promoted. Thus, this study aimed to investigate the prevalence of AS consumption among men and women resistance training practitioners after the COVID-19 pandemic.

Study design and population

The survey was conducted in Kermanshah, Iran, a city with an approximate population of 1 million. The survey obtained information on the number and location of gyms in Kermanshah from the Regional Council of Physical Education in the city between May and July 2023. A total of 356 fitness centers that offered resistance training were identified, out of which 286 centers were included in the study. With a confidence interval of 95% and assuming a p = q = 50% probability, a total of 100 resistance training centers were calculated, with an error margin of 7.9%, and used to estimate the population of resistance training practitioners in the city. The gyms were selected randomly and systematically from the five administrative regions of the city, based on the proportion of the number of gyms in each region. The gym management was contacted and explained about the study before obtaining their consent. Individuals aged 18 years and above, training for resistance exercise during morning, afternoon, or night hours were identified in each center. On average, 568 resistance training practitioners were identified per gym. A total number of 4,198 individuals were selected proportionately from each gym based on the number of resistance training practitioners, with a sampling error of 1.25% and a confidence interval of 95%. After screening out incomplete responses, 3603 individuals (1,855 men and 1,748 women) were included in the final analyses. At the commencement of the questionnaires, the participants were provided with information regarding the objectives of the study, and the confidential handling of data, and participants completed the consent form. Also, all educated participants and the legal guardians of illiterate participants were asked to complete the written informed consent at the beginning of the study. The study was conducted in adherence to the seventh and current modification (World Medical Association, 2013) of the Declaration of Helsinki. All experimental protocols were approved by the Committee of Research in Public Sports Board, Kermanshah, Iran.

Data collection

A self-administered questionnaire, consisting of 32 questions, was devised through a scholarly literature review of relevant articles [ 14 , 15 ]. The Questionnaire included the following variables: gender, age, profession, marital status, schooling, socioeconomic status, practice time of resistance training, duration, and purpose of training, nutritional monitoring, use of supplements, and use of AS. The questionnaire underwent a validation process, which determined its clarity, content, and construct indices. The questionnaire’s construction and content were evaluated and validated by professional health practitioners, while the clarity aspect was reviewed by individuals sharing the same traits, including class, age, and lifestyle of the intended research population. A pilot study was conducted to assess the questionnaire’s feasibility for use among the target populace.

To standardize the approach and application of the questionnaire, a pre-training session was conducted with the researchers. Following the pre-training, a pilot study with 40 individuals was carried out at the Kani Gym, which was not included in the survey data. Data collection was conducted throughout the working day by researchers positioned at the entrance of the gym and dressed in uniform to be easily identified. To approach participants, they were explained the research purpose, either at the beginning or end of their workout. Participants who agreed to participate in the study signed an informed consent form. The researchers provided clarification for any queries or ambiguities related to the questionnaire before allowing the participants to complete the form independently, without interference.

Statistical analysis

Statistical analysis in this study was performed using SPSS statistical software (version 21; SPSS Inc., Chicago, IL, USA) with a significance level of P  < 0.05. The normality of distribution was assessed with the Kolmogorov-Smirnov test. Both descriptive statistics, including mean, standard deviation, and percentage, and deductive statistics the Chi-square method, were utilized for analysis.

A total of 3,603 (1,855 men and 1,748 women) resistance training practitioners from various regions participated in the survey. A total of (number) participants took part in this study, of which 46% were aged between 18 and 29 years old (46.15% men and 45.08% women), 34.08% were aged between 30 and 44 years old (34.17% men and 33.98% women), 14.13% were aged between 45 and 59 years old (14.33% men and 13.90% women), and 6.16% were aged over 60 years old (5.34% men and 7.04% women). Also, 27.59% were single (30.02% men and 25.06% women), and 72.41% were married (69.98% men and 74.94% women). Furthermore, 0.72% of the participants were illiterate (0.81% men and 0.63% women), while 19.18% had a bachelor’s degree (18.01% men and 20.42% women). The majority of the participants, 80.10%, were university-educated (18.18% men and 78.95% women). Regarding employment status, 44.93% of the participants were employed (68.46% men and 19.97% women), 30.01% were enrolled as students (27.01% men and 33.18% women), and 25.06% were unemployed (4.53% men and 46.85% women). Only a small proportion of the sample, 3.99%, reported being smokers (4.96% men and 2.98% women), while 13.60% of the participants were hospitalized due to COVID-19 (15.94% men and 12.08% women) (Table  1 ).

The χ2 test was conducted to examine potential gender differences for all of these variables. The results demonstrated that employment status was the only variable with a statistically significant gender difference, with a higher proportion of men being employed compared to women ( p  < 0.05). No significant differences were found in the distribution of other variables based on gender (Table  1 ).

In this study, 49.99% of the participants had 6 to 12 months of experience with resistance training, and 64.25% of them underwent three training sessions per week. The results of analysis using the χ2 test revealed no significant difference in the duration of bodybuilding, frequency per week, and engagement in other activities between men and women. However, a significant difference in the purpose of performing resistance exercises was found, with 51.37% of men attending the gym for hypertrophy and 55.94% of women attending for weight loss. These findings suggest that men and women exhibit similar patterns of engagement in resistance training, but their motivations for doing so may differ (Table  2 ).

Table  3 presents the results showing that 53.05% of men and 41.99% of women used anabolic and androgenic steroids, with consumption methods differing between genders; 50.30% of men used it via injection, while 49.05% of women used it orally. The results of the χ2 test demonstrated a significant difference in the amount and consumption method of anabolic and androgenic steroid use between men and women. Furthermore, it was found that Testosterone was used by 29.47% of men, while Winstrol was used by 31.20% of women. These findings provide insight into gender-based differences in the use of anabolic and androgenic steroids and suggest that gender-specific strategies may be necessary to address this practice.

Resistance training is a popular form of exercise that has gained significant attention in recent years due to its numerous health benefits. The current study aims to investigate the exploring the prevalence of AS use among men and women resistance training practitioners after the COVID-19 pandemic. The results of the present study revealed a sample of 3,603 individuals, with approximately equal representation of men and women (51.42% versus 48.58%, respectively). The age distribution of participants showed that resistance training is popular among young adults, with 46% of participants aged between 18 and 29 years old. The originality of our study lies in its comprehensive analysis of the characteristics and gender differences of resistance training practitioners from various regions. This age range was nearly uniformly split between men and women; this finding is significant as it indicates that resistance training is equally popular among both genders, particularly among young adults, with 46% of participants aged between 18 and 29 years old. The findings of the present study indicate that the majority of participants were university-educated, which is consistent with previous research demonstrating that a higher level of education is associated with higher participation in exercise and sports [ 16 , 17 ]. Additionally, the results showed that the majority of participants were married, which suggests that resistance training may be a popular form of exercise for those with responsibilities such as marriage and children. In terms of employment status, these results suggest that there is a gender difference, with a higher proportion of men being employed compared to women. This finding is consistent with previous research demonstrating that men are more likely to be employed than women [ 18 ], and may reflect societal norms and gender roles. Finally, this study revealed a low prevalence of smoking among resistance training practitioners, which is encouraging given the detrimental health effects of smoking. However, a relatively high rate of hospitalization due to COVID-19 was found among the sample, which could be attributed to increased exposure to the virus in fitness facilities. This emphasizes the importance of implementing and promoting preventive measures to mitigate the risk of COVID-19 transmission in fitness facilities Overall, this study contributes to a better understanding of the characteristics and gender differences of resistance training practitioners from various regions. These findings suggest that resistance training is popular among both genders, particularly among young adults, and can be practiced by individuals with diverse educational and marital backgrounds. This is significant as it broadens our understanding of the demographic profile of resistance training practitioners. However, future research should investigate the motivations and expectations of resistance training practitioners, as well as the factors that influence how this form of exercise is adopted and maintained over time.

Also, the results of this study showed that a high percentage of participants had between 6 and 12 months of resistance training experience, and the majority underwent three weekly training sessions. Furthermore, these results showed no significant differences in the length of bodybuilding, frequency per week, and engagement in other activities between genders. These findings suggest that men and women exhibit similar exercise habits in resistance training. However, a significant difference in motivations between genders was found. The gym was attended by over half of the men (51.37%) for hypertrophy, while over half of the women (55.94%) attended for weight loss. Thus, these findings indicate that the motivations behind resistance training may differ between genders. It is worth noting that despite these differences in motivations, both men and women seem to have an equal level of engagement in resistance training. These findings have important implications for resistance training interventions. For example, hypertrophy may be less of a motivator for women in resistance training, while emphasizing weight loss may be more effective in increasing women’s participation in resistance training programs. However, more research is needed to determine the most effective ways to motivate men and women differently in resistance training interventions.

However, this study highlights the importance of considering gender differences in motivations for resistance training. While men and women exhibit similar exercise habits, their motivations may differ significantly. These findings may have important implications for resistance training interventions aimed at increasing participation and adherence in both men and women. Further research is needed to identify effective methods of motivating men and women in resistance training interventions. These results suggest that there are significant differences between men and women in terms of both the prevalence and consumption method of steroid use. Specifically, 53.05% of men and 41.99% of women reported using anabolic and androgenic steroids. This finding is significant as it highlights the need for gender-specific interventions to address steroid use. The size of the sample, participants, and gyms used in the literature varied considerably. For instance, a study conducted in Germany approximately 15 years ago involved 113 gyms and 621 individuals and reported a prevalence of AS use of 13.5% [ 14 ]. In Stockholm, Sweden, the prevalence was 3.8% with 64 gyms and 1746 individuals [ 19 ]. On the other hand, in Al-Ain, United Arab Emirates, the prevalence was 22.1% with 18 gyms and 154 individuals [ 20 ]. However, some studies had smaller sample sizes; for example, a study in El Paso, United States, evaluated three gyms and 516 individuals, revealing a prevalence of 11.0% [ 21 ]. Several factors, such as the sample distribution, the regional characteristics, and the individual characteristics of the samples, could have contributed to the variability in the prevalence of AS use among these studies. For instance, a study in the Netherlands that involved 92 gyms and 718 individuals reported a prevalence of AS use of 1% [ 22 ]. These findings are consistent with previous studies that have found that men are more likely to use anabolic and androgenic steroids than women [ 23 , 24 , 25 ]. The reason for lower consumption of AS among women is often due to their desire not to become excessively muscular or develop male characteristics [ 26 ]. On the other hand, men use AS not only to attain their desired body but also to gain status, admiration, and popularity in their social circle [ 27 ]. Furthermore, using AS helps them to be recognized and accepted by their peers [ 28 ]. These results also revealed important gender-based differences in the methods of steroid consumption, with 50.30% of men using intravenous injection and 49.05% of women using oral consumption. These differences may be due to various factors such as differences in physiology, availability, and perceived effectiveness. Of particular importance is the use of Testosterone by men and Winstrol by women, which were found to be the most commonly used steroids among the respective genders. While the reasons for these gender-based differences are unclear, they may reflect differences in physique ideals or perceived benefits or side effects.

Strength and limitations

These findings have important implications for the development of interventions to address anabolic and androgenic steroid use. The fact that gender-based differences were found in both the prevalence and consumption method of steroid use highlights the need for gender-specific interventions that take into account the unique factors driving steroid use among men and women. For instance, interventions targeting men may need to focus on reducing intravenous injection use, while interventions targeting women may need to focus on reducing oral consumption. While this study provides valuable insights into gender-specific differences in anabolic and androgenic steroid use, it is important to note that the sample used in this study was limited to a specific population and may not be representative of the broader population. Additionally, self-reported data are subject to social desirability bias and may not reflect the true prevalence of anabolic and androgenic steroid use. Future studies should aim to replicate these findings with larger, more representative samples, and employ more objective measures of steroid use such as biological markers.

In conclusion, our study significantly contributes to the understanding of resistance training practices among both genders, particularly among young adults. It underscores that resistance training is not limited to a specific demographic but is embraced by individuals with diverse educational and marital backgrounds. A key finding of our research is the distinct motivations for resistance training between men and women, with hypertrophy being a primary driver for men and weight loss for women. This divergence in motivations necessitates the development of gender-specific resistance training interventions to enhance participation and adherence. Furthermore, our study unveils critical gender differences in the prevalence and methods of anabolic steroid (AS) use. Men reported higher usage rates and a preference for intravenous injection, while women predominantly opted for oral consumption. These findings are pivotal, highlighting the need for gender-specific considerations when designing interventions and educational programs to address AS use among resistance training practitioners. Our research, therefore, provides valuable insights that can guide the development of more effective, gender-tailored strategies in the field of resistance training.

Future studies

In future studies, several suggestions can be considered to enhance the straightness of research on anabolic steroid use among resistance training practitioners. First, adopting a longitudinal approach would provide valuable insights into the changes in steroid use over time post-pandemic, identifying shifts in prevalence, patterns, and influencing factors. Also, supplementing quantitative data with in-depth interviews would offer a deeper understanding of motivations, perceptions, and experiences related to steroid use. Moreover, comparing steroid use across different training settings, such as home-based workouts, commercial gyms, or community centers, would allow for a comparison of prevalence rates and factors associated with steroid use within these environments. Additionally, exploring psychological factors such as body image dissatisfaction, social pressure, or self-esteem would provide a more comprehensive understanding of the motivations behind steroid use. Lastly, investigating the effectiveness of educational initiatives aimed at raising awareness and assessing their impact on attitudes, knowledge, and behaviors related to steroid use would assist in designing evidence-based preventive strategies. Implementing these suggestions would contribute to a more comprehensive and robust understanding of anabolic steroid use among resistance training practitioners.

Data availability

The datasets used and/or analyzed during the current study are available from the corresponding author upon reasonable request.

Abbreviations

Anabolic Steroid

Coronavirus Disease 2019

Rahim HA, Hoseini R, Hoseini Z, Abbas EN, Kareem DA. Health-related factors of the Iraqi adult population during the 2020 COVID-19 pandemic: physical activity, eating behavior, quality of life, general health, and mood states cross-talk. BMC Public Health. 2023;23(1):1046.

Article   PubMed   PubMed Central   Google Scholar  

Cairney J, Dudley D, Kwan M, Bulten R, Kriellaars D. Physical literacy, physical activity, and health: toward an evidence-informed conceptual model. Sports Med. 2019;49:371–83.

Article   PubMed   Google Scholar  

Dor-Haim H, Katzburg S, Revach P, Levine H, Barak S. The impact of COVID-19 lockdown on physical activity and weight gain among active adult population in Israel: a cross-sectional study. BMC Public Health. 2021;21(1):1–10.

Article   Google Scholar  

Zoob Carter BN, Boardley ID, Van de Ven K. The impact of the COVID-19 pandemic on male strength athletes who use non-prescribed anabolic-androgenic steroids. Front Psychiatry. 2021;12:636706.

Barany QL, Sofi A, Al-Barwary MK. Neurosteroid, natural and anabolic steroids: physiological, immunological and histopathological study on Diabetic albino rats. Egypt J Veterinary Sci. 2023;54(5):863–81.

Google Scholar  

Arazi H, Hosseini R. The prevalence of anabolic-androgenic steroids abuse, knowledge and attitue of their side effects, and attitude toward them among the bodybuilding athletes in Rasht. J Guilan Univ Med Sci. 2011;20(80):34–41.

Pereira E, Moysés SJ, Ignácio SA, Mendes DK, Silva DSD, Carneiro E, et al. Prevalence and profile of users and non-users of anabolic steroids among resistance training practitioners. BMC Public Health. 2019;19:1–8.

Article   CAS   Google Scholar  

Sagoe D, Molde H, Andreassen CS, Torsheim T, Pallesen S. The global epidemiology of anabolic-androgenic steroid use: a meta-analysis and meta-regression analysis. Ann Epidemiol. 2014;24(5):383–98.

Fayyazi Bordbar MR, Abdollahian E, Samadi R, Dolatabadi H. Frequency of use, awareness, and attitudes toward side effects of anabolic–androgenic steroids consumption among male medical students in Iran. Subst Use Misuse. 2014;49(13):1751–8.

Hoseini M, Yousefi B, Khazaei AA. The prevalence of anabolic-androgenic steroids abuse, knowledge and attitude of their side effects, and attitude toward them among the female bodybuilding athletes in Kermanshah. J Adv Biomedical Sci. 2020;10(3):2439–47.

de Zeeuw TI, Brunt TM, van Amsterdam J, van de Ven K, van den Brink W. Anabolic androgenic steroid use patterns and steroid use disorders in a sample of male gym visitors. Eur Addict Res. 2023;29(2):99–108.

Hammoud S, van den Bemt BJ, Jaber A, Kurdi M. Chronic anabolic androgenic steroid administration reduces global longitudinal strain among off-cycle bodybuilders. Int J Cardiol. 2023;381:153–60.

Crisp P, Sims J. COVID-19 and anabolic-androgenic steroids (AAS) as immunosuppresors: is it time to revisit Government Policy and Governance arrangements on AAS? Archives Sports Med. 2020;4(2):245–6.

Striegel H, Simon P, Frisch S, Roecker K, Dietz K, Dickhuth H-H, et al. Anabolic ergogenic substance users in fitness-sports: a distinct group supported by the health care system. Drug Alcohol Depend. 2006;81(1):11–9.

da Silva PR, Machado Júnior LC, Figueiredo VC, Cioffi AP, Prestes MC, Czepielewski MA. Prevalence of the use of anabolic agents among strength training apprentices in Porto Alegre, RS. Arquivos Brasileiros De Endocrinologia Metabologia. 2007;51:104–10.

Hebert JJ, Møller NC, Andersen LB, Wedderkopp N. Organized sport participation is associated with higher levels of overall health-related physical activity in children (CHAMPS study-DK). PLoS ONE. 2015;10(8):e0134621.

Fox CK, Barr-Anderson D, Neumark‐Sztainer D, Wall M. Physical activity and sports team participation: associations with academic outcomes in middle school and high school students. J Sch Health. 2010;80(1):31–7.

Helfand BK, Webb M, Gartaganis SL, Fuller L, Kwon C-S, Inouye SK. The exclusion of older persons from vaccine and treatment trials for coronavirus disease 2019—missing the target. JAMA Intern Med. 2020;180(11):1546–9.

Article   CAS   PubMed   PubMed Central   Google Scholar  

Leifman H, Rehnman C, Sjöblom E, Holgersson S. Anabolic androgenic steroids—use and correlates among gym users—an assessment study using questionnaires and observations at gyms in the Stockholm region. Int J Environ Res Public Health. 2011;8(7):2656–74.

Al-Falasi O, Al-Dahmani K, Al-Eisaei K, Al-Ameri S, Al-Maskari F, Nagelkerke N, et al. Knowledge, attitude and practice of anabolic steroids use among gym users in Al-Ain district, United Arab Emirates. Open Sports Med J. 2008;2:75–81.

STREET C. Steroids from Mexico: educating the strength and conditioning community. J Strength Conditioning Res. 2000;14(3):289–94.

Stubbe JH, Chorus AM, Frank LE, de Hon O, van der Heijden PG. Prevalence of use of performance enhancing drugs by fitness centre members. Drug Test Anal. 2014;6(5):434–8.

Article   CAS   PubMed   Google Scholar  

Lundholm L, Käll K, Wallin S, Thiblin I. Use of anabolic androgenic steroids in substance abusers arrested for crime. Drug Alcohol Depend. 2010;111(3):222–6.

Ip EJ, Barnett MJ, Tenerowicz MJ, Kim JA, Wei H, Perry PJ. Women and anabolic steroids: an analysis of a dozen users. Clin J Sport Med. 2010;20(6):475–81.

Santos AF, Mendonça PMH, Santos LA, Silva NF, Tavares JKL. Anabolic steroids: concepts according to muscular activity practisers in Aracaju (SE). Psicologia em Estudo. 2006;11:371–80.

Kanayama G, Hudson JI, Pope HG Jr. Illicit anabolic–androgenic steroid use. Horm Behav. 2010;58(1):111–21.

Iriart JAB, Chaves JC, Orleans RG. Body cult and use of anabolic steroids by bodybuilders. Cadernos De saúde Publica. 2009;25:773–82.

Hilkens L, Cruyff M, Woertman L, Benjamins J, Evers C. Social media, body image and resistance training: creating the perfect ‘me’with dietary supplements, anabolic steroids and SARM’s. Sports medicine-open. 2021;7(1):1–13.

Download references

Acknowledgements

We would like to thank the subjects for their willing participation in this study.

The authors declared that the research did not receive any financial grants.

Author information

Authors and affiliations.

Department of Exercise Physiology, Faculty of Sport Sciences, Razi University, Kermanshah, P.O. Box. 6714414971, Iran

Rastegar Hoseini & Zahra Hoseini

You can also search for this author in PubMed   Google Scholar

Contributions

RH: contributed to the study conception, design, investigation, data analysis, and writing of the manuscript. ZH: contributed to the data acquisition, interpretation, data analysis, and revision of the manuscript. All authors have approved the final version of the manuscript and agreed to be accountable for all aspects of the study.

Corresponding author

Correspondence to Rastegar Hoseini .

Ethics declarations

Ethics approval and consent to participate.

Participants were provided with detailed information regarding the research purpose, procedures, risks, and benefits. They provided informed consent to participate in the study and were informed that they could withdraw from the study at any time without any penalty. The confidentiality and anonymity of participants were maintained throughout the study, and the data collected were used for research purposes only. All research procedures were conducted in compliance with ethical principles as outlined in the Declaration of Helsinki, the applicable regulations, and the guidelines provided by the IRB/IEC. The authors confirm that all methods were performed following relevant guidelines and regulations. All experimental protocols were approved by the ethics committee of Research in Public Sports Board, Kermanshah, Iran. Also, all educated participants and the legal guardian of illiterate participants were asked to complete the written informed consent at the beginning of the study.

Consent for publication

Not applicable.

Competing interests

The authors declare no competing interests.

Additional information

Publisher’s note.

Springer Nature remains neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Rights and permissions

Open Access This article is licensed under a Creative Commons Attribution 4.0 International License, which permits use, sharing, adaptation, distribution and reproduction in any medium or format, as long as you give appropriate credit to the original author(s) and the source, provide a link to the Creative Commons licence, and indicate if changes were made. The images or other third party material in this article are included in the article’s Creative Commons licence, unless indicated otherwise in a credit line to the material. If material is not included in the article’s Creative Commons licence and your intended use is not permitted by statutory regulation or exceeds the permitted use, you will need to obtain permission directly from the copyright holder. To view a copy of this licence, visit http://creativecommons.org/licenses/by/4.0/ . The Creative Commons Public Domain Dedication waiver ( http://creativecommons.org/publicdomain/zero/1.0/ ) applies to the data made available in this article, unless otherwise stated in a credit line to the data.

Reprints and permissions

About this article

Cite this article.

Hoseini, R., Hoseini, Z. Exploring the prevalence of anabolic steroid use among men and women resistance training practitioners after the COVID-19 pandemic. BMC Public Health 24 , 798 (2024). https://doi.org/10.1186/s12889-024-18292-5

Download citation

Received : 28 July 2023

Accepted : 05 March 2024

Published : 13 March 2024

DOI : https://doi.org/10.1186/s12889-024-18292-5

Share this article

Anyone you share the following link with will be able to read this content:

Sorry, a shareable link is not currently available for this article.

Provided by the Springer Nature SharedIt content-sharing initiative

  • Resistance training
  • Young adults
  • Awareness campaigns

BMC Public Health

ISSN: 1471-2458

research articles about steroids

Information

  • Author Services

Initiatives

You are accessing a machine-readable page. In order to be human-readable, please install an RSS reader.

All articles published by MDPI are made immediately available worldwide under an open access license. No special permission is required to reuse all or part of the article published by MDPI, including figures and tables. For articles published under an open access Creative Common CC BY license, any part of the article may be reused without permission provided that the original article is clearly cited. For more information, please refer to https://www.mdpi.com/openaccess .

