WikiJournal of Medicine/Alternative androgen pathways

From Wikiversity
Jump to navigation Jump to search

WikiJournal of Medicine
Open access • Publication charge free • Public peer review • Wikipedia-integrated

WikiJournal of Medicine is an open-access, free-to-publish, Wikipedia-integrated academic journal for Medical and Biomedical topics. <seo title=" WJM, WikiJMed, Wiki.J.Med., WikiJMed, Wikiversity Journal of Medicine, WikiJournal Medicine, Wikipedia Medicine, Wikipedia medical journal, WikiMed, Wikimedicine, Wikimedical, Medicine, Biomedicine, Free to publish, Open access, Open-access, Non-profit, online journal, Public peer review "/>

<meta name='citation_doi' value='10.15347/WJM/2023.003'>

Article information

Authors: Maxim G Masiutin[i] , Maneesh K Yadav  

See author information ▼
  1. maxim@masiutin.com

Abstract

Steroidogenic routes to androgens have been discovered and characterized over the last two decades that fall outside the Δ4 and Δ5 "classical androgen pathways" to testosterone and 5α-dihydrotestosterone. There has been considerable investigation into these routes that has come with natural inconsistencies and overlap in naming that can make it difficult to discover information about them as might be needed in a clinical context. This expository review uses "alternative androgen pathways" to include what has been called the "backdoor" pathway to 5α-dihydrotestosterone, the 5α-dione pathway and pathways to 11-oxygenated steroids. A brief history of what led to the discovery of these pathways, basic information about the steroids and proteins involved in their biosynthesis as well as a summary of clinically significant findings is provided. PubChem CIDs for all steroids have been compiled to help authors avoid naming errors in their work. Modest suggestions for future work in these pathways are also given at the end. Patient comprehension and the clinical diagnosis of relevant conditions such as hyperandrogenism can be impaired by the lack of clear and consistent knowledge of alternative androgen pathways; the authors hope this review will accurately disseminate such knowledge to facilitate the beneficial treatment of such patients.


Introduction

The classical androgen pathway is composed of the steroidogenic adrenal and gonadal metabolic pathways that transform cholesterol to the androgen testosterone (T), which is then transformed into the potent androgen 5α-dihydrotestosterone (DHT). Broadly, androgens are understood to exert their primary effects through binding to cytosolic androgen receptor (AR) that is translocated to the nucleus upon androgen binding and ultimately results in the transcriptional regulation of a number of genes via Androgen Responsive Elements.[1] This androgen response mechanism is perhaps best known and characterized in the context of male sexual differentiation and puberty, but plays a role in a variety of tissue types and processes.[2]

In 2003, a "backdoor" androgen pathway to DHT that did not proceed through T was discovered in the tammar wallaby.[3] Shortly after this study, the pathway was further characterized, and its potential clinical relevance in conditions involving androgen biosynthesis in humans was proposed.[4] In the years following, other distinct pathways to potent 11-oxygenated androgens have been characterized and also been described as "backdoor".[5]

The relatively recent characterization of these "alternative androgen pathways" in the literature can confound the search for clinical information in cases where androgen steroidogenesis is relevant. In addition to the reuse of the term "backdoor", studies across different androgen pathways have also used different names for the same molecules. While such naming inconsistencies are notoriously common when it comes to biomolecules,[6] instances of incorrect names are easy to find and are almost certainly attributable to complex naming rules for organic molecules. In addition, pathway definitions can sometimes differ in the precise initial/terminal molecules, and the inclusion/exclusion of such points can hinder queries in electronic pathway databases.

Alternative androgen pathways are now known to be responsible for the production of biologically active androgens in humans, and there is growing evidence that they play a role in clinical conditions associated with hyperandrogenism. Understanding androgen steroidogenesis at the level of detail presented in this paper and establishing consensus names and pathway specifications would facilitate access to information towards diagnosis and patient comprehension.

History

Backdoor Pathways to 5α-Dihydrotestosterone

In 1987, Eckstein et al.[7] incubated rat testicular microsomes in the presence of radiolabelled steroids and demonstrated that 5α-androstane-3α,17β-diol is preferentially produced from 17α-hydroxyprogesterone (17OHP). While "androstanediol" was used to denote both 5α-androstane-3α,17β-diol and 5α-androstane-3β,17β-diol, "3α-diol" is used here to abbreviate 5α-androstane-3α,17β-diol in this paper as it is a common convention and emphasizes it as the 3α-reduced derivative of DHT. The function of 3α-diol was not known at that time.

Tammar wallaby pouch young do not show sexually dimorphic circulating levels of T and DHT during prostate development which suggested that another androgenization mechanism was responsible. In 2000, Shaw et al.[8] demonstrated that circulating 3α-diol mediates prostate development in these pouch young via conversion to DHT in target tissues. While 3α-diol's AR binding affinity is five orders of magnitude lower than DHT (generally described as AR inactive), it was known 3α-diol can be oxidized back to DHT via the action of a number of dehydrogenases.[9]

In 2003, Wilson et al.[3] incubated the testes of tammar wallaby pouch young with radiolabelled progesterone to show that 5α-reductase expression in this tissue enabled a novel pathway from 17OHP to 3α-diol without T as an intermediate. The sequence of this transformation along steroids is shown below (the full names can be found in § Abbreviations and Identifiers as well as #Figure 2):

  • 17OHP → 17OHDHP → 5α-Pdiol → AST → 3α-diol

In 2004, Mahendroo et al.[10] demonstrated that an overlapping novel pathway is operating in mouse testes, generalizing what had been demonstrated in tammar wallaby:

  • P4 → 5α-DHP → AlloP5→ 5α-Pdiol → AST → 3α-diol

The term "backdoor pathway" was coined by Auchus in 2004[4] and defined as a route to DHT that: 1) bypasses conventional intermediates androstenedione (A4) and T; 2) involves 5α-reduction of 21-carbon (C21) pregnanes to 19-carbon (C19) androstanes; and 3) involves the 3α-oxidation of 3α-diol to DHT. The backdoor pathway explained how androgens are produced under certain normal and pathological conditions in humans when the classical androgen pathway cannot fully explain the observed consequences. The pathway Auchus defined adds DHT to the terminus of the pathway described by Wilson et al. in 2003:

  • 17OHP → 17OHDHP → 5α-Pdiol → AST → 3α-diol → DHT

The clinical relevance of these results was demonstrated in 2012 for the first time when Kamrath et al.[11] attributed the urinary metabolites to the androgen backdoor pathway from 17OHP to DHT in patients with steroid 21-hydroxylase (encoded by the gene CYP21A2) enzyme deficiency.

5α-Dione Pathway

In 2011, Chang et al.[12] demonstrated that an alternative pathway to DHT was dominant and possibly essential in castration-resistant prostate cancer (CRPC) by presenting evidence from cell culture and xenograft models:

  • A4 → 5α-dione → DHT

While this pathway was described as the "5α-dione pathway" in a 2012 review,[13] the existence of such a pathway in the prostate was hypothesized in a 2008 review by Luu-The et al.[14]

A modern outlook of the synthesis of the backdoor pathways to DHT and the 5α-dione pathway is shown in #Figure 2.

11-Oxygenated Androgen Pathways

11-Oxygenated androgens are the products of a distinct alternative androgen pathway. The 11-oxygenated C19 steroids 11β-hydroxyandrostenedione (11OHA4) and 11-ketoandrostenedione (11KA4) were known since the 1950s to be products of the human adrenal but were understood as androgen inactive in humans. Their role as substrates to potent androgens had been overlooked in humans until recently, though they were long known in teleost fishes.[15]

Rege et al. in 2013[16] measured 11-oxygenated androgen levels in healthy women and showed that both 11-ketodihydrotestosterone (11KT) and 11β-hydroxytestosterone (11OHT) could activate the AR.

In 2013, Storbeck et al.[17] demonstrated the existence of 11-oxygenated androgen pathways in androgen-dependent prostate cancer cell culture. The authors demonstrated that A4 transformed to 11OHA4 in the adrenal can ultimately be converted into 11KT and 11-ketodihydrotestosterone (11KDHT) as shown in #Figure 3. The authors found that 11KT activity is comparable to that of T, and 11-ketodihydrotestosterone (11KDHT) activity is comparable to that of DHT, while the activities of 11OHT and 5α-dihydro-11β-hydroxytestosterone (11OHDHT) were observed to be about half of T and DHT, respectively. However, androgen activity in that study was only assessed at a single concentration of 1 nM.[17] Full dose responses for 11KT and 11KDHT were characterized in a study by Pretorius et al. in 2016[18] that showed 11KT and 11KDHT both bind and activate the AR with affinities, potencies, and efficacies that are similar to that of T and DHT, respectively.