Feature papers represent the most advanced research with significant potential for high impact in the field. A Feature Paper should be a substantial original Article that involves several techniques or approaches, provides an outlook for future research directions and describes possible research applications.

Feature papers are submitted upon individual invitation or recommendation by the scientific editors and must receive positive feedback from the reviewers.

Editor’s Choice articles are based on recommendations by the scientific editors of MDPI journals from around the world. Editors select a small number of articles recently published in the journal that they believe will be particularly interesting to readers, or important in the respective research area. The aim is to provide a snapshot of some of the most exciting work published in the various research areas of the journal.

Original Submission Date Received: .

  • Active Journals
  • Find a Journal
  • Proceedings Series
  • For Authors
  • For Reviewers
  • For Editors
  • For Librarians
  • For Publishers
  • For Societies
  • For Conference Organizers
  • Open Access Policy
  • Institutional Open Access Program
  • Special Issues Guidelines
  • Editorial Process
  • Research and Publication Ethics
  • Article Processing Charges
  • Testimonials
  • Preprints.org
  • SciProfiles
  • Encyclopedia

healthcare-logo

Article Menu

  • Subscribe SciFeed
  • Recommended Articles
  • PubMed/Medline
  • Google Scholar
  • on Google Scholar
  • Table of Contents

Find support for a specific problem in the support section of our website.

Please let us know what you think of our products and services.

Visit our dedicated information section to learn more about MDPI.

JSmol Viewer

Adverse effects of anabolic-androgenic steroids: a literature review.

research articles about steroids

1. Introduction

2. phisiology of aass, 3. pathophysiology of aass, 4. aas use and adverse effects, 4.1. autopsy findings, 4.2. brain and behavior, 4.3. cardiovascular system, 4.5. urinary system, 4.6. muscoloskeletal system, 4.7. reproductive system, 4.8. hematologic consequences, 4.9. aass and cancer, 5. conclusions, author contributions, institutional review board statement, informed consent statement, data availability statement, acknowledgments, conflicts of interest.

  • Roman, M.; Roman, D.L.; Ostafe, V.; Ciorsac, A.; Isvoran, A. Computational assessment of pharmacokinetics and biological effects of some anabolic and androgen steroids. Pharm. Res. 2018 , 35 , 41. [ Google Scholar ] [ CrossRef ] [ PubMed ]
  • Patanè, F.G.; Liberto, A.; Maglitto, A.N.M.; Malandrino, P.; Esposito, M.; Amico, F.; Cocimano, G.; Li Rosi, G.; Condorelli, D.; Di Nunno, N.; et al. Nandrolone Decanoate: Use, Abuse and Side Effects. Medicina 2020 , 56 , 606. [ Google Scholar ] [ CrossRef ] [ PubMed ]
  • Mullen, C.; Whalley, B.J.; Schifano, F.; Baker, J.S. Anabolic androgenic steroid abuse in the United Kingdom: An update. Br. J. Pharmacol. 2020 , 177 , 2180–2198. [ Google Scholar ] [ CrossRef ]
  • Basaria, S. Androgen abuse in athletes: Detection and consequences. J. Clin. Endocrinol. Metab. 2010 , 95 , 1533–1543. [ Google Scholar ] [ CrossRef ] [ PubMed ] [ Green Version ]
  • El Osta, R.; Almont, T.; Diligent, C.; Hubert, N.; Eschwège, P.; Hubert, J. Anabolic steroids abuse and male infertility. Basic Clin. Androl. 2016 , 26 , 2. [ Google Scholar ] [ CrossRef ] [ Green Version ]
  • Joseph, J.F.; Parr, M.K. Synthetic androgens as designer supplements. Curr. Neuropharmacol. 2015 , 13 , 89–100. [ Google Scholar ] [ CrossRef ] [ Green Version ]
  • Hakansson, A.; Mickelsson, K.; Wallin, C.; Berglund, M. Anabolic androgenic steroids in the general population: User characteristics and associations with substance use. Eur. Addict. Res. 2012 , 18 , 83–90. [ Google Scholar ] [ CrossRef ]
  • Kanayama, G.; Kaufman, M.J.; Pope, H.G., Jr. Public health impact of androgens. Curr. Opin. Endocrinol. Diabetes Obes. 2018 , 25 , 218. [ Google Scholar ] [ CrossRef ]
  • Rahnema, C.D.; Crosnoe, L.E.; Kim, E.D. Designer steroids-over-the-counter supplements and their androgenic component: Review of an increasing problem. Andrology 2015 , 3 , 150–155. [ Google Scholar ] [ CrossRef ] [ Green Version ]
  • Kanayama, G.; Hudson, J.I.; Pope, H.G., Jr. Illicit anabolic–androgenic steroid use. Horm. Behav. 2010 , 58 , 111–121. [ Google Scholar ] [ CrossRef ] [ Green Version ]
  • Anawalt, B.D. Diagnosis and management of anabolic androgenic steroid use. J. Clin. Endocrinol. Metab. 2019 , 104 , 2490–2500. [ Google Scholar ] [ CrossRef ] [ PubMed ]
  • Kicman, A.T. Pharmacology of anabolic steroids. Br. J. Pharmacol. 2008 , 154 , 502–521. [ Google Scholar ] [ CrossRef ]
  • Sessa, F.; Salerno, M.; Di Mizio, G.; Bertozzi, G.; Messina, G.; Tomaiuolo, B.; Pisanelli, D.; Maglietta, F.; Ricci, P.; Pomara, C. Anabolic androgenic steroids: Searching new molecular biomarkers. Front. Pharmacol. 2018 , 9 , 1321. [ Google Scholar ] [ CrossRef ] [ PubMed ]
  • Alsiö, J.; Birgner, C.; Björkblom, L.; Isaksson, P.; Bergström, L.; Schiöth, H.B.; Lindblom, J. Impact of nandrolone decanoate on gene expression in endocrine systems related to the adverse effects of anabolic androgenic steroids. Basic Clin. Pharmacol. Toxicol. 2009 , 105 , 307–314. [ Google Scholar ] [ CrossRef ] [ PubMed ]
  • Cheung, A.S.; Grossmann, M. Physiological basis behind ergogenic effects of anabolic androgens. Mol. Cell. Endocrinol. 2018 , 464 , 14–20. [ Google Scholar ] [ CrossRef ]
  • Fortunato, R.S.; Marassi, M.P.; Chaves, E.A.; Nascimento, J.H.; Rosenthal, D.; Carvalho, D.P. Chronic administration of anabolic androgenic steroid alters murine thyroid function. Med. Sci. Sports Exerc. 2006 , 38 , 256–261. [ Google Scholar ] [ CrossRef ] [ PubMed ]
  • Nieschlag, E.; Vorona, E. Doping with anabolic androgenic steroids(AAS): Adverse effects on non-reproductive organs and functions. Rev. Endocr. Metab. Disord. 2015 , 16 , 199–211. [ Google Scholar ] [ CrossRef ]
  • Arazi, H.; Mohammadjafari, H.; Asadi, A. Use of anabolic androgenic steroids produces greater oxidative stress responses to resistance exercise in strength-trained men. Toxicol. Rep. 2017 , 4 , 282–286. [ Google Scholar ] [ CrossRef ]
  • Frankenfeld, S.P.; Oliveira, L.P.; Ortenzi, V.H.; Rego-Monteiro, I.C.; Chaves, E.A.; Ferreira, A.C.; Leitao, A.C.; Carvalho, D.P.; Fortunato, R.S. The anabolic androgenic steroid nandrolone decanoate disrupts redox homeostasis in liver, heart and kidney of male Wistar rats. PLoS ONE 2014 , 9 , e102699. [ Google Scholar ] [ CrossRef ]
  • Frankenfeld, S.P.; de Oliveira, L.P.; Ignacio, D.L.; Coelho, R.G.; Mattos, M.N.; Ferreira, A.C.; Carvalho, D.P.; Fortunato, R.S. Nandrolone decanoate inhibits gluconeogenesis and decreases fasting glucose in Wistar male rats. J. Endocrinol. 2014 , 220 , 143–153. [ Google Scholar ] [ CrossRef ] [ Green Version ]
  • Chaves, E.A.; Pereira, P.P., Jr.; Fortunato, R.S.; Masuda, M.O.; de Carvalho, A.C.; de Carvalho, D.P.; Oliveira, M.F.; Nascimento, J.H. Nandrolone decanoate impairs exercise-induced cardioprotection: Role of antioxidant enzymes. J. Steroid Biochem. Mol. Biol. 2006 , 99 , 223–230. [ Google Scholar ] [ CrossRef ] [ PubMed ]
  • Cerretani, D.; Neri, M.; Cantatore, S.; Ciallella, C.; Riezzo, I.; Turillazzi, E.; Fineschi, V. Looking for organ damages due to anabolic-androgenic steroids(AAS): Is oxidative stress the culprit? Mini Rev. Org. Chem. 2013 , 10 , 393–399. [ Google Scholar ] [ CrossRef ]
  • Turillazzi, E.; Neri, M.; Cerretani, D.; Cantatore, S.; Frati, P.; Moltoni, L.; Busardò, F.P.; Pomara, C.; Riezzo, I.; Fineschi, V. Lipid peroxidation and apoptotic response in rat brain areas induced by long-term administration of nandrolone: The mutual crosstalk between ROS and NF-kB. J. Cell. Mol. Med. 2016 , 20 , 601–612. [ Google Scholar ] [ CrossRef ] [ PubMed ]
  • Basile, J.R.; Binmadi, N.O.; Zhou, H.; Yang, Y.; Paoli, A.; Proia, P. Supraphysiological doses of performance enhancing anabolic-androgenic steroids exert direct toxic effects on neuron-like cells. Front. Cell Neurosci. 2013 , 7 , 69. [ Google Scholar ] [ CrossRef ] [ Green Version ]
  • Joukar, S.; Vahidi, R.; Farsinejad, A.; Asadi-Shekaari, M.; Shahouzehi, B. Ameliorative effects of endurance exercise with two different intensities on nandrolone decanoate-induced neurodegeneration in rats: Involving redox and apoptotic systems. Neurotox. Res. 2017 , 32 , 41–49. [ Google Scholar ] [ CrossRef ]
  • D’Errico, S.; Di Battista, B.; Di Paolo, M.; Fiore, C.; Pomara, C. Renal heat shock proteins over-expression due to anabolic androgenic steroids abuse. Mini Rev. Med. Chem. 2011 , 11 , 446–450. [ Google Scholar ] [ CrossRef ]
  • Chrostowski, K.; Kwiatkowska, D.; Pokrywka, A.; Stanczyk, D.; Wójcikowska-Wójcik, B.; Grucza, R. Renin-angiotensin-aldosterone system in bodybuilders using supraphysiological doses of anabolic-androgenic steroids. Biol. Sport 2011 , 28 , 11. [ Google Scholar ] [ CrossRef ] [ Green Version ]
  • D’Andrea, A.; Caso, P.; Salerno, G.; Scarafile, R.; De Corato, G.; Mita, C.; Di Salvo, G.; Severino, S.; Cuomo, S.; Liccardo, B.; et al. Left ventricular early myocardial dysfunction after chronic misuse of anabolic androgenic steroids: A Doppler myocardial and strain imaging analysis. Br. J. Sports Med. 2007 , 41 , 149–155. [ Google Scholar ] [ CrossRef ]
  • Lopes, R.A.; Neves, K.B.; Pestana, C.R.; Queiroz, A.L.; Zanotto, C.Z.; Chignalia, A.Z.; Valim, Y.M.; Silveira, L.R.; Curti, C.; Tostes, R.C. Testosterone induces apoptosis in vascular smooth muscle cells via extrinsic apoptotic pathway with mitochondria-generated reactive oxygen species involvement. Am. J. Physiol. Heart Circ. Physiol. 2014 , 306 , H1485–H1494. [ Google Scholar ] [ CrossRef ] [ Green Version ]
  • Hackett, D.A.; Johnson, N.A.; Chow, C.M. Training practices and ergogenic aids used by male bodybuilders. J. Strength Cond. Res. 2013 , 27 , 1609–1617. [ Google Scholar ] [ CrossRef ]
  • Sjöqvist, F.; Garle, M.; Rane, A. Use of doping agents, particularly anabolic steroids, in sports and society. Lancet 2008 , 371 , 1872–1882. [ Google Scholar ] [ CrossRef ]
  • Turillazzi, E.; Perilli, G.; Di Paolo, M.; Neri, M.; Riezzo, I.; Fineschi, V. Side effects of AAS abuse: An overview. Mini Rev. Med. Chem. 2011 , 11 , 374–389. [ Google Scholar ] [ CrossRef ] [ PubMed ]
  • Van Amsterdam, J.; Opperhuizen, A.; Hartgens, F. Adverse health effects of anabolic–androgenic steroids. Regul. Toxicol. Pharmacol. 2010 , 57 , 117–123. [ Google Scholar ] [ CrossRef ] [ PubMed ]
  • Petersson, A.; Garle, M.; Holmgren, P.; Druid, H.; Krantz, P.; Thiblin, I. Toxicological findings and manner of death in autopsied users of anabolic androgenic steroids. Drug Alcohol Depend. 2006 , 81 , 241–249. [ Google Scholar ] [ CrossRef ] [ PubMed ]
  • Frati, P.; Busardo, F.P.; Cipolloni, L.; De Dominicis, E.; Fineschi, V. Anabolic androgenic steroid(AAS) related deaths: Autoptic, histopathological and toxicological findings. Curr. Neuropharmacol. 2015 , 13 , 146–159. [ Google Scholar ] [ CrossRef ] [ Green Version ]
  • Darke, S.; Torok, M.; Duflou, J. Sudden or unnatural deaths involving anabolic-androgenic steroids. J. Forensic Sci. 2014 , 59 , 1025–1028. [ Google Scholar ] [ CrossRef ]
  • Di Paolo, M.; Agozzino, M.; Toni, C.; Luciani, A.B.; Molendini, L.; Scaglione, M.; Inzani, F.; Pasotti, M.; Buffi, F.; Arbustini, E. Sudden anabolic steroid abuse-related death in athletes. Int. J. Cardiol. 2007 , 114 , 114–117. [ Google Scholar ] [ CrossRef ] [ PubMed ]
  • Hernández-Guerra, A.I.; Tapia, J.; Menéndez-Quintanal, L.M.; Lucena, J.S. Sudden cardiac death in anabolic androgenic steroids abuse: Case report and literature review. Forensic Sci. Res. 2019 , 4 , 267–273. [ Google Scholar ] [ CrossRef ]
  • Caraci, F.; Pistarà, V.; Corsaro, A.; Tomasello, F.; Giuffrida, M.L.; Sortino, M.A.; Nicoletti, F.; Copani, A. Neurotoxic properties of the anabolic androgenic steroids nandrolone and methandrostenolone in primary neuronal cultures. J. Neurosci. Res. 2011 , 89 , 592–600. [ Google Scholar ] [ CrossRef ]
  • Karimooy, F.N.; Bideskan, A.E.; Pour, A.M.; Hoseini, S.M. Neurotoxic Effects of Stanozolol on Male Rats’ Hippocampi: Does Stanozolol cause apoptosis? Biomol. Concepts 2019 , 10 , 73–81. [ Google Scholar ] [ CrossRef ]
  • Gomes, F.G.; Fernandes, J.; Campos, D.V.; Cassilhas, R.C.; Viana, G.M.; D’Almeida, V.; de Moraes Rego, M.K.; Buainain, P.I.; Cavalheiro, E.A.; Arida, R.M. The beneficial effects of strength exercise on hippocampal cell proliferation and apoptotic signaling is impaired by anabolic androgenic steroids. Psychoneuroendocrinology 2014 , 50 , 106–117. [ Google Scholar ] [ CrossRef ] [ PubMed ]
  • Sessa, F.; Salerno, M.; Cipolloni, L.; Bertozzi, G.; Messina, G.; Di Mizio, G.; Asmundo, A.; Pomara, C. Anabolic-androgenic steroids and brain injury: miRNA evaluation in users compared to cocaine abusers and elderly people. Aging 2020 , 12 . [ Google Scholar ] [ CrossRef ]
  • Bjørnebekk, A.; Westlye, L.T.; Walhovd, K.B.; Jørstad, M.L.; Sundseth, Ø.Ø.; Fjell, A.M. Cognitive performance and structural brain correlates in long-term anabolic-androgenic steroid exposed and nonexposed weightlifters. Neuropsychology 2019 , 33 , 547. [ Google Scholar ] [ CrossRef ] [ PubMed ]
  • Bjørnebekk, A.; Walhovd, K.B.; Jørstad, M.L.; Due-Tønnessen, P.; Hullstein, I.R.; Fjell, A.M. Structural brain imaging of long-term anabolic-androgenic steroid users and nonusing weightlifters. Biol. Psychiatry 2017 , 82 , 294–302. [ Google Scholar ] [ CrossRef ] [ Green Version ]
  • Kaufman, M.J.; Janes, A.C.; Hudson, J.I.; Brennan, B.P.; Kanayama, G.; Kerrigan, A.R.; Jensen, J.E.; Pope, H.G., Jr. Brain and cognition abnormalities in long-term anabolic-androgenic steroid users. Drug Alcohol Depend. 2015 , 152 , 47–56. [ Google Scholar ] [ CrossRef ] [ Green Version ]
  • Kanayama, G.; Kean, J.; Hudson, J.I.; Pope, H.G., Jr. Cognitive deficits in long-term anabolic-androgenic steroid users. Drug Alcohol Depend. 2013 , 130 , 208–214. [ Google Scholar ] [ CrossRef ] [ Green Version ]
  • Kaufman, M.J.; Kanayama, G.; Hudson, J.I.; Pope, H.G., Jr. Supraphysiologic-dose anabolic–androgenic steroid use: A risk factor for dementia? Neurosci. Biobehav. Rev. 2019 , 100 , 180–207. [ Google Scholar ] [ CrossRef ]
  • Piacentino, D.; Kotzalidis, G.D.; Del Casale, A.; Aromatario, M.R.; Pomara, C.; Girardi, P.; Sani, G. Anabolic-androgenic steroid use and psychopathology in athletes. A systematic review. Curr. Neuropharmacol. 2015 , 13 , 101–121. [ Google Scholar ] [ CrossRef ] [ Green Version ]
  • Rashid, H.; Ormerod, S.; Day, E. Anabolic androgenic steroids: What the psychiatrist needs to know. Adv. Psychiatr. Treat. 2007 , 13 , 203–211. [ Google Scholar ] [ CrossRef ] [ Green Version ]
  • Joksimovic, J.; Selakovic, D.; Jovicic, N.; Mitrovic, S.; Mihailovic, V.; Katanic, J.; Milovanovic, D.; Rosic, G. Exercise Attenuates Anabolic Steroids-Induced Anxiety via Hippocampal NPY and MC4 Receptor in Rats. Front. Neurosci. 2019 , 13 , 172. [ Google Scholar ] [ CrossRef ]
  • Hauger, L.E.; Sagoe, D.; Vaskinn, A.; Arnevik, E.A.; Leknes, S.; Jørstad, M.L.; Bjørnebekk, A. Anabolic androgenic steroid dependence is associated with impaired emotion recognition. Psychopharmacology 2019 , 236 , 2667–2676. [ Google Scholar ] [ CrossRef ] [ PubMed ]
  • Almeida, O.P.; Yeap, B.B.; Hankey, G.J.; Jamrozik, K.; Flicker, L. Low Free Testosterone Concentration as a Potentially Treatable Cause of Depressive Symptoms in Older Men ; Arch. Gen. Psychiatry: Chicago, IL, USA, 2008; Volume 12, pp. 283–289. ISSN 0003-990X. [ Google Scholar ]
  • Bueno, A.; Carvalho, F.B.; Gutierres, J.M.; Lhamas, C.; Andrade, C.M. A comparative study of the effect of the dose and exposure duration of anabolic androgenic steroids on behavior, cholinergic regulation, and oxidative stress in rats. PLoS ONE 2017 , 12 , e0177623. [ Google Scholar ] [ CrossRef ] [ PubMed ] [ Green Version ]
  • Henderson, L.P.; Penatti, C.A.; Jones, B.L.; Yang, P.; Clark, A.S. Anabolic androgenic steroids and forebrain GABAergic transmission. Neuroscience 2006 , 138 , 793–799. [ Google Scholar ] [ CrossRef ] [ PubMed ]
  • Kouvelas, D.; Pourzitaki, C.; Papazisis, G.; Dagklis, T.; Dimou, K.; Kraus, M.M. Nandrolone abuse decreases anxiety and impairs memory in rats via central androgenic receptors. Int. J. Neuropsychopharmacol. 2008 , 11 , 925–934. [ Google Scholar ] [ CrossRef ] [ PubMed ] [ Green Version ]
  • Melloni, R.H., Jr.; Ricci, L.A. Adolescent exposure to anabolic/androgenic steroids and the neurobiology of offensive aggression: A hypothalamic neural model based on findings in pubertal Syrian hamsters. Horm. Behav. 2010 , 58 , 177–191. [ Google Scholar ] [ CrossRef ] [ PubMed ]
  • Quaglio, G.; Fornasiero, A.; Mezzelani, P.; Moreschini, S.; Lugoboni, F.; Lechi, A. Anabolic steroids: Dependence and complications of chronic use. Intern. Emerg. Med. 2009 , 4 , 289–296. [ Google Scholar ] [ CrossRef ]
  • Elfverson, M.; Johansson, T.; Zhou, Q.; Le Grevès, P.; Nyberg, F. Chronic administration of the anabolic androgenic steroid nandrolone alters neurosteroid action at the sigma-1 receptor but not at the sigma-2 or NMDA receptors. Neuropharmacology 2011 , 61 , 1172–1181. [ Google Scholar ] [ CrossRef ]
  • Sato, S.M.; Schulz, K.M.; Sisk, C.L.; Wood, R.I. Adolescents and androgens, receptors and rewards. Horm. Behav. 2008 , 53 , 647–658. [ Google Scholar ] [ CrossRef ] [ Green Version ]
  • Bertozzi, G.; Sessa, F.; Albano, G.D.; Sani, G.; Maglietta, F.; Roshan, M.H.; Li Volti, G.; Bernardini, R.; Avola, R.; Pomara, C.; et al. The role of anabolic androgenic steroids in disruption of the physiological function in discrete areas of the central nervous system. Mol. Neurobiol. 2018 , 55 , 5548–5556. [ Google Scholar ] [ CrossRef ]
  • Bertozzi, G.; Salerno, M.; Pomara, C.; Sessa, F. Neuropsychiatric and Behavioral Involvement in AAS Abusers. A Literature Review. Medicina 2019 , 55 , 396. [ Google Scholar ] [ CrossRef ] [ Green Version ]
  • Liu, J.D.; Wu, Y.Q. Anabolic-androgenic steroids and cardiovascular risk. Chin. Med. J. 2019 , 132 , 2229. [ Google Scholar ] [ CrossRef ] [ PubMed ]
  • Vasilaki, F.; Tsitsimpikou, C.; Tsarouhas, K.; Germanakis, I.; Tzardi, M.; Kavvalakis, M.; Ozcagli, E.; Kouretas, D.; Tsatsakis, A.M. Cardiotoxicity in rabbits after long-term nandrolone decanoate administration. Toxicol. Lett. 2016 , 241 , 143–151. [ Google Scholar ] [ CrossRef ] [ PubMed ]
  • Pereira-Junior, P.P.; Chaves, E.A.; Costa-e-Sousa, R.H.; Masuda, M.O.; de Carvalho, A.C.; Nascimento, J.H. Cardiac autonomic dysfunction in rats chronically treated with anabolic steroid. Eur. J. Appl. Physiol. 2006 , 96 , 487–494. [ Google Scholar ] [ CrossRef ] [ PubMed ]
  • Riezzo, I.; Di Paolo, M.; Neri, M.; Bello, S.; Cantatore, S.; D’Errico, S.; Dinucci, D.; Parente, R.; Pomara, C.; Rabozzi, R.; et al. Anabolic steroid-and exercise-induced cardio-depressant cytokines and myocardial β1 receptor expression in CD1 mice. Curr. Pharm. Biotechnol. 2011 , 12 , 275–284. [ Google Scholar ] [ CrossRef ]
  • Christou, G.A.; Christou, K.A.; Nikas, D.N.; Goudevenos, J.A. Acute myocardial infarction in a young bodybuilder taking anabolic androgenic steroids: A case report and critical review of the literature. Eur. J. Prev. Cardiol. 2016 , 23 , 1785–1796. [ Google Scholar ] [ CrossRef ]
  • Achar, S.; Rostamian, A.; Narayan, S.M. Cardiac and metabolic effects of anabolic-androgenic steroid abuse on lipids, blood pressure, left ventricular dimensions, and rhythm. Am. J. Cardiol. 2010 , 106 , 893–901. [ Google Scholar ] [ CrossRef ] [ Green Version ]
  • Golestani, R.; Slart, R.H.; Dullaart, R.P.; Glaudemans, A.W.; Zeebregts, C.J.; Boersma, H.H.; Tio, R.A.; Dierckx, R.A. Adverse cardiovascular effects of anabolic steroids: Pathophysiology imaging. Eur. J. Clin. Investig. 2012 , 42 , 795. [ Google Scholar ] [ CrossRef ] [ Green Version ]
  • Alves, M.J.; Dos Santos, M.R.; Dias, R.G.; Akiho, C.A.; Laterza, M.C.; Rondon, M.U.; Moreau, R.L.; Negrao, C.E. Abnormal neurovascular control in anabolic androgenic steroids users. Med. Sci. Sports Exerc. 2010 , 42 , 865–871. [ Google Scholar ] [ CrossRef ]
  • Baggish, A.L.; Weiner, R.B.; Kanayama, G.; Hudson, J.I.; Picard, M.H.; Hutter, A.M., Jr.; Pope, H.G., Jr. Long-term anabolic-androgenic steroid use is associated with left ventricular dysfunction. Circ. Heart Fail. 2010 , 3 , 472–476. [ Google Scholar ] [ CrossRef ] [ Green Version ]
  • Seara, F.A.; Barbosa, R.A.; de Oliveira, D.F.; Gran da Silva, D.L.; Carvalho, A.B.; Freitas Ferreira, A.C.; Nascimento, J.H.; Olivares, E.L. Administration of anabolic steroid during adolescence induces long-term cardiac hypertrophy and increases susceptibility to ischemia/reperfusion injury in adult Wistar rats. J. Steroid Biochem. Mol. Biol. 2017 , 171 , 34–42. [ Google Scholar ] [ CrossRef ]
  • Rasmussen, J.J.; Schou, M.; Madsen, P.L.; Selmer, C.; Johansen, M.L.; Ulriksen, P.S.; Dreyer, T.; Kumler, T.; Plesner, L.L.; Faber, J.; et al. Cardiac systolic dysfunction in past illicit users of anabolic androgenic steroids. Am. Heart J. 2018 , 203 , 49–56. [ Google Scholar ] [ CrossRef ]
  • Nottin, S.; Nguyen, L.D.; Terbah, M.; Obert, P. Cardiovascular effects of androgenic anabolic steroids in male bodybuilders determined by tissue Doppler imaging. Am. J. Cardiol. 2006 , 97 , 912–915. [ Google Scholar ] [ CrossRef ]
  • Rocha, F.L.; Carmo, E.C.; Roque, F.R.; Hashimoto, N.Y.; Rossoni, L.V.; Frimm, C.; Aneas, I.; Negrao, C.E.; Krieger, J.E.; Oliveira, E.M. Anabolic steroids induce cardiac renin-angiotensin system and impair the beneficial effects of aerobic training in rats. Am. J. Physiol. Heart Circ. Physiol. 2007 , 293 , H3575–H3583. [ Google Scholar ] [ CrossRef ] [ Green Version ]
  • Akçakoyun, M.; Alizade, E.; Gündoğdu, R.; Bulut, M.; Tabakci, M.M.; Açar, G.; Avci, A.; Simsek, Z.; Fidan, S.; Demir, S.; et al. Long-term anabolic androgenic steroid use is associated with increased atrial electromechanical delay in male bodybuilders. Biomed. Res. Int. 2014 . [ Google Scholar ] [ CrossRef ]
  • Angell, P.; Chester, N.; Green, D.; Somauroo, J.; Whyte, G.; George, K. Anabolic steroids and cardiovascular risk. Sports Med. 2012 , 42 , 119–134. [ Google Scholar ] [ CrossRef ]
  • Marocolo, M.; Silva-Neto, J.A.; Neto, O.B. Acute interruption of treatment with nandrolone decanoate is not sufficient to reverse cardiac autonomic dysfunction and ventricular repolarization disturbances in rats. Steroids 2018 , 132 , 12–17. [ Google Scholar ] [ CrossRef ]
  • Phillis, B.D.; Abeywardena, M.Y.; Adams, M.J.; Kennedy, J.A.; Irvine, R.J. Nandrolone potentiates arrhythmogenic effects of cardiac ischemia in the rat. Toxicol. Sci. 2007 , 99 , 605–611. [ Google Scholar ] [ CrossRef ]
  • Olivares, E.L.; Silveira, A.L.; Fonseca, F.V.; Silva-Almeida, C.; Côrtes, R.S.; Pereira-Junior, P.P.; Nascimento, J.H.; Reis, L.C. Administration of an anabolic steroid during the adolescent phase changes the behavior, cardiac autonomic balance and fluid intake in male adult rats. Physiol. Behav. 2014 , 126 , 15–24. [ Google Scholar ] [ CrossRef ] [ Green Version ]
  • Esperón, C.G.; Lopez-Cancio, E.; Bermejo, P.G.; Dávalos, A. Anabolic Androgenic Steroids and Stroke. In Neuropathology of Drug Addictions and Substance Misuse , 2nd ed.; Preedy, V.R., Ed.; Academic Press: London, UK, 2016; pp. 981–990. ISBN 978-0-12-800212-4. [ Google Scholar ]
  • Baggish, A.L.; Weiner, R.B.; Kanayama, G.; Hudson, J.I.; Lu, M.T.; Hoffmann, U.; Pope, H.G., Jr. Cardiovascular toxicity of illicit anabolic-androgenic steroid use. Circulation 2017 , 135 , 1991–2002. [ Google Scholar ] [ CrossRef ] [ Green Version ]
  • Bond, P.; Llewellyn, W.; Van Mol, P. Anabolic androgenic steroid-induced hepatotoxicity. Med. Hypotheses 2016 , 93 , 150–153. [ Google Scholar ] [ CrossRef ]
  • Neri, M.; Bello, S.; Bonsignore, A.; Cantatore, S.; Riezzo, I.; Turillazzi, E.; Fineschi, V. Anabolic androgenic steroids abuse and liver toxicity. Mini Rev. Med. Chem. 2011 , 11 , 430–437. [ Google Scholar ] [ CrossRef ] [ PubMed ]
  • Solimini, R.; Rotolo, M.C.; Mastrobattista, L.; Mortali, C.; Minutillo, A.; Pichini, S.; Pacifici, I.; Palmi, I. Hepatotoxicity associated with illicit use of anabolic androgenic steroids in doping. Eur. Rev. Med. Pharmacol. Sci. 2017 , 21 , 7–16. [ Google Scholar ] [ PubMed ]
  • Vieira, R.P.; Franca, R.F.; Damaceno-Rodrigues, N.R.; Dolhnikoff, M.; Caldini, E.G.; Carvalho, C.R.; Ribeiro, W. Dose-dependent hepatic response to subchronic administration of nandrolone decanoate. Med. Sci. Sports Exerc. 2008 , 40 , 842. [ Google Scholar ] [ CrossRef ] [ PubMed ]
  • Schwingel, P.A.; Cotrim, H.P.; Salles, B.R.; Almeida, C.E.; dos Santos, C.R., Jr.; Nachef, B.; Andrade, A.R.; Zoppi, C.C. Anabolic-androgenic steroids: A possible new risk factor of toxicant-associated fatty liver disease. Liver Int. 2011 , 31 , 348–353. [ Google Scholar ] [ CrossRef ] [ PubMed ]
  • Riezzo, I.; Turillazzi, E.; Bello, S.; Cantatore, S.; Cerretani, D.; Di Paolo, M.; Fiaschi, A.I.; Frati, P.; Neri, M.; Pedretti, M.; et al. Chronic nandrolone administration promotes oxidative stress, induction of pro-inflammatory cytokine and TNF-α mediated apoptosis in the kidneys of CD1 treated mice. Toxicol. Appl. Pharmacol. 2014 , 280 , 97–106. [ Google Scholar ] [ CrossRef ]
  • Kahal, A.; Allem, R. Reversible effects of anabolic steroid abuse on cyto-architectures of the heart, kidneys and testis in adult male mice. Biomed. Pharmacother. 2018 , 106 , 917–922. [ Google Scholar ] [ CrossRef ]
  • Brasil, G.A.; de Lima, E.M.; do Nascimento, A.M.; Caliman, I.F.; de Medeiros, A.R.; Silva, M.S.; de Abreu, G.R.; dos Reis, A.M.; de Andrade, T.U.; Bissoli, N.S. Nandrolone decanoate induces cardiac and renal remodeling in female rats, without modification in physiological parameters: The role of ANP system. Life Sci. 2015 , 137 , 65–73. [ Google Scholar ] [ CrossRef ]
  • Josiak, K.; Jankowska, E.A.; Piepoli, M.F.; Banasiak, W.; Ponikowski, P. Skeletal myopathy in patients with chronic heart failure: Significance of anabolic-androgenic hormones. J. Cachexia Sarcopenia Muscle 2014 , 5 , 287–296. [ Google Scholar ] [ CrossRef ] [ Green Version ]
  • Hoseini, L.; Roozbeh, J.; Sagheb, M.; Karbalay-Doust, S.; Noorafshan, A. Nandrolone decanoate increases the volume but not the length of the proximal and distal convoluted tubules of the mouse kidney. Micron 2009 , 40 , 226–230. [ Google Scholar ] [ CrossRef ]
  • Carson, J.A.; Manolagas, S.C. Effects of sex steroids on bones and muscles: Similarities, parallels, and putative interactions in health and disease. Bone 2015 , 80 , 67–78. [ Google Scholar ] [ CrossRef ] [ Green Version ]
  • Casavant, M.J.; Blake, K.; Griffith, J.; Yates, A.; Copley, L.M. Consequences of use of anabolic androgenic steroids. Pediatr. Clin. N. Am. 2007 , 54 , 677–690. [ Google Scholar ] [ CrossRef ]
  • Andrews, M.A.; Magee, C.D.; Combest, T.M.; Allard, R.J.; Douglas, K.M. Physical effects of anabolic-androgenic steroids in healthy exercising adults: A systematic review and meta-analysis. Curr. Sports Med. Rep. 2018 , 17 , 232–241. [ Google Scholar ] [ CrossRef ]
  • Caminiti, G.; Volterrani, M.; Iellamo, F.; Marazzi, G.; Massaro, R.; Miceli, M.; Mammi, C.; Piepoli, M.; Fini, M.; Rosano, G.M. Effect of long-acting testosterone treatment on functional exercise capacity, skeletal muscle performance, insulin resistance, and baroreflex sensitivity in elderly patients with chronic heart failure: A double-blind, placebo-controlled, randomized study. J. Am. Coll. Cardiol. 2009 , 54 , 919–927. [ Google Scholar ] [ CrossRef ]
  • Marqueti, R.C.; Prestes, J.; Stotzer, U.S.; Paschoal, M.; Leite, R.D.; Perez, S.E.; de Araujo, H.S. MMP-2, jumping exercise and nandrolone in skeletal muscle. Int. J. Sports Med. 2008 , 29 , 559–563. [ Google Scholar ] [ CrossRef ]
  • Paschoal, M.; de Cássia Marqueti, R.; Perez, S.; Selistre-de-Araujo, H.S. Nandrolone inhibits VEGF mRNA in rat muscle. Int. J. Sports Med. 2009 , 30 , 775–778. [ Google Scholar ] [ CrossRef ]
  • Marqueti, R.C.; Prestes, J.; Wang, C.C.; Ramos, O.H.; Perez, S.E.; Nakagaki, W.R.; Carvalho, H.F.; Selistre-de-Araujo, H.S. Biomechanical responses of different rat tendons to nandrolone decanoate and load exercise. Scand. J. Med. Sci. Sports 2011 , 21 , e91–e99. [ Google Scholar ] [ CrossRef ]
  • Marqueti, R.C.; Prestes, J.; Paschoal, M.; Ramos, O.H.; Perez, S.E.; Carvalho, H.F.; Selistre-de-Araujo, H.S. Matrix metallopeptidase 2 activity in tendon regions: Effects of mechanical loading exercise associated to anabolic-androgenic steroids. Eur. J. Appl. Physiol. 2008 , 104 , 1087. [ Google Scholar ] [ CrossRef ]
  • Pomara, C.; Barone, R.; Marino Gammazza, A.; Sangiorgi, C.; Barone, F.; Pitruzzella, A.; Locorotondo, N.; Di Gaudio, F.; Salerno, M.; Maglietta, F.; et al. Effects of nandrolone stimulation on testosterone biosynthesis in leydig cells. J. Cell. Physiol. 2016 , 231 , 1385–1391. [ Google Scholar ] [ CrossRef ] [ Green Version ]
  • Armstrong, J.M.; Avant, R.A.; Charchenko, C.M.; Westerman, M.E.; Ziegelmann, M.J.; Miest, T.S.; Trost, L.W. Impact of anabolic androgenic steroids on sexual function. Transl. Androl. Urol. 2018 , 7 , 483. [ Google Scholar ] [ CrossRef ]
  • Coward, R.M.; Rajanahally, S.; Kovac, J.R.; Smith, R.P.; Pastuszak, A.W.; Lipshultz, L.I. Anabolic steroid induced hypogonadism in young men. J. Urol. 2013 , 190 , 2200–2205. [ Google Scholar ] [ CrossRef ]
  • Barone, R.; Pitruzzella, A.; Marino Gammazza, A.; Rappa, F.; Salerno, M.; Barone, F.; Sangiorgi, C.; D’Amico, D.; Locorotondo, N.; Di Gaudio, F.; et al. Nandrolone decanoate interferes with testosterone biosynthesis altering blood–testis barrier components. J. Cell. Mol. Med. 2017 , 21 , 1636–1647. [ Google Scholar ] [ CrossRef ] [ PubMed ]
  • García-Manso, J.M.; Esteve, T.V. Consequences of the Use of Anabolic-Androgenic Steroids for Male Athletes’ Fertility. In Exercise and Human Reproduction ; Vaamonde, D., du Plessis, S.S., Agarwal, A., Eds.; Springer: New York, NY, USA, 2016; pp. 153–165. ISBN 978-1-4939-3402-7. [ Google Scholar ]
  • Rahnema, C.D.; Lipshultz, L.I.; Crosnoe, L.E.; Kovac, J.R.; Kim, E.D. Anabolic steroid–induced hypogonadism: Diagnosis and treatment. Fertil. Steril. 2014 , 101 , 1271–1279. [ Google Scholar ] [ CrossRef ] [ PubMed ]
  • Tatem, A.J.; Beilan, J.; Kovac, J.R.; Lipshultz, L.I. Management of anabolic steroid-induced infertility: Novel strategies for fertility maintenance and recovery. World J. Mens. Health 2020 , 38 , 141–150. [ Google Scholar ] [ CrossRef ] [ PubMed ]
  • Agledahl, I.; Brodin, E.; Svartberg, J.; Hansen, J.B. Plasma free tissue factor pathway inhibitor(TFPI) levels and TF-induced thrombin generation ex vivo in men with low testosterone levels. Thromb. Haemost. 2009 , 101 , 471–477. [ Google Scholar ] [ CrossRef ] [ Green Version ]
  • Liljeqvist, S.; Helldén, A.; Bergman, U.; Söderberg, M. Pulmonary embolism associated with the use of anabolic steroids. Eur. J. Intern. Med. 2008 , 19 , 214–215. [ Google Scholar ] [ CrossRef ]
  • Thiblin, I.; Garmo, H.; Garle, M.; Holmberg, L.; Byberg, L.; Michaëlsson, K.; Gedeborg, R. Anabolic steroids and cardiovascular risk: A national population-based cohort study. Drug Alcohol Depend. 2015 , 152 , 87–92. [ Google Scholar ] [ CrossRef ]
  • Jin, H.; Lin, J.; Fu, L.; Mei, Y.F.; Peng, G.; Tan, X.; Wang, D.M.; Wang, W.; Li, Y.G. Physiological testosterone stimulates tissue plasminogen activator and tissue factor pathway inhibitor and inhibits plasminogen activator inhibitor type 1 release in endothelial cells. Biochem. Cell Biol. 2007 , 85 , 246–251. [ Google Scholar ] [ CrossRef ]
  • Chang, S.; Münster, A.B.; Gram, J.; Sidelmann, J.J. Anabolic androgenic steroid abuse: The effects on thrombosis risk, coagulation, and fibrinolysis. Semin. Thromb. Hemost. 2018 , 44 , 734–746. [ Google Scholar ] [ CrossRef ]
  • Bertozzi, G.; Sessa, F.; Maglietta, F.; Cipolloni, L.; Salerno, M.; Fiore, C.; Fortarezza, P.; Ricci, P.; Turillazzi, E.; Pomara, C. Immunodeficiency as a side effect of anabolic androgenic steroid abuse: A case of necrotizing myofasciitis. Forensic Sci. Med. Pathol. 2019 , 15 , 616–621. [ Google Scholar ] [ CrossRef ]
  • Salerno, M.; Cascio, O.; Bertozzi, G.; Sessa, F.; Messina, A.; Monda, V.; Cipolloni, L.; Biondi, A.; Daniele, A.; Pomara, C. Anabolic androgenic steroids and carcinogenicity focusing on Leydig cell: A literature review. Oncotarget 2018 , 9 , 19415. [ Google Scholar ] [ CrossRef ] [ Green Version ]
  • Hashimoto, H.; Vertino, P.M.; Cheng, X. Molecular coupling of DNA methylation and histone methylation. Epigenomics 2010 , 2 , 657–669. [ Google Scholar ] [ CrossRef ] [ PubMed ] [ Green Version ]
  • Medei, E.; Marocolo, M.; de Carvalho Rodrigues, D.; Arantes, P.C.; Takiya, C.M.; Silva, J.; Rondinelli, E.; Goldenberg, R.C.; Campos de Carvalho, A.C.; Nascimento, J.H. Chronic treatment with anabolic steroids induces ventricular repolarization disturbances: Cellular, ionic and molecular mechanism. J. Mol. Cell. Cardiol. 2010 , 49 , 165–175. [ Google Scholar ] [ CrossRef ] [ PubMed ]
  • Schwarzenbach, H. Impact of physical activity and doping on epigenetic gene regulation. Drug Test. Anal. 2011 , 3 , 682–687. [ Google Scholar ] [ CrossRef ] [ PubMed ]
  • Souza, L.D.; da Cruz, L.A.; Cerqueira, E.D.; Meireles, J.R. Micronucleus as biomarkers of cancer risk in anabolic androgenic steroids users. Hum. Exp. Toxicol. 2017 , 36 , 302–310. [ Google Scholar ] [ CrossRef ] [ PubMed ]
  • Agriesti, F.; Tataranni, T.; Pacelli, C.; Scrima, R.; Laurenzana, I.; Ruggieri, V.; Cela, O.; Mazzoccoli, C.; Salerno, M.; Sessa, F.; et al. Nandrolone induces a stem cell-like phenotype in human hepatocarcinoma-derived cell line inhibiting mitochondrial respiratory activity. Sci. Rep. 2020 , 10 , 1–17. [ Google Scholar ] [ CrossRef ] [ Green Version ]
  • Tentori, L.; Graziani, G. Doping with growth hormone/IGF-1, anabolic steroids or erythropoietin: Is there a cancer risk? Pharmacol. Res. 2007 , 55 , 359–369. [ Google Scholar ] [ CrossRef ]
  • Kanayama, G.; Pope, H.G., Jr.; Hudson, J.I. Associations of anabolic-androgenic steroid use with other behavioral disorders: An analysis using directed acyclic graphs. Psychol. Med. 2018 , 48 , 2601–2608. [ Google Scholar ] [ CrossRef ]
  • Brooks, J.H.; Ahmad, I.; Easton, G. Anabolic steroid use. Br. Med. J. 2016 , 355 . [ Google Scholar ] [ CrossRef ]
  • Grönbladh, A.; Nylander, E.; Hallberg, M. The neurobiology and addiction potential of anabolic androgenic steroids and the effects of growth hormone. Brain Res. Bull. 2016 , 126 , 127–137. [ Google Scholar ] [ CrossRef ]
  • Riezzo, I.; De Carlo, D.; Neri, M.; Nieddu, A.; Turillazzi, E.; Fineschi, V. Heart disease induced by AAS abuse, using experimental mice/rats models and the role of exercise-induced cardiotoxicity. Mini Rev. Med. Chem. 2011 , 11 , 409–424. [ Google Scholar ] [ CrossRef ]
  • Pagonis, T.A.; Angelopoulos, N.V.; Koukoulis, G.N.; Hadjichristodoulou, C.S. Psychiatric side effects induced by supraphysiological doses of combinations of anabolic steroids correlate to the severity of abuse. Eur. Psychiatry 2006 , 21 , 551–562. [ Google Scholar ] [ CrossRef ]
  • Baker, J.S.; Graham, M.R.; Davies, B. Steroid and prescription medicine abuse in the health and fitness community: A regional study. Eur. J. Intern. Med. 2006 , 17 , 479–484. [ Google Scholar ] [ CrossRef ]
  • Nagata, J.M.; Ganson, K.T.; Gorrell, S.; Mitchison, D.; Murray, S.B. Association between legal performance-enhancing substances and use of anabolic-androgenic steroids in young adults. JAMA Pediatr. 2020 . [ Google Scholar ] [ CrossRef ]
  • Bates, G.; Van Hout, M.C.; Teck, J.T.W.; McVeigh, J. Treatments for people who use anabolic androgenic steroids: A scoping review. Harm Reduct. J. 2019 , 16 , 75. [ Google Scholar ] [ CrossRef ] [ PubMed ] [ Green Version ]
  • Sessa, F.; Salerno, M.; Bertozzi, G.; Cipolloni, L.; Messina, G.; Aromatario, M.; Polo, L.; Turillazzi, E.; Pomara, C. miRNAs as novel biomarkers of chronic kidney injury in anabolic-androgenic steroid users: An experimental study. Front. Pharmacol. 2020 , 11 , 1454. [ Google Scholar ] [ CrossRef ] [ PubMed ]
  • Torrisi, M.; Pennisi, G.; Russo, I.; Amico, F.; Esposito, M.; Liberto, A.; Cocimano, G.; Salerno, M.; Li Rosi, G.; Di Nunno, N.; et al. Sudden Cardiac Death in Anabolic-Androgenic Steroid Users: A Literature Review. Medicina 2020 , 56 , 587. [ Google Scholar ] [ CrossRef ] [ PubMed ]
  • Mottram, D.R.; Alan, J.G. Anabolic steroids. Best Pract. Res. Clin. Endocrinol. Metab. 2000 , 14 , 55–69. [ Google Scholar ] [ CrossRef ]
  • Albano, G.D.; Sessa, F.; Messina, A.; Monda, V.; Bertozzi, G.; Maglietta, F.; Giugliano, P.; Vacchiano, G.; Marsala, G.; Salerno, M. AAS and organs damage: A focus on Nandrolone effects. Acta Med. Mediterr. 2017 , 6 , 939–946. [ Google Scholar ] [ CrossRef ]
  • Sessa, F.; Maglietta, F.; Bertozzi, G.; Salerno, M.; Di Mizio, G.; Messina, G.; Montana, A.; Ricci, P.; Pomara, C. Human brain injury and mirnas: An experimental study. Int. J. Mol. Sci. 2019 , 20 , 1546. [ Google Scholar ] [ CrossRef ] [ Green Version ]
  • Monda, V.; Salerno, M.; Sessa, F.; Bernardini, R.; Valenzano, A.; Marsala, G.; Zammit, C.; Avola, R.; Carotenuto, M.; Messina, G.; et al. Functional changes of orexinergic reaction to psychoactive substances. Mol. Neurobiol. 2018 , 55 , 6362–6368. [ Google Scholar ] [ CrossRef ]
  • Sessa, F.; Franco, S.; Picciocchi, E.; Geraci, D.; Chisari, M.G.; Marsala, G.; Polito, A.N.; Sorrentino, M.; Tripi, G.; Salerno, M.; et al. Addictions substance free during lifespan. Acta Med. Mediter. 2018 , 34 , 2081–2087. [ Google Scholar ]
  • Monda, V.; Salerno, M.; Fiorenzo, M.; Villano, I.; Viggiano, A.; Sessa, F.; Triggiani, A.I.; Cibelli, G.; Valenzano, A.; Marsala, G.; et al. Role of sex hormones in the control of vegetative and metabolic functions of middle-aged women. Front. Physiol. 2017 , 8 , 773. [ Google Scholar ] [ CrossRef ] [ PubMed ] [ Green Version ]
  • Pomara, C.; Neri, M.; Bello, S.; Fiore, C.; Riezzo, I.; Turillazzi, E. Neurotoxicity by synthetic androgen steroids: Oxidative stress, apoptosis, and neuropathology: A review. Curr. Neuropharmacol. 2015 , 13 , 132–145. [ Google Scholar ] [ CrossRef ] [ PubMed ] [ Green Version ]
  • Pomara, C.; D’Errico, S.; Riezzo, I.; De Cillis, G.P.; Fineschi, V. Sudden cardiac death in a child affected by Prader-Willi syndrome. Int. J. Legal Med. 2005 , 119 , 153–157. [ Google Scholar ] [ CrossRef ] [ PubMed ]
  • Fineschi, V.; Neri, M.; Di Donato, S.; Pomara, C.; Riezzo, I.; Turillazzi, E. An immunohistochemical study in a fatality due to ovarian hyperstimulation syndrome. Int. J. Legal Med. 2006 , 120 , 293–299. [ Google Scholar ] [ CrossRef ] [ PubMed ]
  • Neri, M.; Riezzo, I.; Pomara, C.; Schiavone, S.; Turillazzi, E. Oxidative-nitrosative stress and myocardial dysfunctions in sepsis: Evidence from the literature and postmortem observations. Mediat. Inflamm. 2016 , 2016 , 3423450. [ Google Scholar ] [ CrossRef ] [ Green Version ]