Barnard et al.[5] in 2017 demonstrated metabolic pathways from C21 steroids to 11KDHT that bypasses A4 and T in vitro in a prostate cancer derived cell line, an aspect that is similar to that of the backdoor pathway to DHT. These newly discovered pathways to 11-oxygenated androgens were also described as "backdoor" pathways due to this similarity, and were further characterized in subsequent studies.[19][20]

A diagram of the 11-oxygenated androgen pathways is shown in #Figure 3.

Definition

The term "alternative androgen pathway" is used in this paper to refer to any pathway that produces potent androgens without a T intermediate. This subsumes all three groups of androgen pathways described in the previous section. A new term that describes the three groups of pathways (as well as future discoveries) will allow a single entry point into scientific information when alternatives to the classical androgen pathway[21][22] must be considered.

Nomenclature and Background

The complex naming rules of organic chemistry have resulted in the use of incorrect steroid names in published studies. Many examples of errors in systematic names can be found via Google Scholar[23][24] for molecules described in this paper, motivating this expository section on steroid nomenclature to facilitate the use of correct names.

Almost all biologically relevant steroids can be presented as a derivative of a parent hydrocarbon structure that serves as a skeleton.[25] These parent structures have specific names, such as pregnane, androstane, etc. The derivatives carry various functional groups called suffixes or prefixes after the respective numbers indicating their position in the steroid nucleus.[26] The widely-used trivial steroid names such as progesterone, testosterone or cortisol can also be used as base names to derive new names, however, by adding prefixes only rather than suffixes e.g., the steroid 17α-hydroxyprogesterone has a hydroxy group (-OH) at position 17 of the steroid nucleus comparing to progesterone. The letters α and β[27] denote absolute stereochemistry at chiral centers (a specific nomenclature distinct from the R/S convention[28] of organic chemistry). In steroids drawn from the standard perspective used in this paper, α-bonds are depicted on figures as dashed wedges and β-bonds as solid wedges.

 Numbering of carbon atoms up to position 21 (positions 18 and 19 are omitted) in a hypothetical steroid nucleus, as defined by the Nomenclature of Steroids.


The molecule "11-deoxycortisol" is an example of a derived name that uses cortisol as a parent structure without an oxygen atom (hence "deoxy") attached to position 11 (as a part of a hydroxy group).[29] The numbering of positions of carbon atoms in the steroid nucleus is set in a template found in the Nomenclature of Steroids[30] that is used regardless of whether an atom is present in the steroid in question. Although the nomenclature defines more than 30 positions, this review only needs numbering for the first 21 (#Figure 1).

Unsaturated carbons (generally, ones that are a part of a double bond) in the steroid nucleus are indicated by changing -ane to -ene.[31] This change was traditionally done in the parent name, adding a prefix to denote the position, with or without Δ (Greek capital delta) which designates unsaturation, for example, 4-pregnene-11β,17α-diol-3,20-dione (also Δ4-pregnene-11β,17α-diol-3,20-dione) or 4-androstene-3,11,17-trione (also Δ4-androstene-3,11,17-trione). However, the Nomenclature of Steroids recommends the locant of a double bond to be always adjacent to the syllable designating the unsaturation, therefore, having it as a suffix rather than a prefix, and without the use of the Δ character, i.e. pregn-4-ene-11β,17α-diol-3,20-dione or androst-4-ene-3,11,17-trione. The double bond is designated by the lower-numbered carbon atom, i.e. "Δ4-" or "4-ene" means the double bond between positions 4 and 5. The saturation of carbons of a parent steroid can be done by adding "dihydro-" prefix,[32] i.e. saturation of carbons 4 and 5 of testosterone with two hydrogens is 4,5α-dihydrotestosterone or 4,5β-dihydrotestosterone. Generally, when there is no ambiguity, one number of a hydrogen position from a steroid with a saturated bond may be omitted, leaving only the position of the second hydrogen atom, e.g., 5α-dihydrotestosterone or 5β-dihydrotestosterone. The Δ5-steroids are those with a double bond between carbons 5 and 6 (#Figure 1) and the Δ4 steroids are those with a double bond between carbons 4 and 5.[33][31] The classical androgen pathway is generally described in terms of a sequence of Δ5 compounds:

  • P5 → 17OHP5 → DHEA→ A5

and Δ4 compounds:

  • P4 → 17OHP → AlloP5→ A4 → T

The abbreviations like "P4" and "A4" refer to Δ4-steroids, while "P5" and "A5" refer to Δ5-steroids.

The suffix -ol denotes a hydroxy group, while the suffix -one denotes an oxo group. When two or three identical groups are attached to the base structure at different positions, the suffix is ​​indicated as -diol or -triol for hydroxy, and -dione or -trione for oxo groups, respectively. For example, 5α-pregnane-3α,17α-diol-20-one has a hydrogen atom at the 5α position (hence the "5α-" prefix), two hydroxy groups (-OH) at the 3α and 17α positions (hence "3α,17α-diol" suffix) and an oxo group (=O) at the position 20 (hence the "20-one" suffix). However, erroneous use of suffixes can be found, e.g., "5α-pregnan-17α-diol-3,11,20-trione"[23] [sic] — since it has just one hydroxy group (at 17α) rather than two, then the suffix should be -ol, rather than -diol, so that the correct name to be "5α-pregnan-17α-ol-3,11,20-trione".

According to the rule set in the Nomenclature of Steroids, the terminal "e" in the parent structure name should be elided before the vowel (the presence or absence of a number does not affect such elision).[26] This means, for instance, that if the suffix immediately appended to the parent structure name begins with a vowel, the trailing "e" is removed from that name. An example of such removal is "5α-pregnan-17α-ol-3,20-dione", where the last "e" of "pregnane" is dropped due to the vowel ("o") at the beginning of the suffix -ol. Some authors incorrectly use this rule, eliding the terminal "e" where it should be kept, or vice versa.[24]

The term "11-oxygenated" refers to the presence of an oxygen atom as an oxo (=O) or hydroxy (-OH) substituent at carbon 11. "Oxygenated" is consistently used within the chemistry of the steroids[34] since the 1950s.[35] Some studies use the term "11-oxyandrogens"[36][37] as an abbreviation for 11-oxygenated androgens, to emphasize that they all have an oxygen atom attached to carbon at position 11.[38][39] However, in chemical nomenclature, the prefix "oxy" is associated with ether functional groups, i.e., a compound with an oxygen atom connected to two alkyl or aryl groups (R-O-R),[40] therefore, using "oxy" within the name of a steroid class may be misleading. One can find clear examples of "oxygenated" to refer to a broad class of organic molecules containing a variety of oxygen containing functional groups in other domains of organic chemistry,[41] and it is appropriate to use this convention here.

Even though "keto" is a standard prefix in organic chemistry, the 1989 recommendations of the Joint Commission on Biochemical Nomenclature discourage the application of the prefix "keto" for steroid names, and favor the prefix "oxo" (e.g., 11-oxo steroids rather than 11-keto steroids), because "keto" includes the carbon that is part of the steroid nucleus and the same carbon atom should not be specified twice.[42]

Biochemistry

A more detailed description of each alternative androgen pathway introduced in § History is provided below. Proteins are described by their Human Genome Organisation Gene Nomenclature Committee (HGNC) gene symbols[43][44], i.e. the symbols of the genes that they are encoded by (e.g., 5α-reductases type 1 is abbreviated by SRD5A1). Full enzyme/protein names are not shown in the text since the steroid names that are being used are already unwieldy, but they can be found in § Abbreviations and Identifiers.

Backdoor Pathways to 5α-Dihydrotestosterone

While 5α-reduction is the last transformation in the classical androgen pathway, it is the first step in the backdoor pathways to 5α-dihydrotestosterone that acts on either 17OHP or P4 which are ultimately converted to DHT.

 The androgen backdoor pathways from 17α-hydroxyprogesterone or progesterone towards 5α-dihydrotestosterone roundabout testosterone and androstenedione (red arrows), as well as the 5α-dione pathway that starts with 5α-reduction of androstenedione, embedded within classical androgen pathway (black arrows). Transformation arrows are annotated with the HGNC symbol (boxed text) of the gene for the enzyme that catalyzes the transformation. Some additional proteins that are required for specific transformations including steroidogenic acute regulatory protein (STAR), cytochromes b5 (CYB5A), cytochrome P450 reductase (POR) are not shown for clarity. Only negligible A4 is produced from 17OHP,[45] denoted as a dotted arrow.