Click here to enlarge figure

Organ/SystemAuthor(s)Year of PublicationAdverse Effects
Brain and BehaviorBertozzi, et al.2019↓ orbitofrontal cortex; lack of inhibitory control.
Hauger, et al.2019↓ memory, ↓ anxiety ↑ depression.
Joksimovic, et al.2019↓ orbitofrontal cortex; lack of inhibitory control.
Karimooy, et al.2019Neurodegeneration; histopathologic changes in hippocampus.
Bertozzi, et al.2018↑ aggressiveness.
Bjørnebekk, et al.2017Lower cognitive functions, motor and executive performance; ↓ gray matter, cortical and putamen volume.
Bueno, et al.2017Neurodegeneration.
Turillazzi, et al.2016Neurodegeneration; hippocampus, striatum and frontal cortex injury.
Joukar, et al.2017Neurodegeneration.
Kaufman, et al.2015↑↑ right amygdala volume.
Piacentino, et al.2015↓ memory, ↑ aggressiveness, ↓ anxiety ↑ depression.
Gomes, et al.2014Neurodegeneration.
Basile, et al.2013Neurodegeneration.
Kanayama, et al.2013Cognitive dysfunction.
Elfverson, et al.2011↑ aggressiveness.
Melloni, Jr., et al.2010Anxiety-like behavior.
Kouvelas, et al.2008↓ memory, ↓ anxiety.
Sato, et al.2008↑ aggressiveness.
Rashid, et al.2007↓ memory, ↑ aggressiveness, ↓ anxiety ↑ depression.
Henderson, et al.2006behavioral effects.
Cardiovascular systemMarocolo, et al.2018Arrhythmic events; cardiac hypertrophy.
Rasmussen, et al.2018Hypertension; left ventricular hypertrophy.
Baggish, et al.2017↑ left ventricular mass index; ↑ left ventricular walls; myocardial mechanical dysfunction. myocardial infarction.
Seara, et al.2017↑ left ventricular mass index; ↑ left ventricular walls; myocardial mechanical dysfunction.
Christou, et al.2016Abnormal plasma lipoproteins.
Esperón, et al.2016Atherosclerotic plaque.
Vasilaki, et al.2016Cardiac injury
Akçakoyun et al.2014Arrhythmic events.
Angell, et at.2012Arrhythmic events; impairment of cardiac repolarization.
Golestani, et al.2012↓ plasma levels of HDL cholesterol; ↑ LDL cholesterol levels; arrhythmic events.
Chrostowski, et al.2011Increased synthesis of heart muscle, left ventricular hypertrophy and hypertension.
Riezzo, et al.2011Hypertension; left ventricular hypertrophy; pro-atherogenic effects; thrombosis.
Achar, et al.2010↓ plasma levels of HDL cholesterol; ↑ LDL cholesterol levels.
Alves, et al.2010Increased sympathetic activity.
D’Andrea, et al.2007Left ventricular dysfunction.
Phillis, et al.2007Arrhythmic events.
Chaves, et al.2006Oxidative stress.
Nottin, et al.2006↓ left ventricle relaxation properties.
Pereira-Junior, et al.2006Impairment of parasympathetic cardiac modulation; heart rate variability.
LiverSolimini, et al.2017Mitochondrial degeneration of hepatic cells.
Bond, et al.2016Hepatotoxicity.
Neri, et al.2011Hepatotoxicity.
Schwingel, et al.2011Liver fibrosis process.
Vieira, et al.2008Liver fibrosis process.
Urinary systemBrasil, et al.2015Kidney fibrosis.
Riezzo, et al.2014Oxidative stress; Accumulation of mesangial matrix.
D’Errico, et al.2011Accumulation of mesangial matrix.
Muscoloskeletal systemCarson, et al.2015↑ Muscle mass.
Marqueti, et al.2011Tendon damage.
Paschoal, et al.2009Skeletal muscles injuries.
Reproductive systemArmstrong, et al.2018Sexual dysfunctions.
Barone, et al.2017Spermatogenesis alteration.
El Osta, et al.2016Infertility; ↓ number of spermatidis.
García-Manso, et al.2016Changes in number, diameter and thickness of seminiferous tubules.
Pomara, et al.2015Impairment of testosterone production.
Rahnema, et al.2014Hypogonadism.
Coward, et al.2013Hypogonadism.
Hematologic consequencesChang, et al.2018↑ Thrombosis.
Casavant, et al.2007↑ hematocrit, ↑ erythrocytosis.
MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

Albano, G.D.; Amico, F.; Cocimano, G.; Liberto, A.; Maglietta, F.; Esposito, M.; Rosi, G.L.; Di Nunno, N.; Salerno, M.; Montana, A. Adverse Effects of Anabolic-Androgenic Steroids: A Literature Review. Healthcare 2021 , 9 , 97. https://doi.org/10.3390/healthcare9010097

Albano GD, Amico F, Cocimano G, Liberto A, Maglietta F, Esposito M, Rosi GL, Di Nunno N, Salerno M, Montana A. Adverse Effects of Anabolic-Androgenic Steroids: A Literature Review. Healthcare . 2021; 9(1):97. https://doi.org/10.3390/healthcare9010097

Albano, Giuseppe Davide, Francesco Amico, Giuseppe Cocimano, Aldo Liberto, Francesca Maglietta, Massimiliano Esposito, Giuseppe Li Rosi, Nunzio Di Nunno, Monica Salerno, and Angelo Montana. 2021. "Adverse Effects of Anabolic-Androgenic Steroids: A Literature Review" Healthcare 9, no. 1: 97. https://doi.org/10.3390/healthcare9010097

Article Metrics

Article access statistics, further information, mdpi initiatives, follow mdpi.

MDPI

Subscribe to receive issue release notifications and newsletters from MDPI journals

Select Your Interests

Customize your JAMA Network experience by selecting one or more topics from the list below.

  • Academic Medicine
  • Acid Base, Electrolytes, Fluids
  • Allergy and Clinical Immunology
  • American Indian or Alaska Natives
  • Anesthesiology
  • Anticoagulation
  • Art and Images in Psychiatry
  • Artificial Intelligence
  • Assisted Reproduction
  • Bleeding and Transfusion
  • Caring for the Critically Ill Patient
  • Challenges in Clinical Electrocardiography
  • Climate and Health
  • Climate Change
  • Clinical Challenge
  • Clinical Decision Support
  • Clinical Implications of Basic Neuroscience
  • Clinical Pharmacy and Pharmacology
  • Complementary and Alternative Medicine
  • Consensus Statements
  • Coronavirus (COVID-19)
  • Critical Care Medicine
  • Cultural Competency
  • Dental Medicine
  • Dermatology
  • Diabetes and Endocrinology
  • Diagnostic Test Interpretation
  • Drug Development
  • Electronic Health Records
  • Emergency Medicine
  • End of Life, Hospice, Palliative Care
  • Environmental Health
  • Equity, Diversity, and Inclusion
  • Facial Plastic Surgery
  • Gastroenterology and Hepatology
  • Genetics and Genomics
  • Genomics and Precision Health
  • Global Health
  • Guide to Statistics and Methods
  • Hair Disorders
  • Health Care Delivery Models
  • Health Care Economics, Insurance, Payment
  • Health Care Quality
  • Health Care Reform
  • Health Care Safety
  • Health Care Workforce
  • Health Disparities
  • Health Inequities
  • Health Policy
  • Health Systems Science
  • History of Medicine
  • Hypertension
  • Images in Neurology
  • Implementation Science
  • Infectious Diseases
  • Innovations in Health Care Delivery
  • JAMA Infographic
  • Law and Medicine
  • Leading Change
  • Less is More
  • LGBTQIA Medicine
  • Lifestyle Behaviors
  • Medical Coding
  • Medical Devices and Equipment
  • Medical Education
  • Medical Education and Training
  • Medical Journals and Publishing
  • Mobile Health and Telemedicine
  • Narrative Medicine
  • Neuroscience and Psychiatry
  • Notable Notes
  • Nutrition, Obesity, Exercise
  • Obstetrics and Gynecology
  • Occupational Health
  • Ophthalmology
  • Orthopedics
  • Otolaryngology
  • Pain Medicine
  • Palliative Care
  • Pathology and Laboratory Medicine
  • Patient Care
  • Patient Information
  • Performance Improvement
  • Performance Measures
  • Perioperative Care and Consultation
  • Pharmacoeconomics
  • Pharmacoepidemiology
  • Pharmacogenetics
  • Pharmacy and Clinical Pharmacology
  • Physical Medicine and Rehabilitation
  • Physical Therapy
  • Physician Leadership
  • Population Health
  • Primary Care
  • Professional Well-being
  • Professionalism
  • Psychiatry and Behavioral Health
  • Public Health
  • Pulmonary Medicine
  • Regulatory Agencies
  • Reproductive Health
  • Research, Methods, Statistics
  • Resuscitation
  • Rheumatology
  • Risk Management
  • Scientific Discovery and the Future of Medicine
  • Shared Decision Making and Communication
  • Sleep Medicine
  • Sports Medicine
  • Stem Cell Transplantation
  • Substance Use and Addiction Medicine
  • Surgical Innovation
  • Surgical Pearls
  • Teachable Moment
  • Technology and Finance
  • The Art of JAMA
  • The Arts and Medicine
  • The Rational Clinical Examination
  • Tobacco and e-Cigarettes
  • Translational Medicine
  • Trauma and Injury
  • Treatment Adherence
  • Ultrasonography
  • Users' Guide to the Medical Literature
  • Vaccination
  • Venous Thromboembolism
  • Veterans Health
  • Women's Health
  • Workflow and Process
  • Wound Care, Infection, Healing
  • Download PDF
  • Share X Facebook Email LinkedIn
  • Permissions

Steroid Side Effects

  • 1 JAMA , Chicago Illinois
  • 2 Midwestern University, Downers Grove, Illinois

Steroid medications, which are prescribed in many different forms for many different conditions, have a multitude of side effects.

Corticosteroid medications—often just called steroids by clinicians and patients—are used to reduce inflammation and inhibit the immune system. They are also associated with many side effects.

Corticosteroid medications are synthetic versions of the human steroid hormone cortisol , which is produced in the adrenal glands. These are different from the synthetic versions of the human steroid hormone testosterone used by some athletes (anabolic steroids) or the synthetic versions of the human steroid hormone estrogen used by some women after menopause (hormone therapy).

Formulations

Steroids can be taken as a tablet for simple rashes or mild asthma attacks or given intravenously for flares of autoimmune diseases such as inflammatory bowel disease or rheumatoid arthritis. To minimize the side effects of oral or intravenous steroids, steroid treatments that act locally were developed. Examples include

Topical application to the skin for conditions like eczema or psoriasis

Nasal inhalation for allergy symptoms

Inhalation into the lungs to control asthma symptoms

Injection into joints to reduce pain and inflammation

Eye drops to reduce swelling after eye surgery

Side Effects

Some patients find taking steroids to be difficult because of side effects; other patients like how steroids make them feel. Side effects are most common with oral or intravenous steroids, but sometimes enough locally directed steroid is absorbed systemically to cause side effects. Life-threatening side effects include

Infection: Steroids are effective in treating autoimmune diseases because they reduce the ability of the immune system to function ( immunosuppression ). Patients taking steroids are not only more susceptible to infections but more likely to have severe or unusual infections. These patients should be aware of their increased risk of infection, and their physicians may recommend additional anti-infective medications.

Adrenal crisis: Cortisol is produced in the adrenal glands. It has many effects throughout the body, including regulating blood pressure. Because steroids are so similar to cortisol, prolonged use of systemic steroids at higher doses can cause the adrenal glands to stop making cortisol. If the systemic steroid is stopped suddenly, this adrenal suppression and resulting lack of steroid can cause a wide range of symptoms, such as dangerously low blood pressure.

Health care practitioners are cautious in prescribing steroids because of the side effects. They prescribe them only when necessary and for as short a time as possible. Local rather than systemic therapy is preferable and prescribed when possible. If a patient needs to stop taking a systemic steroid after taking it for a long time, they are prescribed a gradually reduced dose to give the adrenal glands time to “wake up” and start producing cortisol again. When longer courses of higher-dose systemic steroids are necessary, as in some autoimmune conditions, the patient is monitored closely for side effects.

For More Information

US National Library of Medicine medlineplus.gov/steroids.html

Conflict of Interest Disclosures: None reported.

Source: Zoorob RJ, Cender D. A different look at corticosteroids. Am Fam Physician . 1998;58(2):443-450.

See More About

Grennan D , Wang S. Steroid Side Effects. JAMA. 2019;322(3):282. doi:10.1001/jama.2019.8506

Manage citations:

© 2024

Artificial Intelligence Resource Center

Cardiology in JAMA : Read the Latest

Browse and subscribe to JAMA Network podcasts!

Others Also Liked

  • Register for email alerts with links to free full-text articles
  • Access PDFs of free articles
  • Manage your interests
  • Save searches and receive search alerts

Thank you for visiting nature.com. You are using a browser version with limited support for CSS. To obtain the best experience, we recommend you use a more up to date browser (or turn off compatibility mode in Internet Explorer). In the meantime, to ensure continued support, we are displaying the site without styles and JavaScript.

  • View all journals
  • Explore content
  • About the journal
  • Publish with us
  • Sign up for alerts
  • Review Article
  • Published: 02 September 2024

Systemic steroids and bronchopulmonary dysplasia: a systematic review and meta-analysis

  • Talkad S. Raghuveer   ORCID: orcid.org/0000-0001-6482-9684 1 ,
  • Rosey E. Zackula   ORCID: orcid.org/0000-0003-2439-8714 2 ,
  • Richa Lakhotia 1 , 3 &
  • Stephanie A. Binder 1 , 3  

Journal of Perinatology ( 2024 ) Cite this article

1 Altmetric

Metrics details

  • Medical research

It is unclear if systemic steroids decrease the risk of Bronchopulmonary Dysplasia (BPD) while increasing the risk of neurodevelopmental impairment (NDI).

Conduct a systematic review of randomized controlled trials of systemic steroids to evaluate the risk of BPD, mortality, and NDI in premature infants ≤30 weeks.

Data sources

MEDLINE, EBSCOhost, Web of Science, Cochrane Library, Embase, and CINAHL.

Study selection

Randomized clinical trials of Dexamethasone (DEX) or Hydrocortisone (HC) to prevent BPD in premature infants ≤ 30 weeks.

Data extraction and synthesis

Data were extracted using the Preferred Reporting Items for Systematic Reviews and Meta-Analysis guidelines. Random-effects meta-analyses and multivariable meta-regression were conducted.

Main outcomes and measures

Primary outcomes were BPD, mortality, and NDI. Secondary outcomes were hypertension, hyperglycemia, sepsis, intestinal perforation, necrotizing enterocolitis (NEC), and retinopathy of prematurity (ROP). The a priori hypothesis was that steroids would reduce the risk of BPD without increasing NDI.

There were 6377 preterm infants in the 44 (32 DEX, 13 HC) selected studies. DEX significantly reduced the risk of BPD, RR = 0.66, (95% CI, 0.56–0.78). The most effective DEX regimen was medium cumulative dose (2 to 3 mg/kg), RR = 0.43 (95% CI, 0.29–0.65); day of initiation <8 days: RR = 0.68, (95% CI, 0.59–0.79); and treatment for ≥14 days: RR = 0.67 (95% CI, 0.55–0.80). HC did not significantly decrease the risk of BPD, RR = 0.98, (95% CI, 0.87–1.10). Neither DEX, (RR = 0.92, 95% CI, 0.78–1.09) nor HC (RR = 0.83, 95% CI, 0.68–1.01) decrease the risk of mortality. The risk of CP was not increased by either DEX (RR = 1.09, 95% CI, 0.55–2.17) or HC (RR = 1.18, 95% CI, 0.75–1.87). There were no significant differences between steroids and placebo for MDI/PDI scores. Multivariable meta-regression models showed that DEX significantly reduced the risk of BPD without increased risk of CP. DEX increased the risk of hypertension and hyperglycemia. Studies showed high heterogeneity, differing treatment regimen, missing data and different rates of follow-up.

Conclusion and relevance

DEX, but not HC, significantly decreased the risk of BPD. Neither steroid showed an increased risk of NDI or mortality.

This is a preview of subscription content, access via your institution

Access options

Subscribe to this journal

Receive 12 print issues and online access

251,40 € per year

only 20,95 € per issue

Buy this article

  • Purchase on SpringerLink
  • Instant access to full article PDF

Prices may be subject to local taxes which are calculated during checkout

research articles about steroids

Similar content being viewed by others

research articles about steroids

Timing of postnatal steroids for bronchopulmonary dysplasia: association with pulmonary and neurodevelopmental outcomes

research articles about steroids

The beneficial effect of prophylactic hydrocortisone treatment in extremely preterm infants improves upon adjustment of the baseline characteristics

research articles about steroids

Risk factors of early pulmonary hypertension and its clinical outcomes in preterm infants: a systematic review and meta-analysis

Stoll BJ, Hansen NI, Bell EF, Walsh MC, Carlo WA, Shankaran S. et al. Trends in care practices, morbidity, and mortality of extremely preterm neonates, 1993-2012. JAMA. 2015;314:1039–51. https://doi.org/10.1001/jama.2015.10244 .

Article   CAS   PubMed   PubMed Central   Google Scholar  

Bell EF, Hintz SR, Hansen NI, Bann CM, Wyckoff MH, DeMauro SB, et al. Mortality, in-hospital morbidity, care practices, and 2-year outcomes for extremely preterm infants in the US, 2013-2018. JAMA. 2022;327:248–63. https://doi.org/10.1001/jama.2021.23580 .

Article   PubMed   Google Scholar  

Jensen EA, Edwards EM, Greenberg LT, Soll RF, Ehret DEY, Horbar JD. Severity of bronchopulmonary dysplasia among very preterm infants in the United States. Pediatrics. 2021;148. https://doi.org/10.1542/peds.2020-030007 .

Hwang JS, Rehan VK. Recent advances in bronchopulmonary dysplasia: pathophysiology, prevention, and treatment. Lung. 2018;196:129–38. https://doi.org/10.1007/s00408-018-0084-z .

Article   PubMed   PubMed Central   Google Scholar  

Levy PT, Dioneda B, Holland MR, Sekarski TJ, Lee CK, Mathur A, et al. Right ventricular function in preterm and term neonates: reference values for right ventricle areas and fractional area of change. J Am Soc Echocardiogr. 2015;28:559–69. https://doi.org/10.1016/j.echo.2015.01.024 .

Bui CB, Pang MA, Sehgal A, Theda C, Lao JC, Berger PJ, et al. Pulmonary hypertension associated with bronchopulmonary dysplasia in preterm infants. J Reprod Immunol. 2017;124:21–29. https://doi.org/10.1016/j.jri.2017.09.013 .

Donda K, Agyemang CO, Adjetey NA, Agyekum A, Princewill N, Ayensu M, et al. Tracheostomy trends in preterm infants with bronchopulmonary dysplasia in the United States: 2008-2017. Pediatr Pulmonol. 2021;56:1008–17. https://doi.org/10.1002/ppul.25273 .

Collaco JM, McGrath-Morrow SA. Respiratory phenotypes for preterm infants, children, and adults: bronchopulmonary dysplasia and more. Ann Am Thorac Soc. 2018;15:530–8. https://doi.org/10.1513/AnnalsATS.201709-756FR .

Manimtim WM, Agarwal A, Alexiou S, Levin JC, Aoyama B, Austin ED, et al. Respiratory outcomes for ventilator-dependent children with bronchopulmonary dysplasia. Pediatrics. 2023;151. https://doi.org/10.1542/peds.2022-060651 .

Cheong JLY, Doyle LW. An update on pulmonary and neurodevelopmental outcomes of bronchopulmonary dysplasia. Semin Perinatol. 2018;42:478–84. https://doi.org/10.1053/j.semperi.2018.09.013 .

Schmidt B, Asztalos EV, Roberts RS, Robertson CM, Sauve RS, Whitefield MF. Impact of bronchopulmonary dysplasia, brain injury, and severe retinopathy on the outcome of extremely low-birth-weight infants at 18 months: results from the trial of indomethacin prophylaxis in preterms. JAMA. 2003;289:1124–9. https://doi.org/10.1001/jama.289.9.1124 .

Sriram S, Schreiber MD, Msall ME, Kuban KCK, Joseph RM, O’Shea TM, et al. Cognitive development and quality of life associated with BPD in 10-year-olds born preterm. Pediatrics. 2018;141:e20172719. https://doi.org/10.1542/peds.2017-2719 .

Doyle LW, Ranganathan S, Mainzer RM, Cheong JLY. Victorian infant collaborative study G. Relationships of severity of bronchopulmonary dysplasia with adverse neurodevelopmental outcomes and poor respiratory function at 7-8 years of age. J Pediatr. 2024;269:114005. https://doi.org/10.1016/j.jpeds.2024.114005 .

Jensen EA, Dysart K, Gantz MG, McDonald S, Bamat NA, Keszler M, et al. The diagnosis of bronchopulmonary dysplasia in very preterm infants. an evidence-based approach. Am J Respir Crit Care Med. 2019;200:751–9. https://doi.org/10.1164/rccm.201812-2348OC .

Mammel MC, Green TP, Johnson DE, Thompson TR. Controlled trial of dexamethasone therapy in infants with bronchopulmonary dysplasia. Lancet. 1983;1:1356–8. https://doi.org/10.1016/s0140-6736(83)92139-6 .

Article   CAS   PubMed   Google Scholar  

American Academy of Pediatrics. Committee on fetus and newborn. Postnatal corticosteroids to treat or prevent chronic lung disease in preterm infants. Pediatrics. 2002;109:330–8. https://doi.org/10.1542/peds.109.2.330 .

Article   Google Scholar  

Walsh MC, Yao Q, Horbar JD, Carpenter JH, Lee SK, Ohlsson A. Changes in the use of postnatal steroids for bronchopulmonary dysplasia in 3 large neonatal networks. Pediatrics. 2006;118:e1328–35. https://doi.org/10.1542/peds.2006-0359 .

Yoder BA, Harrison M, Clark RH. Time-related changes in steroid use and bronchopulmonary dysplasia in preterm infants. Pediatrics. 2009;124:673–9. https://doi.org/10.1542/peds.2008-2793 .

Abiramalatha T, Ramaswamy VV, Bandyopadhyay T, Somanath SH, Shaik NB, Pullattayil AK, et al. Interventions to prevent bronchopulmonary dysplasia in preterm neonates: an umbrella review of systematic reviews and meta-analyses. JAMA Pediatr. 2022;176:502–16. https://doi.org/10.1001/jamapediatrics.2021.6619 .

Ramaswamy VV, Bandyopadhyay T, Nanda D, Bandiya P, Ahmed J, Garg A, et al. Assessment of postnatal corticosteroids for the prevention of bronchopulmonary dysplasia in preterm neonates: a systematic review and network meta-analysis. JAMA Pediatr. 2021;175:e206826. https://doi.org/10.1001/jamapediatrics.2020.6826 .

Doyle LW, Davis PG, Morley CJ, McPhee A, Carlin JB. Low-dose dexamethasone facilitates extubation among chronically ventilator-dependent infants: a multicenter, international, randomized, controlled trial. Pediatrics. 2006;117:75–83. https://doi.org/10.1542/peds.2004-2843 .

Job S, Clarke P. Current UK practices in steroid treatment of chronic lung disease. Arch Dis Child Fetal Neonatal Ed. 2015;100:F371. https://doi.org/10.1136/archdischild-2014-308060 .

Cummings JJ, Pramanik AK. Postnatal corticosteroids to prevent or treat chronic lung disease following preterm birth. Pediatrics. 2022;149:e2022057530. https://doi.org/10.1542/peds.2022-057530 .

Page MJ, Moher D, Bossuyt PM, Boutron I, Hoffmann TC, Mulrow CD, et al. PRISMA 2020 explanation and elaboration: updated guidance and exemplars for reporting systematic reviews. BMJ. 2021;372:n160. https://doi.org/10.1136/bmj.n160 .

Onland W, Cools F, Kroon A, Ramful D, El Moussawi F, Nicaise C, et al. Effect of hydrocortisone therapy initiated 7 to 14 days after birth on mortality or bronchopulmonary dysplasia among very preterm infants receiving mechanical ventilation: a randomized clinical trial. JAMA. 2019;321:354–363. https://doi.org/10.1001/jama.2018.21443 .

Doyle LW, Cheong JL, Ehrenkranz RA, Halliday HL. Late (>7 days) systemic postnatal corticosteroids for prevention of bronchopulmonary dysplasia in preterm infants. Cochrane Database Syst Rev. 2017;10:CD001145. https://doi.org/10.1002/14651858.CD001145.pub4 .

Baud O, Biran V, Trousson C, Leroy E, Mohamed D, Alberti C. Two-year outcomes after prophylactic hydrocortisone in extremely preterm neonates. EAPS Congress 2016. Eur J Pediatr. 2016;175:1393–880. https://doi.org/10.1007/s00431-016-2785-8 .

Clauss C, Thomas S, Khodak I, Tack V, Akerman M, Hanna N, et al. Hydrocortisone and bronchopulmonary dysplasia: variables associated with response in premature infants. J Perinatol. 2020;40:1349–57. https://doi.org/10.1038/s41372-020-0680-7 .

Htun ZT, Schulz EV, Desai RK, Marasch JL, McPherson CC, Mastrandrea LD. et al. Postnatal steroid management in preterm infants with evolving bronchopulmonary dysplasia. J Perinatol. 2021;41:1783–96. https://doi.org/10.1038/s41372-021-01083-w .

Sterne J, Savović J, Page M, Elbers R, Blencowe N, Boutron I, et al. RoB 2: a revised tool for assessing risk of bias in randomised trials. BMJ. 2019;366:14898. https://doi.org/10.1136/bmj.14898 .

GRADEpro Guideline Development Tool [Software]. McMaster University and Evidence Prime, 2022. Available from gradepro.org. 2022.

Viechtbauer W. Conducting meta-analyses in R with metafor package. J Stat Softw. 2010;36:1–48.

Harrer M, Cuijpers P, Furukawa TA, Ebert DD. Doing meta-analysis with R: a hands-on guide. 1st ed. Chapman & Hall/CRC Press; 2021.

Borenstein M, Hedges LV, Higgins JP, Rothstein HR. A basic introduction to fixed-effect and random-effects models for meta-analysis. Res Synth Methods. 2010;1:97–111. https://doi.org/10.1002/jrsm.12 .

Christiansen S, Iverson C, Flanagin A, Livingston EH, Fishcer L, Manno C, et al. AMA manual of style: A guide for authors and editors. American Medical Association manual of style. 11th edition. New York, NY: Oxford University Press; 2020.

Doyle LW, Cheong JL, Hay S, Manley BJ, Halliday HL. Early (< 7 days) systemic postnatal corticosteroids for prevention of bronchopulmonary dysplasia in preterm infants. Cochrane Database Syst Rev. 2021;10:Cd001146. https://doi.org/10.1002/14651858.CD001146.pub6 .

Doyle LW, Cheong JL, Hay S, Manley BJ, Halliday HL. Late (≥ 7 days) systemic postnatal corticosteroids for prevention of bronchopulmonary dysplasia in preterm infants. Cochrane Database Syst Rev. 2021;11:Cd001145. https://doi.org/10.1002/14651858.CD001145.pub5 .

Zeng LN, Tian JH, Song FJ, Li WR, Jiang LC, Gui G, et al. Corticosteroids for the prevention of bronchopulmonary dysplasia in preterm infants: a network meta-analysis. Arch Dis Child Fetal Neonatal Ed. 2018;103:F506–11. https://doi.org/10.1136/archdischild-2017-313759 .

Hay S, Ovelman C, Zupancic JA, Doyle LW, Onland W, Konstantinidis M, et al. Systemic corticosteroids for the prevention of bronchopulmonary dysplasia, a network meta-analysis. Cochrane Database Syst Rev. 2023;8:CD013730. https://doi.org/10.1002/14651858.CD013730.pub2 .

van de Loo M, van Kaam A, Offringa M, Doyle LW, Cooper C, Onland W. Corticosteroids for the prevention and treatment of bronchopulmonary dysplasia: an overview of systemic reviews. Cochrane Database Syst Rev. 2024:CD013271. https://doi.org/10.1002/14651858.CD013271.pub2 .

Onland W, Offringa M, Jaegere APD, Van Kaam AH. Finding the optimal postnatal dexamethasone regimen for preterm infants at risk of bronchopulmonary dysplasia: a systematic review of placebo-controlled trials. Pediatrics. 2009;123:367–77. https://doi.org/10.1542/peds.2008-0016 .

Doyle LW, Halliday HL, Ehrenkranz RA, Davis PG, Sinclair JC. An update on the impact of postnatal systemic corticosteroids on mortality and cerebral palsy in preterm infants: effect modification by risk of bronchopulmonary dysplasia. J Pediatr. 2014;165:1258–60. https://doi.org/10.1016/j.jpeds.2014.07.049 .

Greenberg RG, McDonald SA, Laughon MM, Tanaka D, Jensen E, Van Meurs K, et al. Online clinical tool to estimate risk of bronchopulmonary dysplasia in extremely preterm infants. Arch Dis Child Fetal Neonatal Ed. 2022. https://doi.org/10.1136/archdischild-2021-323573 .

Shimotsuma T, Tomotaki S, Akita M, Araki R, Tomotaki H, Iwanaga K, et al. Severe Bronchopulmonary Dysplasia adversely affects brain growth in preterm infants. Neonatology. 2024. https://doi.org/10.1159/000538527 .

Download references

Acknowledgements

The authors thank the contributions of Hayrettin Okut, PhD, who advised us on statistical aspects of model development, including programming in R.

Author information

Authors and affiliations.

Department of Pediatrics, University of Kansas School of Medicine—Wichita, Wichita, KS, USA

Talkad S. Raghuveer, Richa Lakhotia & Stephanie A. Binder

Office of Research, University of Kansas School of Medicine—Wichita, Wichita, KS, USA

Rosey E. Zackula

Pediatrix Medical Group of Kansas, Wichita, KS, USA

Richa Lakhotia & Stephanie A. Binder

You can also search for this author in PubMed   Google Scholar

Contributions

TSR: conceptualized, formulated the research methodology, wrote the original draft, reviewed and edited the manuscript, supervised/administered the whole project and guarantor for the project. REZ: conceptualized, formulated the research methodology, performed formal analysis, curated data and wrote the original draft, reviewed and edited the manuscript. RL: conceptualized, formulated the research methodology, wrote the original draft, reviewed and edited the manuscript. SAB: conceptualized, formulated the research methodology, wrote the original draft, reviewed and edited the manuscript. All authors approved the final manuscript as submitted and agree to be accountable for all aspects of the work.

Corresponding author

Correspondence to Talkad S. Raghuveer .

Ethics declarations

Competing interests.

TSR, REZ, RL, and SAB have no relevant conflicts to disclose. This study was not supported by funding from any source.

Additional information

Publisher’s note Springer Nature remains neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Supplementary information

Supplemental materials, rights and permissions.

Springer Nature or its licensor (e.g. a society or other partner) holds exclusive rights to this article under a publishing agreement with the author(s) or other rightsholder(s); author self-archiving of the accepted manuscript version of this article is solely governed by the terms of such publishing agreement and applicable law.

Reprints and permissions

About this article

Cite this article.

Raghuveer, T.S., Zackula, R.E., Lakhotia, R. et al. Systemic steroids and bronchopulmonary dysplasia: a systematic review and meta-analysis. J Perinatol (2024). https://doi.org/10.1038/s41372-024-02097-w

Download citation

Received : 06 June 2024

Revised : 29 July 2024

Accepted : 16 August 2024

Published : 02 September 2024

DOI : https://doi.org/10.1038/s41372-024-02097-w

Share this article

Anyone you share the following link with will be able to read this content:

Sorry, a shareable link is not currently available for this article.

Provided by the Springer Nature SharedIt content-sharing initiative

Quick links

  • Explore articles by subject
  • Guide to authors
  • Editorial policies

research articles about steroids

  • Alzheimer's disease & dementia
  • Arthritis & Rheumatism
  • Attention deficit disorders
  • Autism spectrum disorders
  • Biomedical technology
  • Diseases, Conditions, Syndromes
  • Endocrinology & Metabolism
  • Gastroenterology
  • Gerontology & Geriatrics
  • Health informatics
  • Inflammatory disorders
  • Medical economics
  • Medical research
  • Medications
  • Neuroscience
  • Obstetrics & gynaecology
  • Oncology & Cancer
  • Ophthalmology
  • Overweight & Obesity
  • Parkinson's & Movement disorders
  • Psychology & Psychiatry
  • Radiology & Imaging
  • Sleep disorders
  • Sports medicine & Kinesiology
  • Vaccination
  • Breast cancer
  • Cardiovascular disease
  • Chronic obstructive pulmonary disease
  • Colon cancer
  • Coronary artery disease
  • Heart attack
  • Heart disease
  • High blood pressure
  • Kidney disease
  • Lung cancer
  • Multiple sclerosis
  • Myocardial infarction
  • Ovarian cancer
  • Post traumatic stress disorder
  • Rheumatoid arthritis
  • Schizophrenia
  • Skin cancer
  • Type 2 diabetes
  • Full List »

share this!

September 6, 2024

This article has been reviewed according to Science X's editorial process and policies . Editors have highlighted the following attributes while ensuring the content's credibility:

fact-checked

peer-reviewed publication

trusted source

Steroids found in Scottish prisons increase from 1% to 10% in four years

by University of Dundee

Steroids found in Scottish prisons increase from 1% to 10% in four years

Researchers at the University's Leverhulme Research Center of Forensic Science (LRCFS), part of the School of Science and Engineering, made the discovery while working in collaboration with the Scottish Prison Service (SPS).

Their findings have been reported in a new publication , "Changing trends in anabolic-androgenic steroid use within Scottish prisons," in the journal Drug Testing and Analysis .

The researchers analyzed 3,896 suspected drug samples , all of which had been seized at Scottish prisons between January 2019 and August 2023.

Of the samples which were analyzed in 2019, less than 1% were found to contain anabolic-androgenic steroids (AAS), compared to more than 10% in 2023.

Various types of AAS were found and they were the third most common drug detected in Scottish prisons in 2023.

Most of these steroid drugs (77%) were in tablet form, of various vivid colors.

AAS compounds were also found in powders, herbal material, a fragmented soap bar-type sample, and within vapes. In many cases, ASS was found along with other illicit substances.

Steroids found in Scottish prisons increase from 1% to 10% in four years

Dr. Lorna Nisbet, senior lecturer at LRCFS who was involved in the study, said, "The research shows a significant rise in steroid compounds within prisons, and they are being detected in forms we would not typically expect, such as herbal material and vapes.

"Lots of these materials contain a combination of different drugs, in varying amounts, making it difficult for individuals to know exactly what they are taking or at what dosage.

"By consuming these drugs, individuals may be unknowingly engaging in polydrug use and the effects of these drugs when taken in combination with others can be particularly problematic."

Polydrug use—taking more than one substance at a time—was prevalent in 81% of all drug misuse fatalities in 2023, new data on drug-related deaths recently released by National Records of Scotland revealed.

"Polydrug use, can increase the toxic effects of drugs, prolong a drug's effects on the person and increase negative side effects ," Lorna said.

"It is important that people working within prison settings know what drugs are potentially being used within the prison to allow them to provide the necessary support and address emerging risks more effectively.

"That is one of the reasons behind our collaborative work with SPS as it has allowed for the quick identification of changing drug use patterns and trends within the Scottish Prison landscape."

Lorna added, "Steroids are not routinely tested for in the United Kingdom but this data suggests that their use may be on the rise, and that more monitoring of these drugs may be needed."

Explore further

Feedback to editors

research articles about steroids

Low-impact yoga and exercise found to help older women manage urinary incontinence

20 hours ago

research articles about steroids

Missouri patient tests positive for bird flu despite no known exposure to animals

research articles about steroids

Falling for financial scams? It may signal early Alzheimer's disease

Sep 6, 2024

research articles about steroids

Cognitive behavioral therapy enhances brain circuits to relieve depression

research articles about steroids

New molecular sensor enables fluorescence imaging for assessing sarcoma severity

research articles about steroids

Noninvasive focused ultrasound show potential for combating chronic pain

research articles about steroids

Study finds TGF-beta and RAS signaling are both required for lung cancer metastasis

research articles about steroids

Research team successfully maps the brain-spinal cord connection in humans

research articles about steroids

Alzheimer's study reveals critical differences in memory loss progression based on the presence of specific proteins

research articles about steroids

Chemical screen identifies PRMT5 as therapeutic target for paclitaxel-resistant triple-negative breast cancer

Related stories.

research articles about steroids

Mandatory drug testing in prisons is ineffective and harmful, says researcher

Jul 8, 2024

research articles about steroids

Risks faced by women using anabolic steroids in Australia

Aug 28, 2024

research articles about steroids

UK government moves to ban 'zombie drug' xylazine

Sep 4, 2024

Drug-related deaths in Scotland reach record high

Dec 15, 2020

Anabolic steroids linked to higher rates of premature death in men

Nov 21, 2018

research articles about steroids

Steroid use associated with serious side effects among adolescents and young adults

Nov 21, 2022

Recommended for you

research articles about steroids

Public health researchers find decriminalization of drug possession was not associated with Oregon overdose spike

Sep 5, 2024

research articles about steroids

Fentanyl vaccine heads for clinical trials, with goal of saving lives

Sep 2, 2024

research articles about steroids

Mechanisms of how morphine relieves pain mapped out

Aug 29, 2024

research articles about steroids

Access to opioid agonist treatment in prisons saves lives, researchers say

research articles about steroids

Machine learning predicts which patients will continue taking opioids after hand surgery

research articles about steroids

Researchers find e-cigarette use disrupts the nasal microbiome

Aug 26, 2024

Let us know if there is a problem with our content

Use this form if you have come across a typo, inaccuracy or would like to send an edit request for the content on this page. For general inquiries, please use our contact form . For general feedback, use the public comments section below (please adhere to guidelines ).

Please select the most appropriate category to facilitate processing of your request

Thank you for taking time to provide your feedback to the editors.

Your feedback is important to us. However, we do not guarantee individual replies due to the high volume of messages.

E-mail the story

Your email address is used only to let the recipient know who sent the email. Neither your address nor the recipient's address will be used for any other purpose. The information you enter will appear in your e-mail message and is not retained by Medical Xpress in any form.

Newsletter sign up

Get weekly and/or daily updates delivered to your inbox. You can unsubscribe at any time and we'll never share your details to third parties.

More information Privacy policy

Donate and enjoy an ad-free experience

We keep our content available to everyone. Consider supporting Science X's mission by getting a premium account.

E-mail newsletter

  • +44 7418625617

Annals of Clinical Trials and Vaccines Research

Research Article - Annals of Clinical Trials and Vaccines Research (2023) Volume 13, Issue 4

  • Download PDF

The Role of Steroids in Clinical Practice: Benefits, Risks, and Considerations for Therapeutic Use

Walter Lewis *

Department of Pharmaceutical Chemistry, Medical University of Graz, Austria

*Corresponding Author: Walter Lewis Department of Pharmaceutical Chemistry, Medical University of Graz, Austria E-mail: [email protected]

Received: 01-August-2023, Manuscript No. actvr-23-108542; Editor assigned: 3-August-2023, PreQC No. actvr-23-108542 (PQ); Reviewed: 17-August-2023, QC No. actvr-23-108542; Revised: 22-August-2023, Manuscript No. actvr-23-108542 (R); Published: 28-August-2023; DOI: 10.37532/ ACTVR.2023.13(4).128-131

Steroids, a class of organic compounds, have long been an essential component of clinical practice due to their potent anti-inflammatory and immunosuppressive properties. This article provides an in-depth analysis of the various applications of steroids in medical settings, exploring their benefits, risks, and crucial considerations for therapeutic use. From managing chronic inflammatory conditions to addressing acute emergencies, steroids have proven to be valuable tools in the hands of healthcare professionals. However, their side effects and potential for misuse demand a nuanced approach to prescribing and monitoring. This review aims to equip clinicians with up-to-date knowledge to make informed decisions regarding steroid therapy, ensuring optimal patient outcomes while minimizing adverse effects.

Biomarkers • Personalized medicine • Adaptive trial designs • Artificial intelligence • Regulatory approvals

Introduction

Steroids, also known as corticosteroids or glucocorticoids, are synthetic drugs that mimic the actions of naturally occurring hormones in the body. In clinical practice, steroids have a wide range of therapeutic applications due to their potent anti-inflammatory, immunosuppressive, and metabolic effects. This article provides a comprehensive overview of the various uses of steroids in medical practice, highlighting their benefits and potential side effects. Steroids have been a mainstay in clinical medicine for decades, and their applications continue to evolve as new research sheds light on their mechanisms of action. This article delves into the history and development of steroids as therapeutic agents and explores their role in managing various medical conditions. Steroids remain indispensable in clinical practice, playing a vital role in managing a wide range of medical conditions [ 1 , 2 ].

Their potent anti-inflammatory and immunosuppressive effects make them valuable tools for alleviating symptoms and improving patients’ quality of life. However, healthcare professionals must carefully consider the risks and benefits of steroid therapy, tailoring treatment plans to individual patient needs. Continued research and advancements in steroid therapy are likely to refine their applications in the future. Steroids, a class of potent anti-inflammatory and immunosuppressive agents, have long played a crucial role in various clinical specialties. This comprehensive review article explores the diverse therapeutic applications of steroids in clinical practice and highlights the associated considerations and potential risks. From managing autoimmune disorders and allergic reactions to addressing various inflammatory conditions, steroids have proven their efficacy across a wide range of medical conditions [ 3 , 4 ].

Material & Methods

The introduction provides a historical overview of the discovery and development of steroids, outlining their initial application and subsequent expansion into various medical fields. It emphasizes the significance of understanding the pharmacology and mechanism of action of steroids to optimize their use in clinical settings. This section delves into the pharmacokinetics and pharmacodynamics of steroids, explaining how different formulations and routes of administration can influence their efficacy and side effect profile. The focus is on the varying potency of different steroid compounds and the importance of tailoring treatments to individual patient needs.

The article emphasizes the importance of weighing the potential benefits against the risks when prescribing steroids. It addresses common side effects, such as weight gain, osteoporosis, mood changes, and increased susceptibility to infections. Additionally, special considerations are discussed, including steroid tapering, alternate-day dosing, and monitoring patients on longterm steroid therapy.

The conclusion highlights the significant impact of steroids on clinical practice and patient outcomes, while also advocating for a balanced and cautious approach to their use. It emphasizes the need for further research and continued efforts to refine steroid therapies to maximize their benefits while minimizing adverse effects. Steroids, also known as corticosteroids or glucocorticoids, have been a cornerstone in the management of various medical conditions for decades. These powerful anti-inflammatory agents have demonstrated remarkable efficacy in treating a wide range of diseases, spanning from autoimmune disorders to respiratory illnesses. This article aims to provide an indepth exploration of the use of steroids in clinical practice, shedding light on their mechanisms of action, indications, dosing strategies, potential side effects, and the latest research trends.

This section delves into the mechanisms through which steroids exert their effects on the immune system and inflammatory response. It discusses their impact on gene expression, cytokine modulation, and inhibition of immune cell activation. Additionally, the article examines how steroids interfere with the synthesis of prostaglandins and Leukotrienes, influencing the overall inflammatory cascade. Autoimmune diseases such as rheumatoid arthritis, systemic lupus erythematous, and multiple sclerosis can lead to debilitating symptoms. This part of the article explores the role of steroids as an essential component of the treatment regimen for these conditions. It also highlights the importance of early intervention and the balance between symptomatic relief and long-term management.

Respiratory conditions, including asthma and chronic obstructive pulmonary disease (COPD), often involve airway inflammation. The article examines the use of steroids in these cases, focusing on their ability to reduce airway inflammation, improve lung function, and prevent exacerbations. Dermatological conditions like eczema, psoriasis, and allergic reactions can significantly impact a patient’s quality of life. This section explores the role of topical and systemic steroids in managing these conditions, emphasizing the importance of appropriate application and potential adverse effects on the skin. The correct dosing of steroids is crucial to achieve therapeutic benefits while minimizing side effects. This part of the article discusses various dosing strategies for different conditions, as well as the importance of close patient monitoring during treatment to detect and manage potential complications.