17α-Hydroxyprogesterone Pathway

The first step of this pathway is the 5α-reduction of 17OHP to 5α-pregnan-17α-ol-3,20-dione (17OHDHP, since it is also known as 17α-hydroxy-dihydroprogesterone). The reaction is catalyzed by SRD5A1.[46][47] 17OHDHP is then converted to 5α-pregnane-3α,17α-diol-20-one (5α-Pdiol) via 3α-reduction by a 3α-hydroxysteroid dehydrogenase isozyme (AKR1C2 and AKR1C4)[21][48] or HSD17B6, that also has 3α-reduction activity.[49][50] 5α-Pdiol is also known as 17α-hydroxyallopregnanolone or 17OH-allopregnanolone. 5α-Pdiol is then converted to 5α-androstan-3α-ol-17-one (AST) by 17,20-lyase activity of CYP17A1 which cleaves a side-chain (C17-C20 bond) from the steroid nucleus, converting a C21 steroid (a pregnane) to C19 steroid (an androstane or androgen). AST is 17β-reduced to 5α-androstane-3α,17β-diol (3α-diol) by HSD17B3 or AKR1C3.[22] The final step is 3α-oxidation of 3α-diol in target tissues to DHT by an enzyme that has 3α-hydroxysteroid oxidase activity, such as AKR1C2,[51] HSD17B6, HSD17B10, RDH16, RDH5, and DHRS9.[47] This oxidation is not required in the classical androgen pathway.

The pathway can be summarized as:

  • 17OHP → 17OHDHP → 5α-Pdiol → AST → 3α-diol → DHT

Progesterone Pathway

The pathway from P4 to DHT is similar to that described above from 17OHP to DHT, but the initial substrate for 5α-reductase here is P4 rather than 17OHP. Placental P4 in the male fetus is the feedstock for the backdoor pathway found operating in multiple non-gonadal tissues.[21]

The first step in this pathway is 5α-reduction of P4 towards 5α-dihydroprogesterone (5α-DHP) by SRD5A1. 5α-DHP is then converted to allopregnanolone (AlloP5) via 3α-reduction by AKR1C2 or AKR1C4. AlloP5 is then converted to 5α-Pdiol by the 17α-hydroxylase activity of CYP17A1. This metabolic pathway proceeds analogously to DHT as the 17α-hydroxyprogesterone pathway described the previous subsection.

The pathway can be summarized as:

  • P4 → 5α-DHP → AlloP5 → 5α-Pdiol → AST → 3α-diol → DHT

5α-Dione Pathway

5α-reduction is also the initial transformation of the 5α-dione pathway where A4 is converted to androstanedione (5α-dione) by SRD5A1 and then directly to DHT by either HSD17B3 or AKR1C3. While this pathway is unlikely to be of biological relevance in healthy humans, it has been found operating in castration-resistant prostate cancer.[12]

5α-dione can also be transformed into AST, which can then either be converted back to 5α-dione or be transformed into DHT along the common part of the backdoor pathways to DHT (i.e., via 3α-diol).[12][52]

This pathway can be summarized as:

  • A4 → 5α-dione → DHT

11-Oxygenated Androgen Pathways

 Abbreviated routes to 11-oxygenated androgens with transformations (black arrow) annotated with the HGNC symbol (boxed text) of the gene for the enzyme that catalyzes the transformation. The four groups of steroids are distinguished by the C17 substituent configuration associated with four distinct precursors, A: the conversion of progesterone B: the conversion of 17-hydroxyprogesterone by CYP11B1, SRD5A, and AKR1C2 in the C11-hydroxy and -oxo C21 steroids C: the conversion of androstenedione D: the conversion of testosterone by CYP11B1, HSD11B, SRD5A, AKR1C3 and AKR1C2/4 in the C11-hydroxy and -oxo C19 steroids. CYP17A1 catalyzes the C21 steroids to C19 steroids. Some transformations which are presumed to exist but not yet shown to exist are depicted with dotted arrows. Some CYP17A1 mediated reactions that transform 11-oxygenated androgens classes (grey box) are omitted for clarity. Δ5 compounds that are transformed to Δ4 compounds are also omitted for clarity.


There are two known 11-oxygenated androgens that are both C19: 11KT and 11KDHT. Some work[53][54][55] suggests that though 11OHT and 11OHDHT may not have significant androgenic activity as they were once thought to possess, they may still be important precursors to androgenic molecules. The relative importance of the androgens depends on their activity, circulating levels and stability.


Routes to 11-oxygenated androgens (#Figure 3) also fall under our definition of alternative androgen pathways. These routes begin with four Δ4 steroid entry points (P4, 17OHP, A4 and T) and can then proceed through a set of transformation sequences that can be organized in a lattice-like structure. The reachability a particular steroid in the lattice depends on the expression of a given enzyme in the tissue where that steroid is synthesized or transported to, which in turn can depend on the health status of the individual. All the steroid products in this lattice have a hydroxy group or an oxo group covalently bound to the carbon atom at position 11 (#Figure 1).

The steroids 11OHA4 and 11KA4 have been established as having minimal androgen activity,[16][17][56] but remain important molecules in this context since they act as androgen precursors.

The complex lattice structure seen in #Figure 3 can be understood broadly as four Δ4 steroid entry points that can undergo a common sequence of three transformations:

  1. 11β-hydroxylation by CYP11B1 in the adrenal cortex,
  2. 5α-reduction by SRD5A1/2,
  3. reversible 3α-reduction/oxidation of the ketone/alcohol by AKR1C2 or AKR1C4.

These steroids correspond to the "C11-hydroxy" column in #Figure 3. This sequence is replicated in the parallel column of "C11-oxo" steroids, in which are a result of 11β-oxidation of the alcohol to a ketone.[17] HSD11B1 catalyzes both oxidation of the 11OH substrates and reduction of the 11K substrates, while HSD11B2 only catalyzes the oxidation. It should be noted that while 11β-hydroxylation by CYP11B2 has been shown in vitro, this isozyme is only known to be expressed in the zona glomerulosa of the adrenal cortex which would be unlikely to encounter the Δ4 substrates.

There are additional transformations in the lattice that cross the derivatives of the entry points. AKR1C3 catalyzes (reversibly in some cases) 17β-reduction of the ketone to transform between 17-oxo and 17β-OH steroids. Pregnanes can also be transformed to 17-oxo steroids via CYP17A1 17α-hydroxylase activity. Some 17α-OH pregnanes can be transformed into 17-oxo steroids via 17,20 lyase activity of CYP17A1.

There are many possible routes to androgens as seen in #Figure 3, but only three are known in healthy humans. A4 is synthesized in the adrenal where it can undergo 11β-hydroxylation to yield 11OHA4,[57][58][59] an important circulating androgen precursor, which is further transformed to 11KA4 and then 11KT (primarily outside the adrenal in peripheral tissue):

  • A4 → 11OHA4 → 11KA4 → 11KT

This route is regarded as the primary 11-oxygenated androgen pathway in healthy humans. It is thought that the T entry point also operates in normal human physiology, but much less that A4:

  • T → 11OHT → 11OHA4 → 11KA4 → 11KT
  • T → 11OHT → 11KT

The diminished role of these pathways is supported by that fact that the adrenal produces significantly more 11OHA4 than OHT.[16][60]

Other Backdoor Androgen Pathways

There is currently no good evidence for androgens from the C21 steroid entry points (P4, 17OHP) operating in healthy humans. These entry points are relevant in the clinical context, as discussed in the next section.

There is one report of 5α-androstane-3β,17β-diol (3β-diol) as a precursor to DHT[61] through a backdoor pathway, but this does not yet seem widely accepted in later reviews.[62]

Clinical Significance

11-Oxygenated Androgens

Characterizing serum concentration profiles of 11-oxygenated androgens in humans is technically challenging but essential for clinical applications. Just as with classical androgens, low concentrations, similarity of different androgens and cross-reactivity of molecules are just some of the barriers that must be mitigated to measure concentrations with sufficient accuracy for clinical utility. Circulating androgen levels do not always reflect tissue steroid concentrations as is evidenced by poor correlations between intra-tissue and blood levels. The analytic chemistry methods used to quantitate androgens can be technically difficult, and not all protocols have been validated to the same degree. In general, most published concentration profiles of 11-oxygenated androgens should be assumed to come with considerable uncertainty. While there is a long history to 11-oxygenated steroids in a number of diseases, there is no profile measurement that is considered a standard clinical tool.

Broadly, higher quality studies with better analytic methods that cover larger numbers of steroid types, careful disease subtyping etc. seem to be required before strong clinical applications can be expected.