Steroids are associated with a range of side effects, including weight gain, mood changes, osteoporosis, and increased susceptibility to infections. This section explores strategies to mitigate these side effects, such as lifestyle modifications, adjunctive therapies, and dose-tapering techniques.

As medical research progresses, new insights into the use of steroids continue to emerge. This final section of the article highlights recent developments and ongoing studies, including novel formulations, targeted therapies, and potential alternative treatments. Steroids play a vital role in clinical practice, offering effective management for various medical conditions. However, their use requires careful consideration, weighing the benefits against potential risks. By understanding their mechanisms of action, appropriate indications, dosing strategies, and monitoring, healthcare professionals can optimize the use of steroids and enhance patient outcomes. Ongoing research and advancements in this field promise a brighter future for patients who rely on steroids as part of their therapeutic journey [ 5 - 7 ].

Steroids, a class of potent anti-inflammatory agents, have been an indispensable component of clinical practice for several decades. This comprehensive review aims to explore the diverse therapeutic applications of steroids and their significant role in managing various medical conditions. From their use as powerful immunosuppressive agents to their essential role in treating hormonal imbalances, steroids have proven to be invaluable in enhancing patient outcomes. Tailoring treatment plans to individual patient needs. Continued research and advancements in steroid therapy are likely to refine their applications in the future. Steroids, a class of potent anti-inflammatory and immunosuppressive agents, have long played a crucial role in various clinical specialties. This comprehensive review article explores the diverse therapeutic applications of steroids in clinical practice and highlights the associated considerations and potential risks. From managing autoimmune disorders and allergic reactions to addressing various inflammatory conditions, steroids have proven their efficacy across a wide range of medical conditions

Steroids, also known as corticosteroids or glucocorticoids, have emerged as an essential pharmacological tool in clinical practice. Originally introduced for their profound anti-inflammatory effects, steroids have since demonstrated their versatility across a wide range of medical disciplines. This review examines the mechanisms of action, indications, and potential side effects of steroids, providing valuable insights for clinicians to optimize patient care. Steroids play a pivotal role in the management of various immunological disorders, including autoimmune diseases and organ transplantations. The article delves into the underlying immunomodulatory mechanisms of steroids, discussing their efficacy in suppressing immune responses and mitigating inflammatory cascades. The correct dosing of steroids is crucial to achieve therapeutic benefits while minimizing side effects. This part of the article discusses various dosing strategies for different conditions, as well as the importance of close patient monitoring during treatment to detect and manage potential complications.

A major area of clinical application for steroids lies in the treatment of respiratory conditions. The review evaluates the use of steroids in asthma, chronic obstructive pulmonary disease (COPD), and other respiratory disorders, outlining their effectiveness in reducing airway inflammation and improving lung function. The management of rheumatic diseases, such as rheumatoid arthritis, systemic lupus erythematous, and vasculitis, has been revolutionized by the introduction of steroids. This section investigates the role of steroids in alleviating joint inflammation and preventing disease progression However; their use requires careful consideration, weighing the benefits against potential risks. By understanding their mechanisms of action, appropriate indications, dosing strategies, and monitoring, healthcare professionals can optimize the use of steroids and enhance patient outcomes [ 8 - 10 ].

Steroids are vital in endocrinology for their ability to supplement hormonal deficiencies and regulate various physiological processes. The article discusses their use in adrenal insufficiency, Addison’s disease, and other endocrine disorders, emphasizing the need for precise dosing and monitoring to avoid adverse effects. Dermatological conditions, including eczema, psoriasis, and severe allergic reactions, often require steroids for their anti-inflammatory and immunosuppressive properties. The review examines the different formulations and routes of administration used in dermatology and highlights potential complications associated with long-term usage.

While steroids offer substantial therapeutic benefits, their prolonged use or inappropriate dosing can lead to a wide range of adverse effects. This section explores the potential risks associated with steroids, such as osteoporosis, glucose intolerance, and immunosuppression, prompting clinicians to carefully weigh the benefits against the risks in individual patients. Steroids represent a cornerstone of modern clinical practice, contributing significantly to the management of various medical conditions. This review provides an in-depth analysis of the diverse therapeutic applications of steroids, along with crucial safety considerations to guide clinicians in utilizing these potent agents judiciously. By understanding the intricacies of steroid therapy, healthcare providers can optimize patient outcomes while minimizing the likelihood of adverse events.

  • Saraswat A. Topical corticosteroid use in children: Adverse effects and how to minimize them . Indian J. Dermatol. Venereol. Leprol. 76, 225–228 (2010).

Indexed at , Google Scholar , Crossref

  • Beggs S. Paediatric analgesia . Aust Prescr. 31, 63–65 (2008).

Google Scholar , Crossref

  • Rossi M, Giorgi G. Domperidone and long QT syndrome . Curr Drug Saf. , 5, 257–262 (2010).
  • Kosek M, Bern C, Guerrant RL. The global burden of diarrhoeal disease, as estimated from studies published between 1992 and 2000 . Bull World Health Organ. 81, 197–204(2003).

Indexed at , Google Scholar

  • Alam N, Najam R. Effect of repeated oral therapeutic doses of methylphenidate on food intake and growth rate in rats . Pak J Pharm. Sci. 28 9–13(2015).

Google Scholar

  • Ryan C, Ross S, Davey P et al . Prevalence and causes of prescribing errors: The PRescribing Outcomes for Trainee Doctors Engaged in Clinical Training (PROTECT) study . PLoS ONE .9, 69-143 (2006).
  • Patrick DM, Marra F, Hutchinson J et al . Per capita antibiotic consumption: How does a North American jurisdiction compare with Europe? Clin Infect Dis . 39, 11-17 (2004).

Indexed at , Google Scholar , CrossRef

  • Li WC. Occurrence, sources, and fate of pharmaceuticals in aquatic environment and soil. Environ. Pollute. 187, 193-201 (2014).
  • Heberer T. Occurrence, fate, and removal of pharmaceutical residues in the aquatic environment:A review of recent research data. Toxicol Lett. 131, 5-17 (2002).
  • Banci L, Ciofi-Baffoni S, Tien M Lignin et al . Peroxidase-catalyzed oxidation of phenolic lignin oligomers . Biochemistry . 38, 3205-3210 (1999).

Annals of Clinical Trials and Vaccines Research

Google Scholar citation report

Citation image

research articles about steroids

Article Tags:

Ten per cent rise in steroids found in scottish prisons.

Ten per cent rise in steroids found in Scottish prisons

Dr Lorna Nisbet

A rise in steroid drugs found in Scottish prisons has been uncovered in a study led by the University of Dundee.

Researchers at the University’s Leverhulme Research Centre of Forensic Science (LRCFS), part of the School of Science and Engineering, made the discovery while working in collaboration with the Scottish Prison Service (SPS).

Their findings have been reported in a new publication, Changing trends in anabolic-androgenic steroid use within Scottish prisons , published by Wiley.

The researchers analysed 3,896 suspected drug samples, all of which had been seized at Scottish prisons between January 2019 and August 2023.

Of the samples which were analysed in 2019, less than one per cent were found to contain anabolic-androgenic steroids (AAS), compared to more than 10 per cent in 2023.

Various types of AAS were found – and they were the third most common drug detected in Scottish prisons in 2023.

Most of these steroid drugs (77 per cent) were in tablet form, of various vivid colours.

AAS compounds were also found in powders, herbal material, a fragmented soap bar-type sample, and within vapes. In many cases ASS was found along with other illicit substances.

Dr Lorna Nisbet, senior lecturer at LRCFS who was involved in the study, said: “The research shows a significant rise in steroid compounds within prisons, and they are being detected in forms we would not typically expect such as herbal material and vapes.

“Lots of these materials contain a combination of different drugs, in varying amounts, making it difficult for individuals to know exactly what they are taking or at what dosage.

“By consuming these drugs, individuals may be unknowingly engaging in polydrug use and the effects of these drugs when taken in combination with others can be particularly problematic.”

Polydrug use – taking more than one substance at a time – was prevalent in 81 per cent of all drug misuse fatalities in 2023, new data on drug-related deaths recently released by National Records of Scotland revealed.

“Polydrug use, can increase the toxic effects of drugs, prolong a drug’s effects on the person and increase negative side effects,” Dr Nisbet said.

“It is important that people working within prison settings know what drugs are potentially being used within the prison to allow them to provide the necessary support and address emerging risks more effectively.

“That is one of the reasons behind our collaborative work with SPS as it has allowed for the quick identification of changing drug use patterns and trends within the Scottish Prison landscape.”

She added: “Steroids are not routinely tested for in the United Kingdom but this data suggests that their use may be on the rise, and that more monitoring of these drugs may be needed.”

Share this article:

Subscribe to our newsletter to not miss articles like this one:.

research articles about steroids

Related Articles

Box image

Drugs offences accused fails in appeal against admissibility of police ‘expert’ evidence

Drugs offence accused fails in legal challenge against admissibility of evidence recovered following ‘forced entry’ into property.

Box image

Drugs gang leader forced to pay back profit

Box image

Drugs death report: Time to reform law and stop punishing addicts

Join more than 18,000 legal professionals in receiving our free daily email newsletter.

Scottish Legal News is your daily service for the latest news, jobs and events, delivered directly to your email inbox.

U.S. flag

An official website of the United States government

The .gov means it’s official. Federal government websites often end in .gov or .mil. Before sharing sensitive information, make sure you’re on a federal government site.

The site is secure. The https:// ensures that you are connecting to the official website and that any information you provide is encrypted and transmitted securely.

  • Publications
  • Account settings

Preview improvements coming to the PMC website in October 2024. Learn More or Try it out now .

  • Advanced Search
  • Journal List
  • J Sports Sci Med
  • v.5(2); 2006 Jun

Medical Issues Associated with Anabolic Steroid Use: Are They Exaggerated?

For the past 50 years anabolic steroids have been at the forefront of the controversy surrounding performance enhancing drugs. For almost half of this time no attempt was made by sports governing bodies to control its use, and only recently have all of the major sports governing bodies in North America agreed to ban from competition and punish athletes who test positive for anabolic steroids. These punitive measures were developed with the primary concern for promotion of fair play and eliminating potential health risks associated with androgenic-anabolic steroids. Yet, controversy exists whether these testing programs deter anabolic steroid use. Although the scope of this paper does not focus on the effectiveness of testing, or the issue of fair play, it is of interest to understand why many athletes underestimate the health risks associated from these drugs. What creates further curiosity is the seemingly well-publicized health hazards that the medical community has depicted concerning anabolic steroidabuse. Is there something that the athletes know, or are they simply naïve regarding the dangers? The focus of this review is to provide a brief history of anabolic steroid use in North America, the prevalence of its use in both athletic and recreational populations and its efficacy. Primary discussion will focus on health issues associated with anabolic steroid use with an examination of the contrasting views held between the medical community and the athletes that are using these ergogenic drugs. Existing data suggest that in certain circumstances the medical risk associated with anabolic steroid use may have been somewhat exaggerated, possibly to dissuade use in athletes.

  • For many years the scientific and medical communities depicted a lack of efficacy and serious adverse effects from anabolic steroid use.
  • Clinical case studies continue to link anabolic steroid administration with myocardial infarct, suicide, and cancer, evidence to support a cause and effect relationship is lacking.
  • It may be other contributing factors (i.e. genetic predisposition, diet, etc.) that play a substantial role and potentiate the harmful effects from anabolic steroids.

Introduction

Anabolic-androgenic steroids (herein referred to as only anabolic steroids) are the man-made derivatives of the male sex hormone testosterone. Physiologically, elevations in testosterone concentrations stimulate protein synthesis resulting in improvements in muscle size, body mass and strength (Bhasin et al., 1996 ; 2001 ). In addition, testosterone and its synthetic derivatives are responsible for the development and maturation of male secondary sexual characteristics (i.e. increase in body hair, masculine voice, development of male pattern baldness, libido, sperm production and aggressiveness).

Testosterone was isolated in the early 20 th century and its discovery led to studies demonstrating that this substance stimulated a strong positive nitrogen balance in castrated dogs and rats (Kochakian, 1950 ). Testosterone, because of its rapid degradation when given through either oral or parenteral administration, poses some limitations as an ergogenic aid. Although its potency is rapidly observed, the high frequency of administration needed becomes problematic. In addition, testosterone has a therapeutic index of 1 meaning there is similarity in the proportion between the anabolic and androgenic effects. As a result it becomes necessary to chemically modify testosterone to retard the degradation process and reduce some of the negative side effects. This allows for maintenance of effective blood concentrations for longer periods of time, may increase its interaction with the androgen receptor, and achieves the desired anabolic and androgenic changes.

Boje, 1939 was the first to suggest that exogenous testosterone administration may enhance athletic performance. By the late 1940’s and 1950’s testosterone compounds were experimented with by some west coast bodybuilders (Yesalis et al., 2000 ). The first dramatic reports of anabolic steroid use occurred following the 1954 world weightlifting championships (Yesalis et al., 2000 ). Use of these drugs spread quickly through the 1960’s and became popular among athletes in a variety of Olympic sports (Dubin, 1990 ). Wide spread use has also been reported in power lifters (Wagman et al., 1995 ), National Football League players (Yesalis et al., 2000 ), collegiate athletes (Yesalis, 1992 ), and recent claims of wide spread use in many sports including Major League Baseball players has made anabolic steroids the number one sports story of 2005 in some markets (Quinn, 2006 ). The ergogenic effects associated with anabolic steroids are presented in Table 1 .

Ergogenic effects associated with anabolic steroid use.

Athletes typically use anabolic steroids in a “stacking” regimen, in which they administer several different drugs simultaneously. The rationale for stacking is to increase the potency of each drug. That is, the potency of one anabolic agent may be enhanced when consumed simultaneously with another anabolic agent. They will use both oral and parenteral compounds. Most users will take anabolic steroids in a cyclic pattern, meaning the athletes will use the drugs for several weeks or months and alternate these cycles with periods of discontinued use. Often the athletes will administer the drugs in a pyramid (step-up) pattern in which dosages are steadily increased over several weeks. Towards the end of the cycle the athlete will ‘step-down’ to reduce the likelihood of negative side effects. At this point, some athletes will discontinue drug use or perhaps initiate another cycle of different drugs (i.e., drugs that may increase endogenous testosterone production to prevent the undesirable drop in testosterone concentrations that follows the removal of the pharmaceutical agents). A recent study has shown that the typical steroid regimen involved 3.1 agents, with a typical cycle ranging from 5 – 10 weeks (Perry et al., 2005 ). The dose that the athlete administers was reported to vary between 5 - 29 times greater than physiological replacement doses (Perry et al., 2005 ). These higher pharmacological dosages appear necessary to elicit the gains that these athletes desire. In a classic study on the dose-response curve of anabolic steroids, Forbes, 1985 demonstrated that the total dose of anabolic steroids have a logarithmic relationship to increases in lean body mass. These results exacerbate the athlete’s philosophy that if a low dose is effective, then more must be better.

Adverse effects associated with anabolic steroid use are listed in Table 2 . For years, the medical and scientific communities attempted to reduce anabolic steroid use by athletes by underscoring their efficacy and focusing on the unhealthy side effects (Biely, 1987 ; Darden, 1983 ; Fahey and Brown, 1973 ; Fowler et al., 1965 ; Golding et al., 1974 ). For the most part, this may have proved to be ineffective and caused athletes to lose trust in the physician’s knowledge of anabolic steroids thereby forcing them to seek advice from friends, internet sites or drug suppliers (Pope et al., 2004 ). However, recent literature has suggested that the medical issues associated with anabolic steroids may be somewhat overstated (Berning et al., 2004 ; Sturmi and Diorio, 1998 ; Street et al., 1996 ) considering that many of the side effects associated with anabolic steroid abuse are reversible upon cessation. It is important to note that there are differences in the side effects associated with anabolic steroid use (i.e.under medical supervision) versus abuse (i.e. consumption of many drugs at high doses).

Adverse effects associated with anabolic steroid use.


The clinical examination of anabolic steroid use is quite limited. Much of the problem in prospectively examining the effects of anabolic steroids on the athletic population is related to the unwillingness of institutional review boards to approve such studies in a non-clinical population. As a result, most of the investigations concerning medical issues associated with anabolic steroid administration have been performed on athletes self-administering the drugs. Anecdotally, it appears that a disproportionate magnitude of use and incidence of adverse effects are evident in bodybuilders (who are also known for consuming several other drugs that relieve some side effects but potentiate other risk factors as well, i.e. diuretics, thyroid hormones, insulin, anti-estrogens, etc.) compared to strength/power athletes. The mindset and motivation of these two types of athletes can be quite different. The strength/power athlete will typically use anabolic steroids to prepare themselves for a season of competition. They will generally cycle the drug to help them reach peak condition at a specific time of the training year. In contrast, bodybuilders use anabolic steroids to enhance muscle growth and definition. Their success is predicated on their aesthetic appearance. As a result many of these athletes may use anabolic steroids excessively for severalyears without cycling off or perhaps minimizing the length of “off cycles” depending on their competition schedule. Recent research has indicated that those athletes exhibit behavior that are consistent with substance dependence disorder (Perry et al., 2005 ). Although the medical issues associated with anabolic steroids may be quite different between these two types of athletes, the scientific literature generally does not differentiate between the two. The following sections will discuss adverse effects on specific physiological systems associated with anabolic-androgenic steroid use. It is important to note that many athletes consume multiple drugs in addition to anabolic steroids. Thus, the unhealthy side effects could be potentiated by the use of drugs such as human growth hormone or IGF-1.

Cardiovascular System

In both the medical and lay literature one of the principal adverse effects generally associated with anabolic steroid use is the increased risk for myocardial infarction. This is primarily based upon several case reports published over the past 20 years describing the occurrence of myocardial infarctions in young and middle-aged body builders or weight lifters attributed to anabolic steroid use and/or abuse (Bowman, 1989 ; Ferenchick and Adelman, 1992 ; Gunes et al., 2004 ; Kennedy and Lawrence, 1993 ; Luke et al., 1990 ; McNutt et al., 1988 ). However, direct evidence showing cause and effect between anabolic steroid administration and myocardial infarction is limited. Many of the case studies reported normal coronary arterial function in anabolic steroid users that experienced an infarct (Kennedy and Lawrence, 1993 ; Luke et al., 1990 ), while others have shown occluded arteries with thrombus formation (Ferenchick and Adelman, 1992 ; Gunes et al., 2004 ; McNutt et al., 1988 ). Still, some of these studies have reported abnormal lipoprotein concentrations with serum cholesterol levels nearly approaching 600 mg·dl -1 (McNutt et al., 1988 ). Interestingly, in most case studies the effects of diet or genetic predisposition for cardiovascular disease were not disseminated and could not be excluded as contributing factors.

Alterations in serum lipids, elevations in blood pressure and an increased risk of thrombosis are additional cardiovascular changes often associated with anabolic steroid use (Cohen et al., 1986 ; Costill et al., 1984 ; Dhar et al., 2005 ; Kuipers et al., 1991 ; LaRoche, 1990 ). The magnitude of these effects may differ depending upon the type, duration, and volume of anabolic steroids used. Interesting to note is that these effects appear to be reversible upon cessation of the drug (Dhar et al., 2005 , Parssinen and Seppala, 2002 ). In instances where the athlete remains on anabolic steroids for prolonged periods of time (e.g “abuse”), the risk for developing cardiovascular disease may increase. Sader and colleagues ( 2001 ) noted that despite low HDL levels in bodybuilders, anabolic steroid use did not appear to cause significant vascular dysfunction. Interestingly, athletes participating in power sports appear to have a higher incidence of cardiovascular dysfunction than other athletes, regardless of androgen use (Tikkanen et al., 1991 ; 1998 ). Thus, a strength/power athlete with underlying cardiovascular abnormalities that begins using anabolic steroids is at a much higher risk for cardiovascular disease. However, anabolic steroid-induced changes in lipid profiles may not, per se, lead to significant cardiovascular dysfunction.

The risk of sudden death from cardiovascular complications in the athlete consuming anabolic steroids can occur in the absence of atherosclerosis. Thrombus formation has been reported in several case studies of bodybuilders self-administering anabolic steroids (Ferenchick, 1991 ; Fineschi et al., 2001 ; McCarthy et al., 2000 ; Sahraian et al., 2004 ). Melchert and Welder, 1995 have suggested that the use of 17α-alkylated steroids (primarily from oral ingestion) likely present the highest risk for thrombus formation. They hypothesized that anabolic steroid consumption can elevate platelet aggregation, possibly through an increase in platelet production of thromboxane A 2 and/or decreasing platelet production of prostaglandin PgI 2, resulting in a hypercoagulable state.

Left ventricular function and anabolic steroid use/abuse has been examined. Climstein and colleagues ( 2003 ) demonstrated that highly strength-trained athletes, with no history of anabolic steroid use exhibited a higher incidence of wave form abnormalities relative to recreationally-trained or sedentary individuals. However, when these athletes self-administered anabolic steroids, a higher percentage of wave form abnormalities were exhibited. Further evidence suggestive of left ventricular dysfunction has been reported in rodent models. A study on rats has shown that 8 weeks of testosterone administration increased left ventricle stiffness and caused a reduction in stroke volume and cardiac performance (LeGros et al., 2000 ). It was hypothesized that the increased stiffness may have been related to formation of crosslinks between adjacent collagen molecules within the heart. Others have suggested that anabolic steroid use may suppress the increases normally shown in myocardial capillary density following prolonged endurance training (Tagarakis et al., 2000 ). However, there are a number of interpretational issues with this study. The changes reported were not statistically significant. In addition, the exercise stimulus employed (prolonged endurance training) is not the primary mode of exercise frequently used by anabolic steroid users. Resistance training, independent of anabolic steroid administration, has been shown to increase left ventricular wall and septal thickness due to the high magnitude of pressure overload (Fleck et al., 1993 ; Fleck, 2003 ; Hoffman, 2002 ). This is known as concentric hypertrophy and does not occur at the expense of left ventricular diameter. In general, cardiac hypertrophy (resulting from a pressure overload, i.e. hypertension) may not be accompanied by a proportional increase in capillary density (Tomanek, 1986 ). Therefore, the potential for a reduction in coronary vasculature density exists for the resistance- trained athlete. However, it does not appear to pose a significant cardiac risk for these athletes. Recent observations have shown a dose-dependent increase in left ventricular hypertrophy (LVH) in anabolic steroid users (Parssinen and Seppala, 2002 ). This may have the potential to exacerbate the reduction in coronary vasculature density. However, the authors have acknowledged that their results may have been potentiated by a concomitant use of human growth hormone by their subjects. Other studies have failed to show additive effects of anabolic steroid administration and LVH in resistance-trained athletes (Palatini et al., 1996 ; Dickerman et al., 1998 ).

Hepatic System

An elevated risk for liver tumors, damage, hepatocellular adenomas, and peliosis hepatitis are often associated with anabolic steroid use or abuse. This is likely due to the liver being the primary site of steroid clearance. In addition, hepatic cancers have been shown to generally occur with higher frequency in males compared to females (El-Serag, 2004 ). It is thought that high endogenous concentrations of testosterone and low estrogen concentrations increase the risk of hepatic carcinomas (Tanaka et al., 2000 ). However, this appears to be prevalent for men with pre-existing liver disease. In normal, healthy men the relationship between testosterone concentrations and liver cancer has not been firmly established. Additional reports of liver cancer and anabolic steroids have been reported in non- athletic populations being treated with testosterone for aplastic anemia (Nakao et al., 2000 ). In regards to liver cancer and disease in athletes consuming anabolic steroids, many concerns have been raised based primarily on several case studies that have documented liver disease in bodybuilders using anabolic steroids (Cabasso, 1994 ; Socas et al., 2005 ; Soe et al., 1992 ).

A few studies have recently questioned the risk to hepatic dysfunction from anabolic steroid use (Dickerman et al., 1999 ). A recent study examining the blood chemistry of bodybuilders self-administering anabolic steroids reported elevations in aspartate aminotransferase (AST), alanine aminotransferase (ALT) and creatine kinase (CK), but no change in the often-regarded more sensitive gamma- glutamyltranspeptidase (GGT) concentration (Dickerman et al., 1999 ). Thus, some experts have questioned these criteria tools because of the difficulty in dissociating the effects of muscle damage resulting from training from potential liver dysfunction. This has prompted some researchers to suggest that steroid-induced hepatotoxicity may be overstated. Another study involved a survey sent to physicians asking them to provide a diagnosis for a 28-year-old anabolic steroid using bodybuilder with abnormal serum chemistry profile (elevations in AST, ALT, CK, but with a normal GGT) (Pertusi et al., 2001 ). The majority of physicians (63%) indicated liver disease as the primary diagnosis as 56% of physicians failed to acknowledge the potential role of muscle damage or disease thereby increasing the likelihood of overemphasized anabolic steroid-induced hepatotoxicity diagnoses. Many case reports involving anabolic steroid administration and hepatic cancers examined individuals who were treated with oral steroids (17α-alkylated) for many years. No cysts or tumors have been reported in athletes using 17β-alkylated steroids. Thus, evidence appears to indicate that the risk for hepatic disease from anabolic steroid use may not be as high as the medical community had originally thought although a risk does exist especially with oral anabolic steroid use or abuse.

Bone and Connective Tissue

The issue of anabolic steroids and bone growth has been examined in both young and adult populations. In both populations, androgens have been successfully used as part of the treatment for growth delay (Albanese et al., 1994 ; Bagatell and Bremner, 1996 ; Doeker et al., 1998 ), and for osteoporosis in women (Geusens et al., 1986 ). Androgens are bi-phasic in that they stimulate endochondral bone formation and induce growth plate closure at the end of puberty. The actions of androgens on the growth plate are mediated to a large extent by aromatization to estrogens (Vanderschueren et al., 2004 ; Hoffman, 2002 ). Anabolic steroid use results in significant elevations in estrogens thought to impact premature closure of the growth plate. The acceleration of growth in adolescents treated with testosterone has raised concern for the premature closure of the epiphyseal plate (NIDA, 1996 ; Sturmi and Diorio, 1998 ). However, there does not appear to be any reports documenting the occurrence of premature stunted growth inadolescents taking anabolic steroids. Interesting, anabolic steroid administration in colts has been reported to delay epiphyseal plate closure (Koskinen and Katila, 1997 ). Although comparisons between humans and animals are difficult to make, suprapharmacological dosages that most athletes use may pose a greater risk than the doses studied to date. Thus, for the adolescent athlete using anabolic steroids the risk of premature epiphyseal plate closure may exist.

Anabolic steroids have been suggested to increase the risk of tendon tears in athletes (David et al., 1994 ; Stannard and Bucknell, 1993 ). Studies in mice have suggested that anabolic steroids may lead to degeneration of collagen (proportional to duration of steroid administration) and potentially lead to a decrease in tensile strength (Michna, 1986 ). In addition, a decrease in collagen synthesis has been reported from anabolic steroid administration in rats (Karpakka et al., 1992 ). The response in humans has been less clear. Mechanical failure has been suggested as a mechanism in anabolic steroid-using athletes. Skeletal muscle adaptations (i.e. hypertrophy and strength increases) take place rather rapidly in comparison to connective tissue. Therefore, tendon injuries in athletes are thought to occur from a rapid increase in training intensity and volume where connective tissue fails to withstand the overload. However, case reports of spontaneous tendon ruptures of weightlifters and athletes are limited.Although experimental data from animal models suggest that anabolic steroids may alter biomechanical properties of tendons, ultrastructural evidence supporting this claim is lacking. One study has shown that high doses of anabolic steroids decrease the degradation and increase the synthesis of type I collagen (Parssinen et al., 2000 ). Evans and colleagues ( 1998 ) performed an ultrastructural analysis on ruptured tendons from anabolic steroid users. They concluded that anabolic steroids did not induce any ultrastructural collagen changes that would increase the risk of tendon ruptures. Although the incidences of tendon rupture in anabolic steroid users should not be discounted, it is important to consider it in relation to the mechanical stress encountered from the rapid increases in muscular performance. Prospective research on anabolic steroid use and connective tissue injury is warranted.

Psychological and Behavioral

An issue that is often raised with anabolic steroid use is the psychological and behavioral effects. Increases in aggressiveness, arousal and irritability have been associated with anabolic steroid use. This has potentially beneficial and harmful implications. Elevations in arousal and self-esteem may be a positive side effect for the athlete. The increase in aggressiveness is a benefit that athletes participating in a contact sport may possess. However, increased aggressiveness may occur outside of the athletic arena thereby posing significant risks for anabolic steroid users and those they come in contact with. Anabolic steroids are associated with mood swings and increases in psychotic episodes. Studies have shown that nearly 60% of anabolic steroid users experience increases in irritability and aggressiveness (Pope and Katz, 1994 ; Silvester, 1995 ). A recent study by Pope and colleagues ( 2000 ) reported that significant elevations in aggressiveness and manic scores were observed following 12 weeks of testosterone cypionate injections in a controlled double-blind cross-over study. Interestingly, the results of this study were not uniform across the subjects. Most subjects showed little psychological effect and few developed prominent effects. A cause and effect relationship has yet to be identified in anabolic steroid users and it does appear that individuals who experience psychological or behavioral changes do recover when steroid use is discontinued (Fudula et al., 2003 ).

Additional Adverse Effects Associated with Anabolic Steroid Use

Other adverse events generally associated with anabolic steroid use include acne, male pattern baldness, gynecomastia, decreased sperm count, testicular atrophy, impotence, and transient infertility. Acne is one of the more common side effects associated with anabolic steroid administration. One study reported that 43% of users experienced acne as a consequence from androgen use (O’Sullivan et al., 2000 ). Few other investigations have been able to prospectively determine the occurrence of side effects associated with androgen administration. Increases in acne are thought to be related to a stimulation of sebaceous glands to produce more oil. The most common sites of acne development are on the face and back. Acne appears to disappear upon cessation of androgen administration.

Male pattern baldness does not appear to be a common adverse effect, but is often discussed as a potential side effect associated with androgen use. This is likely related to the role that androgens have in regulating hair growth (Lee et al., 2005 ). An abnormal expression of a specific cutaneous androgen receptor increases the likelihood of androgenic alopecia (Kaufman and Dawber, 1999 ; Lee et al., 2005 ). Thus, it is likely that androgenic alopecia observed as a result of exogenous androgen use is more prevalent in individuals that have a genetic predisposition to balding.

Gynecomastia is a common adverse effect associated with anabolic steroid use. Research has demonstrated a prevalence rate of 37% in anabolic steroid users (O’Sullivan et al., 2000 ). Gynecomastia isa benign enlargement of the male breast resulting from an altered estrogen-androgen balance, or increased breast sensitivity to a circulating estrogen level. Increases in estrogen production in men are seen primarily through the aromatization of circulating testosterone. Many anabolic steroid users will use anti-estrogens (selective estrogen receptor modulators) such as tamoxifen and clomiphene or anastrozole which is a nonsteroidal aromatase inhibitor to minimize side effects of estrogen and stimulate testosterone production. Once gynecomastia is diagnosed cosmetic surgery is often needed to correct the problem.

Changes in libido appear to be the most common adverse event (approximately 61% of users) reported in a small sample of anabolic steroid users (O’Sullivan et al., 2000 ). Although testosterone is often used in hypogonadal men to restore normal sexual function, increasing testosterone above the normal physiological range does not appear to increase sexual interest or frequency of sexual behavior in healthy men administered anabolic steroids in supraphysiological dosages (up to 500 mg·wk -1 ) for 14 weeks (Yates et al., 1999 ). Other studies confirm unchanged libido following 10 weeks of anabolic steroid administration in dosages ranging up to 200 mg·wk -1 (Schurmeyer, et al., 1984 ). However, reports do indicate that towards the end of an androgen cycle some men may experience loss of libido (O’Sullivan et al., 2000 ). It was thought that the decreased libido was related to the transient hypogonadism which typically occurs during exogenous androgen administration. Decreases in libido as a result of hypogonadism appear to be a function of high baseline levels of sexual functioning and desire (Schmidt et al., 2004 ). This may explain the conflicting reports seen in the literature. Regardless, changes in libido do appear to normalize once baseline endogenous testosterone concentrations return (Schmidt et al., 2004 ).

Another frequent adverse event relating to sexual function in males administering anabolic steroids is reversible azoospermia and oligospermia (Alen and Suominen, 1984 ; Schurmeyer et al., 1984 ). As exogenous androgen use increases, endogenous testosterone production is reduced. As a result, testicular size is reduced within three months of androgen administration (Alen and Suominen, 1984 ). In addition, sperm concentration and the number of spermatozoa in ejaculate may be reduced or eliminated by 7 weeks of administration (Schurmeyer et al., 1984 ). During this time risk for infertility is elevated. However, the changes seen in testicular volume, sperm count and concentration are reversible. Anabolic steroid-induced hypogonadism returns to baseline levels within 4 months following discontinuation of androgen use (Jarow and Lipshultz, 1990 ), and sperm counts and concentration return to normal during this time frame (Alen and Suominen, 1984 ; Schurmeyer et al., 1984 ).

Medical Issues Associated with Female Steroid Use

In female anabolic steroid users the medical issues are quite different than that shown in men. Deepening of the voice, enlargement of the clitoris, decreased breast size, altered menstruation, hirsutism and male pattern baldness are all clinical features common to hyperandrogenism in females (Derman, 1995 ). Androgen excess may occur as the result of polycystic ovary syndrome, congenital adrenal hyperplasia and possibly Cushing’s syndrome (Derman, 1995 ; Redmond, 1995 ). However, these clinical symptoms are seen in young, female athletes that are self-administering anabolic steroids. In contrast to men, many of these adverse events in the female anabolic steroid user may not be transient (Pavlatos et al., 2001 ).

Long Term Health Issues Associated with Anabolic Steroid Administration

The acute health issues associated with anabolic steroid use appear to be transient and more prevalent in individuals with genetic predisposition (e.g. hair loss, heart disease). It is the long-term effects that become a larger issue. However, limited data are available. In one study in mice, anabolic steroids were administered in relative dosages typically used by bodybuilders. However, the duration of the study was 1/5 the life span of the mouse which is relatively greater than that experienced by most athletes self-administering androgens. The results demonstrated a shortened life span of the mice with evidence of liver, kidney and heart pathology (Bronson and Matherne, 1997 ). In a study on Finnish power lifters, investigators examined 62 athletes who finished in the top 5 in various weight classes between the years 1977 and 1982 (Parssinen et al., 2000 ). These investigators reported that during a 12-year follow-up, the mortality rate for the power lifters was 12.1% compared to 3.1% in a control population. They concluded that their study depicted the detrimental long-term health effects from anabolic steroid use. Others have suggested that prolonged anabolic steroid use may increase the risk for premature death, but this may be more relevant in subjects with substance abuse or underlying psychiatric disease (Petersson et al., 2006 ).

The use of anabolic steroids in strength/power athletes has been reported for more than 50 years in North America. As discussed in the beginning of this review, during the 1970’s and 1980’s anecdotal reports on the rampant use of anabolic steroids in professional athletes were prevalent. However, little information is available concerning steroid-related diseases or associated deaths in these former strength/power athletes who are now well into middle age. Regardless, research should focus on these former athletes to ascertain possible long-term effects from androgen use.

Is There a Clinical Role of Androgenic Anabolic Steroids?

The efficacy of anabolic steroids in enhancing muscle strength and lean tissue accruement is no longer an issue for debate. While the issue of medical risks in individuals self-administering anabolic steroids is still being hotly debated, the medical community is no longer denying the potential clinical use of these androgens (Dobs, 1999 ). In recent years clinical treatment with anabolic steroids has increased lean tissue and improved daily functional performance in AIDS patients (Strawford et al., 1999 ) patients receiving dialysis (Johansen et al., 1999 ), patients with chronic obstructive pulmonary disease (Ferreira et al., 1998 ), and patients recovering from a myocardial infarction (Nahrendorf et al., 2003 ). In addition, research has demonstrated a positive effect on healing from muscle contusion injuries (Beiner et al., 1999 ). Although the medical community has generally taken a conservative approach to promoting anabolic steroids as part of a treatment plan in combating diseases involving muscle wasting, the body of knowledge that has developed indicates the potential positive effects of androgen therapy for certain diseased populations.

Conclusions

For many years the scientific and medical communities depicted a lack of efficacy and serious adverse effects from anabolic steroid use. However, competitive athletes continued to experiment with, use, and abuse anabolic steroids on a regular basis to enhance athletic performance despite the potential harmful side effects. The empirical evidence that the athletes viewed may have led to the development of distrust between the athletic and medical communities. Science has been lagging several years behind the experimental practices of athletes. In fact, most athletes consume anabolic steroids on a trial and error approach based on information gained from other athletes, coaches, websites, or gym “gurus.” Science has lacked in its approach to study anabolic steroids because only few studies have examined long-term cyclical patterns, high doses, and the effects of stacking different brands of steroids. These practices are common to the athletic community and not for the medicinal purposes of anabolic steroid therapy. In addition, some athletes (especially bodybuilders) have experimented with drugs unbeknown to the medical community, i.e. insulin, thyroid hormones, and site-specific enhancers such as Synthol and Esiclene to name a few.

When examining the potential medical issues associated with anabolic steroid use, evidence indicates that most known side effects are transient. More so, few studies have been able to directly link anabolic steroids to many of the serious adverse effects listed. Although clinical case studies continue to link anabolic steroid administration with myocardial infarct, suicide, and cancer, the evidence to support a cause and effect relationship is lacking and it may be other contributing factors (i.e. genetic predisposition, diet, etc.) play a substantial role and potentiate the harmful effects from anabolic steroids. Consistent physician monitoring is critical to the athlete who consumes anabolic steroids. However, many athletes may not undergo extensive medical exams prior to androgen administration and few physicians may be willing to provide such monitoring.

The purpose of this review was not to support or condone anabolic steroid use. Rather, the aim was to discuss pertinent medical issues and provide another perspective in light of the fact that many anabolic steroids users do not appear to prioritize the health/safety hazards or potential adverse medical events. In order to maintain credibility with the athlete, it is important to provide accurate information to the athlete in regards to these performance enhancing drugs, and provide education about alternative means and potential risks. Finally, anabolic steroids have been used legitimately for several clinical purposes such as muscle wasting or hypogonadal related diseases.

Biographies

Jay R. HOFFMAN

The College of New Jersey.

Research interests

Sport supplementation, resistance training, eExercise endocrinology.

E-mail: ude.jnct@jnamffoh

Nicholas A. RATAMESS

Sport supplementation, resistance training, exercise endocrinolgy.