A recent study[63] found 11KHDHT levels remained below the limit of detection (<20 pg/mL) and no commercial 11OHDHT standard was available to measure against. In the same study on samples drawn from a diverse pool of participants, serum 11KT levels were found to range between 197.63 pg/mL and 461.40 pg/mL, while 11OHT levels ranged between 87.34 pg/mL and 252.61 pg/mL. 11KT is thought to circulate at similar levels to T in healthy women.[64]

It may be that 11KT is the main androgen in women since it circulates at similar level to T and may[64] or may not[65][66] decline with age as T does. While 11KDHT is equipotent to DHT, circulating levels of 11KDHT are lower than DHT.

Unlike T and A4, 11-oxygenated androgens are not known to be aromatized to estrogens in the human body.[67][68] It is possible that 11-oxygenated estrogens may be produced in some pathological conditions such as feminizing adrenal carcinoma.[69]

Each condition in the following subsections has demonstrated potential roles for 11-oxygenated androgens. While there are no known C21 androgens (11-oxygenated or otherwise), they do arise as precursors to C19 androgens and some C21 steroid levels are thought to be associated with some of these conditions.

Hyperandrogenism

Alternative androgen pathways are not always considered in the clinical evaluation of patients with hyperandrogenism.[70] Hyperandrogenism may lead to presentations including congenital adrenal hyperplasia (CAH), other disorders of sex development (DSDs), polycystic ovary syndrome (PCOS), premature adrenarche and Cushing's disease.[71] Not considering alternative androgen pathways in clinical hyperandrogenism investigations may lead to misdiagnosis.[70]

Despite the prevailing notion that T and DHT are the primary human androgens, this notion only applies to healthy men.[72] Although T has been traditionally used as a biomarker of hyperandrogenism,[73] it correlates poorly with clinical measurements of androgen excess.[72] While T levels can appear normal, ignoring the alternative androgen pathways may lead to diagnostic errors since hyperandrogenism may be caused by potent androgens such as DHT produced by a backdoor pathway and 11-oxygenated androgens.[74][75]

It had been suggested that 11OHA4 and its urinary metabolites could have clinical applications as biomarkers of hyperandrogenism in women.[76] Increased adrenal 11OHA4 production was characterized, using changes in A4:11OHA4 and 11β-hydroxyandrosterone:11β-hydroxyetiocholanolone ratios, in Cushing's disease,[77] hirsutism,[78] CAH and PCOS.[79] These ratios have still not been established as a standard clinical as a diagnostic tool.

Congenital Adrenal Hyperplasia

CAH refers to a group of autosomal recessive disorders characterized by impaired cortisol biosynthesis[80] caused by a deficiency in any of the enzymes required to produce cortisol in the adrenal.[81][82] This deficiency leads to an excessive accumulation of steroid precursors that are converted to androgens and lead to hyperandrogenism. "Classical CAH" is generally considered as a DSD since it can lead to virilization and cliteromegaly in females, but it is potentially fatal in either sex. The "non-classical" or "late onset" forms of CAH are non-fatal and generally present with more subtle symptoms. The potential contributions from alternative androgen pathways remain under-appreciated in CAH.

In CAH due to deficiency of 21-hydroxylase[11] or cytochrome P450 oxidoreductase (POR),[47][83] the associated elevated 17OHP levels result in flux through the backdoor pathway to DHT that begins with 5α-reduction of 17OHP. This pathway may be activated regardless of age and sex.[84] Fetal excess of 17OHP in CAH may contribute to DHT synthesis that leads to external genital virilization in newborn girls with CAH.[47] P4 levels may also be elevated in CAH,[85][86] leading to androgen excess via the backdoor pathway from P4 to DHT.[87] 17OHP and P4 may also serve as substrates to 11-oxygenated androgens in CAH.[20]

Serum levels of the C21 11-oxygenated steroids 11OHP4 and 21dF have been known to be elevated in (non-classical/classical) CAH since about 1990,[88][89] and LC-MS/MS profiles that include these steroids have been proposed for clinical applications.[90] Classical CAH patients receiving glucocorticoid therapy had C19 11-oxygenated steroid serum levels that were elevated 3-4 fold compared to healthy controls.[91] In that same study, the levels of C19 11-oxygenated androgens correlated positively with conventional androgens in women but negatively in men. The levels of 11KT were 4 times higher compared to that of T in women with the condition. In adult women with CAH, the ratio of DHT produced in a backdoor pathway to that produced in a conventional pathway increases as control of androgen excess by glucocorticoid therapy deteriorates.[92] In CAH patients with poor disease control, 11-oxygenated androgens remain elevated for longer than 17OHP, thus serving as a better biomarker for the effectiveness of the disease control.[93][94] In males with CAH, 11-oxygenated androgen levels may indicate the presence testicular adrenal rest tumors.[94][95]

Other Disorders of Sex Development

Both the classical and backdoor androgen pathway to DHT are required for normal human male genital development.[83][96] Deficiencies in the backdoor pathway to DHT from 17OHP or from P4[48][46] lead to underverilization of the male fetus,[97][98] as placental P4 is a precursor to DHT in the backdoor pathway.[21]

A case study[48] of five 46,XY (male) patients from two families demonstrated that atypical genital appearance were attributed to mutations in AKR1C2 and/or AKR1C4, which operate exclusively in the backdoor pathway to DHT. Mutations in the AKR1C3 and genes involved in the classical androgen pathway were excluded as the causes for the atypical appearance. The 46,XX (female) relatives of affected patients, having the same mutations, were phenotypically normal and fertile. Although both AKR1C2 and AKR1C4 are needed for DHT synthesis in a backdoor pathway (#Figure 2), the study found that mutations in AKR1C2 only were sufficient for disruption.[48] However, these AKR1C2/AKR1C4 variants leading to DSD are rare and have been only so far reported in just those two families.[99] This case study emphasizes the role of AKR1C2/4 in the alternative androgen pathways.

Isolated 17,20-lyase deficiency syndrome due to variants in CYP17A1, cytochrome b5, and POR may also disrupt the backdoor pathway to DHT, as the 17,20-lyase activity of CYP17A1 is required for both classical and backdoor androgen pathways. This rare deficiency can lead to DSD in both sexes, with affected girls being asymptomatic until puberty, when they show amenorrhea.[99]

11-oxygenated androgens may play important roles in DSDs.[100][101][47] 11-oxygenated androgen fetal biosynthesis may coincide with the key stages of production of cortisol — at weeks 8–9, 13–24, and from 31 and onwards. In these stages, impaired CYP17A1 and CYP21A2 activity lead to increased ACTH due to cortisol deficiency and the accumulation of substrates for CYP11B1 in pathways to 11-oxygenated androgens and could cause abnormal female fetal development.[100]

Polycystic Ovary Syndrome

PCOS is a heterogeneous and complex disorder that affects women's metabolism and fertility, and is poorly understood in terms of underlying causes. The diagnoses of PCOS that require hyperandrogenism tend to be associated with more severe disease phenotypes.[102] Both the ovary and the adrenal can be involved in the hyperandrogenism.

In PCOS, DHT may be produced in the backdoor androgen pathway from SRD5A1 activity.[103][104] Genes encoding enzymes required for the backdoor pathway to 5α-DHT (AKR1C2/4, SRD5A1/2, RDH16) are expressed at higher levels than normal in the theca cells of the PCOS ovary,[105] and could be responsible for hyperandrogenism.

There is a decades long history of research around the application C19 11-oxygenated steroids (11OHA4, 11KA4, 11OHT and 11KT) as clinical disease markers.[39] Perhaps glibly summarizing this large body of work, there still seems to be a considerable lack of consensus as to which steroid profile might be useful in a clinical context. Some of the controversy can be attributed to the complex diagnostic framework around PCOS.

Premature Adrenarche

Although premature adrenarche has long been considered benign, the early appearance of pubic and/or axillary hair in girls (before 8 years of age) and boys (before 9 years of age) has been linked to several pathological risks, including cardiovascular risks and PCOS. The cause is unknown, but alternative androgen pathways have been implicated. Urinary metabolite analysis has been used to infer that there is increased 17,20-lyase activity of CYP17A1 and that it is likely that it leads to increased metabolic flux through the backdoor androgen pathway to DHT.[106] Another study observed 11KT/T ratios being higher in premature adrenarche girls (3.5 fold) vs. age-matched controls (2.5).[107] These findings may further understanding of the condition and associated risks.