E-mail: ude.jnct@ssematar

  • Albanesey A., Kewley G.D., Long A., Pearl K.N., Robins D.G., Stanhope R. (1994) Oral treatment for constitutional delay of growth and puberty in boys: a randomised trial of an anabolic steroid or testosterone undecanoate . Archives of Disease in Childhood 71 , 315-317 [ PMC free article ] [ PubMed ] [ Google Scholar ]
  • Alen M., Suominen J. (1984) Effect of androgenic and anabolic steroids on spermatogenesis in power athletes . International Journal of Sports Medicine 5 , 189-192 [ Google Scholar ]
  • Bagatell C.J., Bremner W.J. (1996) Androgens in men – uses and abuses . New England Journal of Medicine 334 , 707-714 [ PubMed ] [ Google Scholar ]
  • Beiner J.M., Jokl P., Cholewicki J., Panjabi M.M. (1999) The effect of anabolic steroids and corticosteroids on healing of muscle contusion injury . American Journal of Sports Medicine . 27 , 2-9 [ PubMed ] [ Google Scholar ]
  • Berning J.M., Adams K.J., Stamford B.A. (2004) Anabolic steroid usage in athletics: Facts, fiction, and public relations . Journal of Strength and Conditioning Research 18 , 908-917 [ PubMed ] [ Google Scholar ]
  • Bhasin S., Storer T.W., Berman N., Callegari C., Clevenger B., Phillips J., Bunnell T.J., Tricker R., Shirazi A., Casaburi R. (1996) The effects of supraphysiologic doses of testosterone on muscle size and strength in men . New England Journal of Medicine 335 , 1-7 [ PubMed ] [ Google Scholar ]
  • Bhasin S., Woodhouse L., Storer T.W. (2001) Proof of the effect of testosterone on skeletal muscle . Journal of Endocrinology 170 , 27-38 [ PubMed ] [ Google Scholar ]
  • Biely J.R. (1987) Use of anabolic steroids by athletes. Do the risks outweigh the benefits? Postgraduate Medicine 82 , 71-74 [ PubMed ] [ Google Scholar ]
  • Bowman S.J. (1989) Anabolic steroids and infarction . British Medical Journal 299 , 632 [ Google Scholar ]
  • Boje O. (1939) Doping . Bulletin of the Health Organization of the League of Nations 8 , 439-469 [ Google Scholar ]
  • Bronson F.H., Matherne C.M. (1997) Exposure to anabolic-androgenic steroids shortens life span of male mice . Medicine and Science in Sports and Exercise 29 , 615-619 [ PubMed ] [ Google Scholar ]
  • Cabasso A. (1994) Peliosis hepatis in a young adult bodybuilder . Medicine and Science in Sports and Exercise 26 , 2-4 [ PubMed ] [ Google Scholar ]
  • Climstein M., O’Shea P., Adams K.J., DiBeliso M. (2003) The effects of anabolic-androgenic steroids upon resting and peak exercise left ventricular heart wall motion kinetics in male strength and power athletes . Journal of Science and Medicine and Sport . 6 , 387-397 [ PubMed ] [ Google Scholar ]
  • Cohen J.C., Faber W.M., Spinnler Benade A.J., Noakes T.D. (1986) Altered serum lipoprotein profiles in male and female power lifters ingesting anabolic steroids . Physician and Sportsmedicine 14 , 131-136 [ PubMed ] [ Google Scholar ]
  • Costill D.L., Pearson D.R., Fink W.J. (1984) Anabolic steroid use among athletes: Changes in HDL-C levels . Physician and Sportsmedicine 12 , 113-117 [ Google Scholar ]
  • Darden E. (1983) The facts about anabolic steroids . Athletic Journal March, 100-101 [ Google Scholar ]
  • David H.G., Green J.T., Green A.J., Wilson C.A. (1994) Simultaneous bilateral quadriceps rupture: a complication of anabolic steroid use . Journal of Bone and Joint Surgery, British Volume 77 , 159-160 [ PubMed ] [ Google Scholar ]
  • Derman R.J. (1995) Effects of sex steroids on women’s health: Implications for practitioners . American Journal of Medicine . 98 ( Suppl ), 137S-143S [ PubMed ] [ Google Scholar ]
  • Dhar R., Stout C.W., Link M.S., Homoud M.K., Weinstock J., Estes N.A., III. (2005) Cardiovascular toxicities of performance-enhancing substances in sports . Mayo Clinic Proceedings 80 , 1308-1315 [ PubMed ] [ Google Scholar ]
  • Dickerman R.D., Schaller F., McConathy W.J. (1998) Left ventricular wall thickening does occur in elite power athletes with or without anabolic steroid use . Cardiology 90 , 145-148 [ PubMed ] [ Google Scholar ]
  • Dickerman R.D., Pertusi R.M., Zachariah N.Y., Dufour D.R., McConathy W.J. (1999) Anabolic steroid-induced hepatotoxicity: Is it overstated? Clinical Journal of Sport Medicine 9 , 34-39 [ PubMed ] [ Google Scholar ]
  • Dobs A.S. (1999) Is there a role for androgenic anabolic steroids in medical practice . Journal of American Medical Association 281 , 1326-1327 [ PubMed ] [ Google Scholar ]
  • Doeker B., Müller-Michaels J., Andler W. (1998) Induction of early puberty in a boy after treatment with oxandrolone? Hormone Research 50 , 46-48 [ PubMed ] [ Google Scholar ]
  • Dubin C. (1990) Commission of inquiry into the use of drugs and banned practices intended to increase athletic performance . (Catalogue No. CP32-56/1990E, ISBN 0-660-13610-4) . Ottawa, ON: Canadian Government Publishing Center [ Google Scholar ]
  • El-Serag H.B. (2004) Hepatocellular carcinoma: recent trends in the United States . Gastroenterology 127 , S27-S34 [ PubMed ] [ Google Scholar ]
  • Evans N.A., Bowrey D.J., Newman G.R. (1998) Ultrastructural analysis of ruptured tendon from anabolic steroid users . Injury 29 , 769-773 [ PubMed ] [ Google Scholar ]
  • Fahey T.D., Brown C.H. (1973) The effects of an anabolic steroid on the strength, body composition, and endurance of college males when accompanied by a weight training program . Medicine and Science in Sports 5 , 272-276 [ PubMed ] [ Google Scholar ]
  • Ferenchick G.S. (1991) Anabolic/androgenic steroid abuse and thrombosis: Is there a connection? Medical Hypothesis 35 , 27-31 [ PubMed ] [ Google Scholar ]
  • Ferenchick G.S., Adelman S. (1992) Myocardial infarction associated with anabolic steroid use in a previously healthy 37-year old weight lifter . American Heart Journal 124 , 507-508 [ PubMed ] [ Google Scholar ]
  • Ferreira I.M., Verreschi I.T., Nery L.E., Goldstein R.S., Zamel N., Brooks D., Jardim J.R. (1998) The influence of 6 months of oral anabolic steroids on body mass and respiratory muscle sin undernourished COPD patients . Chest 114 , 19-28 [ PubMed ] [ Google Scholar ]
  • Fineschi V., Baroldi G., Monciotti F., Paglicci Reattelli L., Turillazzi E. (2001) Anabolic steroid abuse and cardiac sudden death: A pathologic study . Archives of Pathology and Laboratory Medicine 125 , 253-255 [ PubMed ] [ Google Scholar ]
  • Fleck S.J., Pattany P.M., Stone M.H., Kraemer W.J., Thrush J., Wong K. (1993) Magnetic resonance imaging determination of left ventricular mass: junior Olympic weightlifters . Medicine and Science in Sports and Exercise 25 , 522-527 [ PubMed ] [ Google Scholar ]
  • Fleck S.J. (2003) Cardiovascular responses to strength training . In: Strength and Power in Sport . : Komi P.V.2nd edition Malden, MA: Blackwell Science; 387-406 [ Google Scholar ]
  • Forbes G.B. (1985) The effect of anabolic steroids on lean body mass: The dose response curve . Metabolism 34 , 571-573 [ PubMed ] [ Google Scholar ]
  • Fowler W.M., Jr., Gardner G.W., Egstrom G.H. (1965) Effect of an anabolic steroid on physical performance in young men . Journal of Applied Physiology 20 , 1038-1040 [ PubMed ] [ Google Scholar ]
  • Fudula P.J., Weinrieb R.M., Calarco J.S., Kampman K.M., Boardman C. (2003) An evaluation of anabolic-androgenic steroid abusers over a period of 1 year: seven case studies . Annals of Clinical Psychiatry 15 , 121-130 [ PubMed ] [ Google Scholar ]
  • Geusens P., Dequeker J., Verstraeten A., Nils J., Van Holsbeeck M. (1986) Bone mineral content, cortical thickness and fracture rate in osteoporotic women after withdrawal of treatment with nandrolone decanoate, 1-alpha hydroxyvitamin D3, or intermittent calcium infusions . Maturitas 8 , 281-289 [ PubMed ] [ Google Scholar ]
  • Golding L.A., Freydinger J.E., Fishel S.S. (1974) The effect of an androgenic-anabolic steroid and a protein supplement on size, strength, weight and body composition in athletes . Physician and Sportsmedicine 2 , 39-45 [ Google Scholar ]
  • Gunes Y., Erbas C, Okuyan E, Babalık E., Gurmen T. (2004) Myocardial infarction with intracoronary thrombus induced by anabolic steroids . Anadolu Kardiyoloji Dergisi 4 , 357-358 [ PubMed ] [ Google Scholar ]
  • Hoffman J.R. (2002) Physiological aspects of sport training and performance . Champaign, IL: Human Kinetics; 15-26 [ Google Scholar ]
  • Jarow J.P., Lipshultz L.I. (1990) Anabolic steroid-induced hypogonadotropic hypogonadism . American Journal of Sports Medicine 18 , 429-431 [ PubMed ] [ Google Scholar ]
  • Johansen K.L., Mulligan K., Schambelan M. (1999) Anabolic effects of nandrolone decanoate in patients receiving dialysis . Journal of American Medical Association 281 , 1275-1281 [ PubMed ] [ Google Scholar ]
  • Karpakka J.A., Pesola M.K., Takala T.E. (1992) The effects of anabolic steroids on collagen synthesis in rat skeletal muscle and tendon. A preliminary report . American Journal of Sports Medicine 20 , 262-266 [ PubMed ] [ Google Scholar ]
  • Kaufman K.D., Dawber R.P. (1999) Finasteride, a type 2 5alpha-reductase inhibitor, in the treatment of men with androgenic alopecia . Expert Opinion of Investigational Drugs 8 , 403-415 [ PubMed ] [ Google Scholar ]
  • Kennedy M.C., Lawrence C. (1993) Anabolic steroid abuse and cardiac death . Medical Journal of Australia 158 , 346-348 [ PubMed ] [ Google Scholar ]
  • Kochakian C.D. (1950) Comparison of protein anabolic property of various androgens in the castrated rat . American Journal of Physiology 160 , 53-61 [ PubMed ] [ Google Scholar ]
  • Koskinen E., Katila T. (1997) Effect of 19-norandrostenololylaurate on serum testosterone concentration, libido, and closure of distal radial growth plate in colts . Acta Veterinaria Scandinavica 38 , 59-67 [ PMC free article ] [ PubMed ] [ Google Scholar ]
  • Kuipers H., Wijnen J.A., Hartgens F., Willems S.M. (1991) Influence of anabolic steroids on body composition, blood pressure, lipid profile, and liver function in amateur bodybuilders . International Journal of Sports Medicine 12 , 413-418 [ PubMed ] [ Google Scholar ]
  • LaRoche G.P. (1990) Steroid anabolic drugs and arterial complications in an athlete – a case history . Angiology 41 , 964-969 [ PubMed ] [ Google Scholar ]
  • Lee P., Zhu C.C., Sadick N.S., Diwan A.H., Zhang P.S., Liu J.S., Prieto V.G. (2005) Expression of androgen receptor coactivator ARA70/ELE1 in androgenic alopecia . Journal of Cutaneous Pathology 32 , 567-571 [ PubMed ] [ Google Scholar ]
  • LeGros T., McConnell D., Murry T., Edavettal M., Racey-Burns L.A., Shepherd R.E., Burns A.H. (2000) The effects of 17 α-methyltestosterone on myocardial function in vitro . Medicine and Science in Sports and Exercise 32 , 897-903 [ PubMed ] [ Google Scholar ]
  • Luke J.L., Farb A., Virmani R., Barry Sample R.H. (1990) Sudden cardiac death during exercise in a weight lifter using anabolic androgenic steroids: pathological and toxicological findings . Journal of Forensic Science 35 , 1441-1447 [ PubMed ] [ Google Scholar ]
  • McCarthy K., Tang A.T., Dalrymple-Hay M.J., Haw M.P. (2000) Ventricular thrombosis and systemic embolism in bodybuilders: etiology and management . Annals of Thoracic Surgery 70 , 658-660 [ PubMed ] [ Google Scholar ]
  • McNutt R.A., Ferenchick G.S., Kirlin P.C., Hamlin N.J. (1988) Acute myocardial infarction in a 22-year-old world class weight lifter using anabolic steroids . American Journal of Cardiology 62 , 164. [ PubMed ] [ Google Scholar ]
  • Melchert R.B., Welder A.A. (1995) Cardiovascular effects of androgenic-anabolic steroids . Medicine and Science in Sports and Exercise 27 , 1252-1262 [ PubMed ] [ Google Scholar ]
  • Michna H. (1986) Organisation of collagen fibrils in tendon: changes induced by an anabolic steroid. I. Functional and ultrastructural studies . Virchows Archive B (Cell Pathology Including Molecular Pathology) 52 , 75-86 [ PubMed ] [ Google Scholar ]
  • Nahrendorf M., Frantz S., Hu K., von zur Muhlen C., Tomaszewski M., Scheuermann H, Kaiser R., Jazbutyte V., Beer S., Bauer W., Neubauer S., Ertl G., Allolio B., Callies F. (2003) Effect of testosterone on post-myocardial infarction remodeling and function . Cardiovascular Research 57 , 370-378 [ PubMed ] [ Google Scholar ]
  • Nakao A., Sakagami K., Nakata Y., Komazawa K., Amimoto T., Nakashima K., Isozaki H., Takakura N., Tanaka N. (2000) Multiple hepatic adenomas caused by long-term administration of androgenic steroids for aplastic anemia in association with familial adenomatous polyposis . Journal of Gastroenterology 35 , 557-562 [ PubMed ] [ Google Scholar ]
  • National Institute on Drug Abuse (1996) Anabolic steroids: A threat to mind and body . U.S. Department of Health and Human Services; NIH publication No. 96-3721 [ Google Scholar ]
  • O’Sullivan A.J., Kennedy M.C., Casey J.H., Day R.O., Corrigan B., Wodak A.D. (2000) Anabolic-androgenic steroids: Medical assessment of present, past and potential users . Medical Journal of Australia 173 , 323-327 [ PubMed ] [ Google Scholar ]
  • Palatini P., Giada F., Garavelli G., Sinisi F., Mario L., Michielleto M., Baldo-Enzi G. (1996) Cardiovascular effects of anabolic steroids in weight-trained subjects . Journal of Clinical Pharmacology 36 , 1132-1140 [ PubMed ] [ Google Scholar ]
  • Parssinen M., Karila T., Kovanen V., Seppala T. (2000) The effect of supraphysiological doses of anabolic androgenic steroids on collagen metabolism . International Journal of Sports Medicine 21 , 406-411 [ PubMed ] [ Google Scholar ]
  • Parssinen M., Seppala T. (2002) Steroid use and long-term health risks in former athletes . Sports Medicine 32 , 83-94 [ PubMed ] [ Google Scholar ]
  • Parssinen M., Kujala U., Vartiainen E., Sarna S., Seppala T. (2000) Increased premature mortality of competitive powerlifters suspected to have used anabolic agents . International Journal of Sports Medicine 21 , 225-227 [ PubMed ] [ Google Scholar ]
  • Pavlatos A.M., Fultz O., Monberg M.J., Vootkur A. (2001) Review of oxymtholone: A 17α-alkylated anabolic-androgenic steroid . Clinical Therapy 23 , 789-801 [ PubMed ] [ Google Scholar ]
  • Perry P.J., Lund B.C., Deninger M.J., Kutscher E.C., Schneider J. (2005) Anabolic steroid use in weightlifters and bodybuilders. An internet survey of drug utilization . Clinical Journal of Sports Medicine 15 , 326-330 [ PubMed ] [ Google Scholar ]
  • Pertusi R., Dickerman R.D., McConathy W.J. (2001) Evaluation of aminotransferases elevations in a bodybuilder using anabolic steroids: hepatitis or rhabdomyolysis . Journal of American Osteopathic Association 101 , 391-394 [ PubMed ] [ Google Scholar ]
  • Petersson A., Garle M., Granath F., Thiblin I. (2006) Morbidity and mortality in patients testing positively for the presence of anabolic androgenic steroids in connection with receiving medical care. A controlled retrospective cohort study . Drug and Alcohol Dependence 81 , 215-220 [ PubMed ] [ Google Scholar ]
  • Pope H.G., Katz D.L. (1994) Psychiatric and medical effects of anabolic-androgenic steroid use. A controlled study of 160 athletes . Archives of General Psychiatry 51 , 375-382 [ PubMed ] [ Google Scholar ]
  • Pope H.G., Kouri E.M., Hudson J.I. (2000) Effects of supraphysiologic doses of testosterone on mood and aggression in normal men . Archives of General Psychiatry 57 , 133-140 [ PubMed ] [ Google Scholar ]
  • Pope H.G., Kanayama G., Ionescu-Pioggia M., Hudson J.I. (2004) Anabolic steroid users’ attitudes towards physicians . Addiction 99 , 1189-1194 [ PubMed ] [ Google Scholar ]
  • Quinn T.J. (2006) Year of the Juice. Steroid scandal top story of 2005 . New York Daily News . January 3 [ Google Scholar ]
  • Redmond G.P. (1995) Androgenic disorders of women: Diagnostic and therapeutic decision making . American Journal of Medicine 98z ( Suppl ), 120S-129S [ PubMed ] [ Google Scholar ]
  • Sader M.A., Griffiths K.A., McCredie R.J., Handelsman D.J., Celermajer D.S. (2001) Androgenic anabolic steroids and arterial structure and function in male bodybuilders . Journal of American College of Cardiology 37 , 224-230 [ PubMed ] [ Google Scholar ]
  • Sahraian M.A., Mottamedi M., Azimi A.R., Moghimi B. (2004) Androgen-induced cerebral venous sinus thrombosis in a young body builder: case report . BMC Neurology 4 , 22. [ PMC free article ] [ PubMed ] [ Google Scholar ]
  • Schmidt P.J., Berlin K.L., Danacceau M.A., Neeren A., Haq N.A., Roca C.A., Rubinow D.R. (2004) The effects of pharmacologically induced hypogonadism on mood in healthy men . Archives of General Psychiatry 61 , 997-1004 [ PubMed ] [ Google Scholar ]
  • Schurmeyer T., Knuth U.A., Belkien L., Nieschlag E. (1984) Reversible azoospermia induced by the anabolic steroid 10-nortestosterone . Lancet 25 , 417-420 [ PubMed ] [ Google Scholar ]
  • Silvester L.J. (1995) Self-perceptions of the acute and long-range effects of anabolic-androgenic steroids . Journal of Strength and Conditioning Research 9 , 95-98 [ Google Scholar ]
  • Socas L., Zumbardo M., Perez-Luzardo O., Ramos A., Perez C., Hernandez J.R., Boada L.D. (2005) Hepatocellular adenomas associated with anabolic androgenic steroid abuse in bodybuilders: a report of two cases and a review of the literature . British Journal of Sports Medicine 39 , e27. [ PMC free article ] [ PubMed ] [ Google Scholar ]
  • Soe K.L., Soe M., Gluud C. (1992) Liver pathology associated with the use of anabolic-androgenic steroids . Liver 12 , 73-79 [ PubMed ] [ Google Scholar ]
  • Stannard J.P., Bucknell A.L. (1993) Rupture of the triceps tendon associated with steroid injections . American Journal of Sports Medicine 21 , 482-485 [ PubMed ] [ Google Scholar ]
  • Strawford A., Barbieri T, Van Loan M., Parks E., Catlin D, Barton N., Neese R., Christiansen M., King J., Hellerstein M.K. (1999) Resistance exercise and supraphysiologic androgen therapy in eugonadal men with HIV-related weight loss . Journal of American Medical Association 281 , 1282-1290 [ PubMed ] [ Google Scholar ]
  • Street C., Antonio J., Cudlipp D. (1996) Androgen use by athletes: a re-evaluation of the health risks . Canadian Journal of Applied Physiology 21 , 421-440 [ PubMed ] [ Google Scholar ]
  • Sturmi J.E., Diorio D.J. (1998) Anabolic agents . Clinics in Sports Medicine 17 , 261-282 [ PubMed ] [ Google Scholar ]
  • Sullivan M.L., Martinez C.M., Gennis P., Gallagher E.J. (1998) The cardiac toxicity of anabolic steroids . Progress in Cardiovascular Diseases . 41 , 1-15 [ PubMed ] [ Google Scholar ]
  • Tagarakis C.V.M., Bloch W., Hartmann G., Hollmann W., Addicks K. (2000) Testosterone-propionate impairs the response of the cardiac capillary bed to exercise . Medicine and Science in Sports and Exercise 32 , 946-953 [ PubMed ] [ Google Scholar ]
  • Tanaka K., Sakai H., Hashizume M., Hirohata T. (2000) Serum testosterone; estradiol ratio and the development of hepatocellular carcinoma among male cirrhotic patients . Cancer Research 60 , 5106-5110 [ PubMed ] [ Google Scholar ]
  • Tikkanen H.O., Hamalainen E., Sarna S., Adlercreutz H., Harkonen M. (1998) Associations between skeletal muscle properties, physical fitness, physical activity and coronary heart disease risk factors in men . Atherosclerosis 137 , 377-389 [ PubMed ] [ Google Scholar ]
  • Tikkanen H.O., Harkonen M., Naveri H. (1991) Relationship of skeletal muscle fiber type to serum high density lipoprotein cholesterol and apolipoprotein A-I levels . Atherosclerosis 90 , 48-57 [ PubMed ] [ Google Scholar ]
  • Tomanek R.J. (1986) Capillary and pre-capillary coronary vascular growth during left ventricular hypertrophy . Canadian Journal of Cardiology 2 , 114-119 [ PubMed ] [ Google Scholar ]
  • Vanderschueren D., Vandenput L., Boonen S., Lindberg M.K., Bouillon R., Ohlsson C. (2004) Androgens and bone . Endocrine Reviews . 25 : 389-425 [ PubMed ] [ Google Scholar ]
  • Wagman D.F., Curry L.A., Cook D.L. (1995) An investigation into anabolic androgenic steroid use by elite U.S. powerlifters . Journal of Strength and Conditioning Research 9 , 149-153 [ Google Scholar ]
  • Yates W.R., Perry P.J., MacIndoe J., Holman T., Ellingrod V. (1999) Psychosexual effects of three doses of testosterone cycling in normal men . Biology Psychiatry 45 , 254-260 [ PubMed ] [ Google Scholar ]
  • Yesalis C.E. (1992) Epidemiology and patterns of anabolic-androgenic steroid use . Psychiatric Annals 22 , 7-18 [ Google Scholar ]
  • Yesalis C.E., Courson S.P., Wright J. (2000) History of anabolic steroid use in sport and exercise . In: Anabolic steroids in sport and exercise . : Yesalis C.E.2nd editionChampaign, IL: Human Kinetics; 51-71 [ Google Scholar ]

IMAGES

  1. Research on Steroids

    research articles about steroids

  2. How Do Steroids Work? Exploring the Effects and Risks of Steroid Use

    research articles about steroids

  3. Britain's anabolic steroid epidemic

    research articles about steroids

  4. (PDF) Impact of steroids on long term use

    research articles about steroids

  5. Steroids in the acutely ill: Evolving recommendations and practice

    research articles about steroids

  6. What Does Science Say About the Risks of Taking Steroids?

    research articles about steroids

VIDEO

  1. What are the best steroids? #motivation #menshealth #fit #gym #men #masculinity #steroiduse

  2. Why I Took Steroids #MensHealth #SteroidUse #FitnessJourney #HealthResearch #Biohacking #menshealth

  3. STEROIDS ARE AWESOME (NEVER Take them)

  4. ANCA-associated GN: the Known and the Unknowns with Dr Andreas Kronbichler

  5. Combination of steroids with ponatinib as frontline therapy for acute lymphoblastic leukaemia

  6. BEFORE USING STEROIDS DO YOUR OWN RESEARCH! #PEDS #steroidssideeffects #bodybuilding

COMMENTS

  1. Steroids: Pharmacology, Complications, and Practice Delivery Issues

    Steroids: Pharmacology, Complications, and Practice ...

  2. Anabolic-androgenic steroids: How do they work and what are the risks

    Anabolic-androgenic steroids (AAS) are a class of natural and synthetic hormones that owe their name to their chemical structure (the steroid nucleus, see Figure 1 ) and the biological effects (anabolic and androgenic) they induce. Anabolic refers to the skeletal muscle-building properties of AAS, whereas androgenic refers to the induction ...

  3. Anabolic Steroids and Cardiovascular Outcomes: The Controversy

    Anabolic steroids (AS) are synthetic derivatives of the male sex hormone testosterone. The use of AS is not limited to bodybuilders and athletes, but non-athletes also use them. It is used to enhance athletic performance, induce muscle hypertrophy, and augment male sexual characteristics. AS use is associated with a wide range of side effects ...

  4. Steroids

    STEROIDS is an international research journal devoted to studies on all chemical and biological aspects of steroidal moieties. The journal focuses on both experimental and theoretical studies on the biology, chemistry, biosynthesis, metabolism, molecular biology, physiology and pharmacology of …. View full aims & scope.

  5. Anabolic Steroids and Other Appearance and Performance Enhancing Drugs

    Anabolic Steroids and Other Appearance and Performance ...

  6. Anabolic-Androgenic Steroid Use in Sports, Health, and Society ...

    SYNOPSIS This consensus statement is an update of the previous position stand from the American College of Sports Medicine (ACSM), published in 1987 ().Since then, a substantial amount of scientific data on anabolic-androgenic steroids (AAS) has emerged and the circumstances of AAS use has evolved in the athletic, recreational, and clinical communities.

  7. The science of steroids

    Steroids are complex lipophilic molecules that have many actions in the body to regulate cellular, tissue and organ functions across the life-span. Steroid hormones such as cortisol, aldosterone, estradiol and testosterone are synthesised from cholesterol in specialised endocrine cells in the adrenal gland, ovary and testis, and released into ...

  8. Anabolic-androgenic steroid use is associated with ...

    Previous research has uncovered medical and psychological effects of anabolic-androgenic steroid (AAS) use, but the specific relationship between AAS use and risk-taking behaviors as well as ...

  9. The Effects of Supraphysiologic Doses of Testosterone on Muscle Size

    Anabolic-androgenic steroids are widely abused by athletes and recreational bodybuilders because of the perception that these substances increase muscle mass and strength, 1-9 but this premise ...

  10. Mortality Among Users of Anabolic Steroids

    This study included 1189 males sanctioned for doping with AAS and 59 450 male control participants. The mean age among AAS users and control participants was 27.4 (SD, 6.9) years. The cohort had a mean follow-up time of 11 years (Table). In the control cohort 5.47% were right censored due to migration, compared with 2.10% for AAS users.

  11. Treatments for people who use anabolic androgenic steroids: a scoping

    Background A growing body of evidence suggests that anabolic androgenic steroids (AAS) are used globally by a diverse population with varying motivations. Evidence has increased greatly in recent years to support understanding of this form of substance use and the associated health harms, but there remains little evidence regarding interventions to support cessation and treat the consequences ...

  12. Adverse Effects of Anabolic-Androgenic Steroids: A Literature Review

    Anabolic-androgenic steroids (AASs), commonly known as anabolic steroids, are a large group of molecules including endogenously produced androgens, such as testosterone, as well as synthetically manufactured derivatives [1]. Testosterone, Nandrolone Decanoate (ND), methandienone, and methenolol, are the most commonly abused androgens [2].

  13. Exploring the prevalence of anabolic steroid use among men and women

    Exploring the prevalence of anabolic steroid use among men ...

  14. Recent Advances in Steroid Research and Nervous System Function

    Steroid research—with the discovery of new steroidogenic sites and innovative mechanisms of steroid action in recent decades—has focused on neurosteroids and neuroactive steroids, encompassing molecular to behavioral domains, with an exponential emphasis on translational and clinical research. ... and will be listed together on the special ...

  15. (PDF) Impact of steroids on long term use

    osteoporosis, proximal myopathy, thinning of th e skin, infection, and delayed wound healing are among side. effects of long-term use. Corticosteroids can cause. depression, agitation, and ...

  16. PDF The Role of Steroids in Clinical Practice: Benefits, Risks, and

    Steroids, also known as corticosteroids or glucocorticoids, are synthetic drugs that mimic the actions of naturally occurring hormones in the body. In clinical practice, steroids have a wide range of therapeutic applications due to their potent anti-inflammatory, immunosuppressive, and metabolic effects. This article provides a comprehensive ...

  17. Steroid use and human performance: Lessons for integrative biologists

    The topic of steroid effects on human athletic performance is germane to an emerging field of research investigating hormonal effects on animals' performance (e.g., sprint speed, endurance capacity, bite-force capacity) (Husak et al. 2009a), as testosterone may exert general effects on performance across widely divergent vertebrate taxa. Our ...

  18. Adverse Effects of Anabolic-Androgenic Steroids: A Literature Review

    Anabolic-androgenic steroids (AASs) are a large group of molecules including endogenously produced androgens, such as testosterone, as well as synthetically manufactured derivatives. AAS use is widespread due to their ability to improve muscle growth for aesthetic purposes and athletes' performance, minimizing androgenic effects. AAS use is very popular and 1-3% of US inhabitants have been ...

  19. The Uses of Anabolic Androgenic Steroids Among Athletes; Its Positive

    Some intramuscularly used steroids are boldenone undecenoate, methenolone acetate, nandrolone decanoate, Sustanon 250 ®, testosterone enanthate, and testosterone cypionate. 46 Anabolic-androgenic steroids are the best-studied class of PEDs that can boost a user's confidence and strength, leading users to overlook the severe, long-lasting ...

  20. Steroid Side Effects

    Steroid Side Effects | Clinical Pharmacy and Pharmacology

  21. Systemic steroids and bronchopulmonary dysplasia: a systematic review

    It is unclear if systemic steroids decrease the risk of Bronchopulmonary Dysplasia (BPD) while increasing the risk of neurodevelopmental impairment (NDI). Conduct a systematic review of randomized ...

  22. Steroids found in Scottish prisons increase from 1% to 10% in four years

    Percentage of samples seized from the Scottish prisons each year found positive for one or more anabolic-androgenic steroids (AASs). Credit: Drug Testing and Analysis (2024). DOI: 10.1002/dta.3790

  23. The Role of Steroids in Clinical Practice: Benefits, Risks, and

    Steroids remain indispensable in clinical practice, playing a vital role in managing a wide range of medical conditions [1, 2]. Their potent anti-inflammatory and immunosuppressive effects make them valuable tools for alleviating symptoms and improving patients' quality of life. However, healthcare professionals must carefully consider the ...

  24. Steroid Hormones and Receptors in Health and Disease

    Steroid hormone effects have been reported for almost 140 years 1 - 3. Research over the past 50 years has led to the discovery of steroid hormone receptors that act via both genomic and nongenomic ("rapid") mechanisms. Steroids are involved in physiology and disease, mediating endocrine, cardiovascular, and reproductive functions and ...

  25. Full article: Radiofrequency vs Steroid Injections for Spinal Facet and

    Radiofrequency vs Steroid Injections for Spinal Facet and Sacroiliac Joint Pain: A Systematic Review and Meta-Analysis ... across various databases yielded a total of 117 studies. After removing duplicate entries, we were left with 46 unique articles. These full texts were then meticulously screened for eligibility based on our inclusion and ...

  26. Ten per cent rise in steroids found in Scottish prisons

    A rise in steroid drugs found in Scottish prisons has been uncovered in a study led by the University of Dundee. Researchers at the University&rsquo;s Leverhulme Research Centre of Forensic Science (LRCFS), part of the School of Science and Engineering, made the discovery while working in collaborat

  27. Medical Issues Associated with Anabolic Steroid Use: Are They

    Gynecomastia is a common adverse effect associated with anabolic steroid use. Research has demonstrated a prevalence rate of 37% in anabolic steroid users (O'Sullivan et al., 2000). Gynecomastia isa benign enlargement of the male breast resulting from an altered estrogen-androgen balance, or increased breast sensitivity to a circulating ...