Cushing's Disease

Cushing's disease itself characterized by cortisol excess, and patients commonly present with symptoms of hyperandrogenism, but these symptoms have not been correlated with serum androgen concentrations. A recent study[77] demonstrated that salivary concentration profiles of 11-oxygenated steroids, including 11OHA4 and 11KT, were elevated in untreated Cushing's patients, and levels were reduced with treatment. The effect was strong enough for the study to claim that hyperandrogenism in Cushing's disease was caused by an excess of 11-oxygenated C19 steroids.

Prostate Cancer

High levels of 11KT, 11KDHT and 11OHDHT have also been detected in prostate cancer tissue[108] and there is some preliminary evidence that C19 11-oxygenated steroids may play an important role at the stage of prostatic carcinogenesis.[109] At the time of writing, it can be difficult to summarize some of the conflicting claims about the relative abundance/importance of different steroids in prostate cancer.

Androgen deprivation is a therapeutic approach to prostate cancer that can be implemented by castration to eliminate gonadal T, but metastatic tumors may then develop into castration-resistant prostate cancer (CRPC). Although castration results in 90-95% decrease of serum T, DHT in the prostate is only decreased by 50%, supporting the notion that the prostate expresses necessary enzymes to produce DHT without testicular T.[14] The 5α-dione pathway was discovered in the context of CPRC (see § History), and is known to mitigate the effects of androgen deprivation therapy.

C19 11-oxygenated androgens contribute significantly to the androgen pool[17][22] and play a previously overlooked role in the reactivation of androgen signaling in the CRPC patient. Serum 11KT levels are higher than any other androgen in 97% of CRPC patients, accounting for 60% of the total active androgen pool, and are not affected by castration.[110]

Benign Prostatic Hyperplasia, Chronic Prostatitis/Chronic Pelvic Pain Syndrome

Androgens play a vital role in the development, growth and maintenance of the prostate.[14] The role of androgens has been considered not only in prostate cancer, but other prostate-related conditions such as BPH[14] and chronic prostatitis/chronic pelvic pain syndrome (CP/CPPS).[111] Just as in prostate cancer, conflicting claims about the relative abundance/importance of different steroids demand resolution.

C21 11-oxygenated steroids have been demonstrated as androgen precursors in BPH cell models (11OHP4 and 11KP4 to 11KDHT), BPH patient tissue biopsy and in their serum.[112]

Future Directions

The role of non-classical CAH in CP/CPPS

Relative steroid serum levels in CP/CPPS have suggested that CYP21A2 deficiency may play a role in the disease and that non-classical (mild) CAH due to CYP21A2 deficiency may be a comorbidity, even though non-classical CAH is generally thought to be asymptomatic in men.[113][114] Given the important role that androgens play in the health of the prostate, at least one study has hypothesized that CP/CPPS may be a consequence of a systemic condition as opposed to a localized condition of the prostate such as an infection, inflammation, or dysfunction.[111][115][116] Considering the potential roles that alternative androgen pathways play in the previously described disease areas, it seems that CP/CPPS would seem to be a good candidate to investigate the same way. The authors are not aware of any work that describes alternative androgen pathways in CP/CPPS.

The biomarkers of disease control in CAH due to CYP21A2 deficiency

Some studies suggested that in CAH patients with poor disease control, 11-oxygenated androgens remain elevated for longer than 17OHP, thus serving as a better biomarker for the effectiveness of the disease control than the traditional indicator 17OHP.[93][94] It may be that 11OHP4 and 21dF, the substrates for 11-oxygenated androgens in pathologic conditions such as CAH rather than 11-oxygenated androgens themselves may serve as better biomarkers for the effectiveness of the disease control in CAH. 21dF has already been proposed as a better biomarker for CAH diagnosis than 17OHP,[117] but the role of 21dF and 11OHP4 as biomarkers of the effectiveness of the disease control in CAH remain to be studied.

SRD5A2 in the backdoor pathway to DHT

While the role of SRD5A1 in the backdoor pathway to DHT is established, it is not clear if SRD5A2 is involved.[11] Some authors[46][47] claim that the reduction of 17OHP to 17OHDHP by SRD5A1 is not "sufficient" or "efficient", as supported by measurements of rat SRD5A2 activity.[118] More recently, it has been shown that recombinant human SRD5A1 and SRD5A2 can catalyze the reduction of 17OHP at comparable rates to the reduction of P4.[5] Given both isozymes may be expressed in fetal tissues of both sexes,[119][120] the action of SRD5A2 in the backdoor pathway to DHT in humans may be worth more exploration.

Immune Response Regulation

Androgens are known to regulate the processes inside AR expressing immune cells and are assumed to play a role in sex differences in immune responses.[121][122][123] The roles of androgens in immune responses have been investigated in many clinical settings including the worldwide COVID-19 pandemic,[124][125] with enough investigation of anti-androgen therapy for meta-analysis.[126]

There seem to be relatively few reports on the links between alternative androgen pathways and the immune system though it has been shown that peripheral mononuclear blood cells produce 11KT from 11KA4 and that 11KT is the predominant androgen in that cell type.[127] Understanding what, if any, effects or therapeutic targets alternative androgen pathways offer in the context of immune processes may become an important research direction as the understanding of these pathways improves.

Abbreviations and Identifiers

Steroids

Abbreviation Systematic Name Trivial Name[i] Other Names and Abbreviations PubChem CID
11dF 4-pregnene-17α,21-diol-3,20-dione 11-deoxycortisol cortodoxone; 11-desoxycortisol; Reichstein's substance S 440707
11K-3αdiol 5α-androstane-3α,17β-diol-11-one 10125601
11K-5αdione 5α-androstane-3,11,17-trione 11-ketoandrostanedione 11-keto-5α-androstanedione 11185733
11KA4 4-androstene-3,11,17-trione 11-ketoandrostenedione androst-4-ene-3,11,17-trione; adrenosterone; Reichstein's substance G 223997
11KAST 5α-androstan-3α-ol-11,17-dione 11-ketoandrosterone 102029
11KDHP4 5α-pregnane-3,11,20-trione 11-ketodihydroprogesterone allopregnanetrione 968899
11KDHT 5α-androstan-17β-ol-3,11-dione 11-ketodihydrotestosterone 5α-dihydro-11-keto testosterone; 5α-dihydro-11-keto-testosterone 11197479
11KP4 4-pregnene-3,11,20-trione 11-ketoprogesterone pregn-4-ene-3,11,20-trione 94166
11KPdiol 5α-pregnane-3α,17α-diol-11,20-dione 92264183
11KPdione 5α-pregnan-17α-ol-3,11,20-trione 99568471
11KT 4-androsten-17β-ol-3,11-dione 11-ketotestosterone 5282365
11OH-3αdiol 5α-androstane-3α,11β,17β-triol 10286384
11OH-5αdione 5α-androstan-11β-ol-3,17-dione 11β-hydroxy-5α-androstanedione adrenosterone-M 59087027
11OHA4 4-androsten-11β-ol-3,17-dione 11β-hydroxyandrostenedione androst-4-en-11β-ol-3,17-dione 94141
11OHAST 5α-androstane-3α,11β-diol-17-one 11β-hydroxyandrosterone 10286365
11OHDHP4 5α-pregnan-11β-ol-3,20-dione 11β-hydroxydihydroprogesterone 11267580
11OHDHT 5α-androstane-11β,17β-diol-3-one 11β-hydroxydihydrotestosterone 5α-dihydro-11β-hydroxytestosterone; 11β,17β-dihydroxy-5α-androstan-3-one 10018051
11OHEt 5β-androstane-3α,11β-diol-17-one 11β-hydroxyetiocholanolone 3α,11β-dihydroxy-5β-androstan-17-one 101849
11OHP4 4-pregnen-11β-ol-3,20-dione 21-deoxycorticosterone pregn-4-en-11β-ol-3,20-dione; 21-desoxycorticosterone; 11β-hydroxyprogesterone 101788
11OHPdiol 5α-pregnane-3α,11β,17α-triol-20-one 99601857
11OHPdione 5α-pregnane-11β,17α-diol-3,20-dione 99572627
11OHT 4-androstene-11β,17β-diol-3-one 11β-hydroxytestosterone 114920
17OHP5 5-pregnene-3β,17α-diol-20-one 17α-hydroxypregnenolone 91451
17OHDHP 5α-pregnan-17α-ol-3,20-dione 17α-hydroxydihydroprogesterone 5α-Pdione 11889565
17OHP 4-pregnen-17α-ol-3,20-dione 17α-hydroxyprogesterone 6238
21dE 4-pregnen-17α-ol-3,11,20-trione 21-deoxycortisone pregn-4-en-17α-ol-3,11,20-trione; 21-desoxycortisone 102178
21dF 4-pregnene-11β,17α-diol-3,20-dione 21-deoxycortisol 1β,17α-dihydroxyprogesterone; pregn-4-ene-11β,17α-diol-3,20-dione; 21-desoxycortisol; 21-deoxyhydrocortisone 92827
3,11diOH-DHP4 5α-pregnane-3α,11β-diol-20-one 3α,11β-dihydroxy-5α-pregnan-20-one 10125849
3α-diol 5α-androstane-3α,17β-diol 5α-adiol; 3α-androstanediol 15818
3β-diol 5α-androstane-3β,17β-diol 3β-androstanediol 242332
5α-DHP 5α-pregnane-3,20-dione 5α-dihydroprogesterone 92810
5α-dione 5α-androstane-3,17-dione androstanedione 222865
5α-Pdiol 5α-pregnane-3α,17α-diol-20-one 17α-hydroxyallopregnanolone 111243
A4 4-androstene-3,17-dione androstenedione androst-4-ene-3,17-dione 6128
A5 5-androstene-3β,17β-diol androstenediol androst-5-ene-3β,17β-diol 10634
A5-S 5-androstene-3β,17β-diol sulfate androstenediol sulfate 13847309
ALDO 4-pregnene-11β,21-diol-3,18,20-trione aldosterone 5839
ALF 5α-pregnan-3α-ol-11,20-dione alfaxalone alphaxalone; alphaxolone; 3α-hydroxy-5α-pregnane-11,20-dione 104845
AlloP5 5α-pregnan-3α-ol-20-one allopregnanolone 92786
AST 5α-androstan-3α-ol-17-one androsterone 5879
CORT 4-Pregnene-11β,21-diol-3,20-dione corticosterone Kendall's compound B; Reichstein's substance H 5753
DHEA 5-androsten-3β-ol-17-one dehydroepiandrosterone 3β-hydroxyandrost-5-en-17-one; androst-5-en-3β-ol-17-one 5881
DHEA-S 5-androsten-3β-ol-17-one sulfate dehydroepiandrosterone sulfate 12594
DHT 5α-androstan-17β-ol-3-one 5α-dihydrotestosterone 10635
DOC 4-pregnen-21-ol,3,20-dione 11-deoxycorticosterone desoxycorticosterone; 11-desoxycorticosterone; Reichstein's substance Q; desoxycortone 6166
F 4-pregnene-11β,17α,21-triol-3,20-dione cortisol hydrocortisone; Kendall's compound F; Reichstein's substance M 5754
P4 4-pregnen-3,20-dione progesterone 5994
P5 5-pregnen-3β-ol-20-one pregnenolone 8955
T 4-androsten-17β-ol-3-one testosterone 6013

Note:

  1. The steroids in this paper are referred to by their trivial names, as specified in the respective column. If a steroid has no conventional trivial name, then the systematic name is used for that purpose.

Enzymes

HGNC Gene Symbol Enzyme Name
AKR1C2 aldo-keto reductase family 1 member C2; 3α-hydroxysteroid dehydrogenase type 3
AKR1C3 aldo-keto reductase family 1 member C3; 3α-hydroxysteroid dehydrogenase type 2; 17β-hydroxysteroid dehydrogenase type 5; HSD17B5
AKR1C4 aldo-keto reductase family 1 member C4; 3α-hydroxysteroid dehydrogenase type 1
CYP11A1 cytochrome P450 cholesterol side-chain cleavage enzyme; P450scc
CYP11B1 cytochrome P450 11β-hydroxylase; P450c11B1
CYP11B2 aldosterone synthase; P450c11B2
CYP17A1 cytochrome P450 17α-hydroxylase/17,20-lyase; P450c17
CYP21A2 cytochrome P450 21-hydroxylase; P450c21
DHRS9 dehydrogenase/reductase SDR family member 9; short-chain dehydrogenases/reductase 9
HSD11B1 11β-hydroxysteroid dehydrogenase type 1
HSD11B2 11β-hydroxysteroid dehydrogenase type 2
HSD17B3 17β-hydroxysteroid dehydrogenase type 3
HSD17B6 17β-hydroxysteroid dehydrogenase type 6; retinol dehydrogenase-like hydroxysteroid dehydrogenase; RL-HSD
HSD17B10 17β-hydroxysteroid dehydrogenase type 10
POR cytochrome P450 oxidoreductase
RDH16 retinol dehydrogenase 16; RODH4
RDH5 retinol dehydrogenase 5
SRD5A1 steroid 5α-reductase type 1; 3-oxo-5α-steroid 4-dehydrogenase type 1
SRD5A2 steroid 5α-reductase type 2; 3-oxo-5α-steroid 4-dehydrogenase type 2
SRD5A3 steroid 5α-reductase type 3; 3-oxo-5α-steroid 4-dehydrogenase type 3

Conditions

Abbreviation Full Name
21OHD 21-hydroxylase deficiency
BPH benign prostatic hyperplasia
CAH congenital adrenal hyperplasia
CP/CPPS chronic prostatitis/chronic pelvic pain syndrome
CRPC castration-resistant prostate cancer
DSD disorder of sex development
PCOS polycystic ovary syndrome

Other Abbreviations

Abbreviation Full Name
ACTH adrenocorticotropic hormone
HGNC Human Genome Organisation Gene Nomenclature Committee
STAR steroidogenic acute regulatory protein

Additional Information

Acknowledgements

The authors would like to thank the reviewers for their immense efforts in guidance and corrections.

Competing Interests

The authors have no competing interest.

Funding

The authors received no financial support for the research, authorship and publication of this article.

References

  1. Gelmann, Edward P. (2022). "Molecular Biology of the Androgen Receptor". Journal of Clinical Oncology 20 (13): 3001–3015. doi:10.1200/JCO.2002.10.018. ISSN 0732-183X. PMID 12089231. https://ascopubs.org/doi/10.1200/JCO.2002.10.018. 
  2. Li, Jin; Al-Azzawi, Farook (2009). "Mechanism of androgen receptor action". Maturitas 63 (2): 142–148. doi:10.1016/j.maturitas.2009.03.008. https://linkinghub.elsevier.com/retrieve/pii/S0378512209000978. 
  3. 3.0 3.1 Wilson, Jean D.; Auchus, Richard J.; Leihy, Michael W.; Guryev, Oleg L.; Estabrook, Ronald W.; Osborn, Susan M.; Shaw, Geoffrey; Renfree, Marilyn B. (2003). "5alpha-androstane-3alpha,17beta-diol is formed in tammar wallaby pouch young testes by a pathway involving 5alpha-pregnane-3alpha,17alpha-diol-20-one as a key intermediate". Endocrinology 144 (2): 575–80. doi:10.1210/en.2002-220721. PMID 12538619. 
  4. 4.0 4.1 Auchus, Richard J. (2004). "The backdoor pathway to dihydrotestosterone". Trends in Endocrinology and Metabolism: TEM 15 (9): 432–8. doi:10.1016/j.tem.2004.09.004. PMID 15519890. 
  5. 5.0 5.1 5.2 Barnard, Lise; Gent, Rachelle; Van Rooyen, Desmaré; Swart, Amanda C. (2017). "Adrenal C11-oxy C21 steroids contribute to the C11-oxy C19 steroid pool via the backdoor pathway in the biosynthesis and metabolism of 21-deoxycortisol and 21-deoxycortisone". The Journal of Steroid Biochemistry and Molecular Biology 174: 86–95. doi:10.1016/j.jsbmb.2017.07.034. PMID 28774496. https://www.sciencedirect.com/science/article/abs/pii/S0960076017302091. 
  6. Pham, Nhung; van Heck, Ruben G. A.; van Dam, Jesse C. J.; Schaap, Peter J.; Saccenti, Edoardo; Suarez-Diez, Maria (2019). "Consistency, Inconsistency, and Ambiguity of Metabolite Names in Biochemical Databases Used for Genome-Scale Metabolic Modelling". Metabolites 9 (2): 28. doi:10.3390/metabo9020028. ISSN 2218-1989. PMID 30736318. PMC 6409771. //www.ncbi.nlm.nih.gov/pmc/articles/PMC6409771/. 
  7. Eckstein, B.; Borut, A.; Cohen, S. (1987). "Metabolic pathways for androstanediol formation in immature rat testis microsomes". Biochimica et Biophysica Acta (BBA) - General Subjects 924 (1): 1–6. doi:10.1016/0304-4165(87)90063-8. ISSN 0006-3002. PMID 3828389. 
  8. Shaw, G.; Renfree, M. B.; Leihy, M. W.; Shackleton, C. H.; Roitman, E.; Wilson, J. D. (2000). "Prostate formation in a marsupial is mediated by the testicular androgen 5 alpha-androstane-3 alpha,17 beta-diol". Proceedings of the National Academy of Sciences of the United States of America 97 (22): 12256–12259. doi:10.1073/pnas.220412297. ISSN 0027-8424. PMID 11035809. PMC 17328. //www.ncbi.nlm.nih.gov/pmc/articles/PMC17328/. 
  9. Penning, Trevor M. (1997). "Molecular Endocrinology of Hydroxysteroid Dehydrogenases". Endocrine Reviews 18 (3): 281–305. doi:10.1210/edrv.18.3.0302. ISSN 0163-769X. PMID 9183566. https://academic.oup.com/edrv/article/18/3/281/2530742. 
  10. Mahendroo, Mala; Wilson, Jean D.; Richardson, James A.; Auchus, Richard J. (2004). "Steroid 5alpha-reductase 1 promotes 5alpha-androstane-3alpha,17beta-diol synthesis in immature mouse testes by two pathways". Molecular and Cellular Endocrinology 222 (1–2): 113–120. doi:10.1016/j.mce.2004.04.009. ISSN 0303-7207. PMID 15249131. 
  11. 11.0 11.1 11.2 Kamrath, Clemens; Hochberg, Ze'ev; Hartmann, Michaela F.; Remer, Thomas; Wudy, Stefan A. (2012). "Increased activation of the alternative "backdoor" pathway in patients with 21-hydroxylase deficiency: evidence from urinary steroid hormone analysis". The Journal of Clinical Endocrinology and Metabolism 97 (3): E367–375. doi:10.1210/jc.2011-1997. ISSN 1945-7197. PMID 22170725. 
  12. 12.0 12.1 12.2 Chang, K.-H.; Li, R.; Papari-Zareei, M.; Watumull, L.; Zhao, Y. D.; Auchus, R. J.; Sharifi, N. (2011). "Dihydrotestosterone synthesis bypasses testosterone to drive castration-resistant prostate cancer". Proceedings of the National Academy of Sciences of the United States of America (Proceedings of the National Academy of Sciences) 108 (33): 13728–13733. doi:10.1073/pnas.1107898108. ISSN 0027-8424. PMID 21795608. PMC 3158152. //www.ncbi.nlm.nih.gov/pmc/articles/PMC3158152/. 
  13. Sharifi, Nima (2012). "The 5α-androstanedione pathway to dihydrotestosterone in castration-resistant prostate cancer". J Investig Med 60 (2): 504–7. doi:10.2310/JIM.0b013e31823874a4. PMID 22064602. PMC 3262939. //www.ncbi.nlm.nih.gov/pmc/articles/PMC3262939/. 
  14. 14.0 14.1 14.2 14.3 Luu-The, Van; Bélanger, Alain; Labrie, Fernand (2008). "Androgen biosynthetic pathways in the human prostate". Best Practice & Research. Clinical Endocrinology & Metabolism (Elsevier BV) 22 (2): 207–221. doi:10.1016/j.beem.2008.01.008. ISSN 1521-690X. PMID 18471780. 
  15. Pretorius, Elzette; Arlt, Wiebke; Storbeck, Karl-Heinz (2017). "A new dawn for androgens: Novel lessons from 11-oxygenated C19 steroids". Mol Cell Endocrinol 441: 76–85. doi:10.1016/j.mce.2016.08.014. PMID 27519632. http://pure-oai.bham.ac.uk/ws/files/30346231/Pretorius_et_al_manuscript.pdf. 
  16. 16.0 16.1 16.2 Rege, Juilee; Nakamura, Yasuhiro; Satoh, Fumitoshi; Morimoto, Ryo; Kennedy, Michael R.; Layman, Lawrence C.; Honma, Seijiro; Sasano, Hironobu et al. (2013). "Liquid chromatography-tandem mass spectrometry analysis of human adrenal vein 19-carbon steroids before and after ACTH stimulation". J Clin Endocrinol Metab 98 (3): 1182–8. doi:10.1210/jc.2012-2912. PMID 23386646. PMC 3590473. //www.ncbi.nlm.nih.gov/pmc/articles/PMC3590473/. 
  17. 17.0 17.1 17.2 17.3 17.4 Storbeck, Karl-Heinz; Bloem, Liezl M.; Africander, Donita; Schloms, Lindie; Swart, Pieter; Swart, Amanda C. (2013). "11β-Hydroxydihydrotestosterone and 11-ketodihydrotestosterone, novel C19 steroids with androgenic activity: a putative role in castration resistant prostate cancer?". Mol Cell Endocrinol 377 (1–2): 135–46. doi:10.1016/j.mce.2013.07.006. PMID 23856005. 
  18. Pretorius, Elzette; Africander, Donita J.; Vlok, Maré; Perkins, Meghan S.; Quanson, Jonathan; Storbeck, Karl-Heinz (2016). "11-Ketotestosterone and 11-Ketodihydrotestosterone in Castration Resistant Prostate Cancer: Potent Androgens Which Can No Longer Be Ignored". PLOS ONE 11 (7): e0159867. doi:10.1371/journal.pone.0159867. PMID 27442248. PMC 4956299. //www.ncbi.nlm.nih.gov/pmc/articles/PMC4956299/. 
  19. van Rooyen, Desmaré; Gent, Rachelle; Barnard, Lise; Swart, Amanda C. (2018). "The in vitro metabolism of 11β-hydroxyprogesterone and 11-ketoprogesterone to 11-ketodihydrotestosterone in the backdoor pathway". The Journal of Steroid Biochemistry and Molecular Biology 178: 203–212. doi:10.1016/j.jsbmb.2017.12.014. PMID 29277707. 
  20. 20.0 20.1 Van Rooyen, Desmaré; Yadav, Rahul; Scott, Emily E.; Swart, Amanda C. (2020). "CYP17A1 exhibits 17αhydroxylase/17,20-lyase activity towards 11β-hydroxyprogesterone and 11-ketoprogesterone metabolites in the C11-oxy backdoor pathway". The Journal of Steroid Biochemistry and Molecular Biology 199: 105614. doi:10.1016/j.jsbmb.2020.105614. PMID 32007561. 
  21. 21.0 21.1 21.2 21.3 O'Shaughnessy, Peter J.; Antignac, Jean Philippe; Le Bizec, Bruno; Morvan, Marie-Line; Svechnikov, Konstantin; Söder, Olle; Savchuk, Iuliia; Monteiro, Ana et al. (2019). "Alternative (backdoor) androgen production and masculinization in the human fetus". PLOS Biology 17 (2): e3000002. doi:10.1371/journal.pbio.3000002. PMID 30763313. PMC 6375548. //www.ncbi.nlm.nih.gov/pmc/articles/PMC6375548/. 
  22. 22.0 22.1 22.2 Storbeck, Karl-Heinz; Mostaghel, Elahe A. (2019). "Canonical and Noncanonical Androgen Metabolism and Activity". Advances in Experimental Medicine and Biology 1210: 239–277. doi:10.1007/978-3-030-32656-2_11. ISBN 978-3-030-32655-5. PMID 31900912. 
  23. 23.0 23.1 "Google Scholar search results for "5α-pregnan-17α-diol-3,11,20-trione" that is an incorrect name". 2022.
  24. 24.0 24.1 "Google Scholar search results for "5α-pregnane-17α-ol-3,20-dione" that is an incorrect name". 2022.
  25. "IUPAC-IUB Joint Commission on Biochemical Nomenclature (JCBN). The nomenclature of steroids. Recommendations 1989". Eur J Biochem 186 (3): 430. 1989. doi:10.1111/j.1432-1033.1989.tb15228.x. PMID 2606099. "3S-1.0. Definition of steroids and sterols Steroids are compounds possessing the skeleton of cyclopenta[a]phenanthrene or a skeleton derived therefrom by one or more bond scissions or ring expansions or contractions. Methyl groups are normally present at C-10 and C-13. An alkyl side chain may also be present at C-17.Sterols are steroids carrying a hydroxyl group at C-3 and most of the skeleton of cholestane." 
  26. 26.0 26.1 "IUPAC-IUB Joint Commission on Biochemical Nomenclature (JCBN). The nomenclature of steroids. Recommendations 1989". Eur J Biochem 186 (3). doi:10.1111/j.1432-1033.1989.tb15228.x. PMID 2606099. "3S-4. FUNCTIONAL GROUPS 3S-4.0. General Nearly all biologically important steroids are derivatives of the parent hydrocarbons (cf. Table 1) carrying various functional groups. [...] Suffixes are added to the name of the saturated or unsaturated parent system (see 33-2.5), the terminal e of -ane, -ene, -yne, -adiene etc. being elided before a vowel (presence or absence of numerals has no effect on such elisions)." 
  27. 3S-1.4. "IUPAC-IUB Joint Commission on Biochemical Nomenclature (JCBN). The nomenclature of steroids. Recommendations 1989". Eur J Biochem 186 (3). 1989. doi:10.1111/j.1432-1033.1989.tb15228.x. PMID 2606099. "3S-1.4. Orientation of projection formulae When the rings of a steroid are denoted as projections onto the plane of the paper, the formula is normally to be oriented as in 2a. An atom or group attached to a ring depicted as in the orientation 2a is termed α (alpha) if it lies below the plane of the paper or β (beta) if it lies above the plane of the paper." 
  28. Favre, Henri; Powell, Warren (2014). "P-91". Nomenclature of Organic Chemistry - IUPAC Recommendations and Preferred Names 2013. The Royal Society of Chemistry. doi:10.1039/9781849733069. ISBN 978-0-85404-182-4. "P-91.2.1.1 Cahn-Ingold-Prelog (CIP) stereodescriptors Some stereodescriptors described in the Cahn-Ingold-Prelog (CIP) priority system, called ‘CIP stereodescriptors’, are recommended to specify the configuration of organic compounds, as described and exemplified in this Chapter and applied in Chapters P-1 through P-8, and in the nomenclature of natural products in Chapter P-10. The following stereodescriptors are used as preferred stereodescriptors (see P-92.1.2): (a) ‘R’ and ‘S’, to designate the absolute configuration of tetracoordinate (quadriligant) chirality centers;" 
  29. Favre, Henri; Powell, Warren (2014). "P-13.8.1.1". Nomenclature of Organic Chemistry - IUPAC Recommendations and Preferred Names 2013. The Royal Society of Chemistry. doi:10.1039/9781849733069. ISBN 978-0-85404-182-4. "P-13.8.1.1 The prefix ‘de’ (not ‘des’), followed by the name of a group or atom (other than hydrogen), denotes removal (or loss) of that group and addition of the necessary hydrogen atoms, i.e., exchange of that group with hydrogen atoms. As an exception, ‘deoxy’, when applied to hydroxy compounds, denotes the removal of an oxygen atom from an –OH group with the reconnection of the hydrogen atom. ‘Deoxy’ is extensively used as a subtractive prefix in carbohydrate nomenclature (see P-102.5.3)." 
  30. "IUPAC-IUB Joint Commission on Biochemical Nomenclature (JCBN). The nomenclature of steroids. Recommendations 1989". Eur J Biochem 186 (3): 430. 1989. doi:10.1111/j.1432-1033.1989.tb15228.x. PMID 2606099. "3S-1.1. Numbering and ring letters Steroids are numbered and rings are lettered as in formula 1" 
  31. 31.0 31.1 "IUPAC-IUB Joint Commission on Biochemical Nomenclature (JCBN). The nomenclature of steroids. Recommendations 1989". Eur J Biochem 186 (3): 436–437. doi:10.1111/j.1432-1033.1989.tb15228.x. PMID 2606099. "3S-2.5 Unsaturation Unsaturation is indicated by changing -ane to -ene, -adiene, -yne etc., or -an- to -en-, -adien-, -yn- etc. Examples: Androst-5-ene, not 5-androstene 5α-Cholest-6-ene 5β-Cholesta-7,9(11)-diene 5α-Cholest-6-en-3β-ol Notes 1) It is now recommended that the locant of a double bond is always adjacent to the syllable designating the unsaturation. [...] 3) The use of Δ (Greek capital delta) character is not recommended to designate unsaturation in individual names. It may be used, however, in generic terms, like ‘Δ5-steroids’" 
  32. Favre, Henri; Powell, Warren (2014). "P-3". Nomenclature of Organic Chemistry - IUPAC Recommendations and Preferred Names 2013. The Royal Society of Chemistry. doi:10.1039/9781849733069. ISBN 978-0-85404-182-4. "P-31.2.2 General methodology ‘Hydro’ and ‘dehydro’ prefixes are associated with hydrogenation and dehydrogenation, respectively, of a double bond; thus, multiplying prefixes of even values, as ‘di’, ‘tetra’, etc. are used to indicate the saturation of double bond(s), for example ‘dihydro’, ‘tetrahydro’; or creation of double (or triple) bonds, as ‘didehydro’, etc. In names, they are placed immediately at the front of the name of the parent hydride and in front of any nondetachable prefixes. Indicated hydrogen atoms have priority over ‘hydro‘ prefixes for low locants. If indicated hydrogen atoms are present in a name, the ‘hydro‘ prefixes precede them." 
  33. Miller, Walter L.; Auchus, Richard J. (2011). "The molecular biology, biochemistry, and physiology of human steroidogenesis and its disorders". Endocr Rev 32 (1): 81–151. doi:10.1210/er.2010-0013. PMID 21051590. PMC 3365799. //www.ncbi.nlm.nih.gov/pmc/articles/PMC3365799/. 
  34. Makin, H.L.J.; Trafford, D.J.H. (1972). "The chemistry of the steroids". Clinics in Endocrinology and Metabolism 1 (2): 333–360. doi:10.1016/S0300-595X(72)80024-0. 
  35. Bongiovanni, A. M.; Clayton, G. W. (1954). "Simplified method for estimation of 11-oxygenated neutral 17-ketosteroids in urine of individuals with adrenocortical hyperplasia". Proc Soc Exp Biol Med 85 (3): 428–9. doi:10.3181/00379727-85-20905. PMID 13167092. 
  36. Slaunwhite, W.Roy; Neely, Lavalle; Sandberg, Avery A. (1964). "The metabolism of 11-Oxyandrogens in human subjects". Steroids 3 (4): 391–416. doi:10.1016/0039-128X(64)90003-0. 
  37. Taylor, Anya E.; Ware, Meredith A.; Breslow, Emily; Pyle, Laura; Severn, Cameron; Nadeau, Kristen J.; Chan, Christine L.; Kelsey, Megan M. et al. (2022). "11-Oxyandrogens in Adolescents With Polycystic Ovary Syndrome". J Endocr Soc 6 (7): bvac037. doi:10.1210/jendso/bvac037. PMID 35611324. PMC 9123281. //www.ncbi.nlm.nih.gov/pmc/articles/PMC9123281/. 
  38. Turcu, Adina F.; Rege, Juilee; Auchus, Richard J.; Rainey, William E. (2020). "11-Oxygenated androgens in health and disease". Nat Rev Endocrinol 16 (5): 284–296. doi:10.1038/s41574-020-0336-x. PMID 32203405. PMC 7881526. //www.ncbi.nlm.nih.gov/pmc/articles/PMC7881526/. 
  39. 39.0 39.1 Barnard, Lise; du Toit, Therina; Swart, Amanda C. (2021). "Back where it belongs: 11β-hydroxyandrostenedione compels the re-assessment of C11-oxy androgens in steroidogenesis". Molecular and Cellular Endocrinology 525: 111189. doi:10.1016/j.mce.2021.111189. ISSN 1872-8057. PMID 33539964. 
  40. Favre, Henri; Powell, Warren (2014). "Appendix 2". Nomenclature of Organic Chemistry - IUPAC Recommendations and Preferred Names 2013. The Royal Society of Chemistry. doi:10.1039/9781849733069. ISBN 978-0-85404-182-4. "oxy* –O– P-15.3.1.2.1.1; P-63.2.2.1.1" 
  41. Barrientos, Eduardo J.; Lapuerta, Magín; Boehman, André L. (2013-08). "Group additivity in soot formation for the example of C-5 oxygenated hydrocarbon fuels". Combustion and Flame 160 (8): 1484–1498. doi:10.1016/j.combustflame.2013.02.024. https://linkinghub.elsevier.com/retrieve/pii/S0010218013000849. 
  42. "IUPAC-IUB Joint Commission on Biochemical Nomenclature (JCBN). The nomenclature of steroids. Recommendations 1989". Eur J Biochem 186 (3): 429–58. 1989. doi:10.1111/j.1432-1033.1989.tb15228.x. PMID 2606099. "The prefix oxo- should also be used in connection with generic terms, e.g., 17-oxo steroids. The term ‘17-keto steroids’, often used in the medical literature, is incorrect because C-17 is specified twice, as the term keto denotes C=O" 
  43. "UMLS Metathesaurus - HGNC (HUGO Gene Nomenclature Committee) - Synopsis".
  44. Tweedie, Susan; Braschi, Bryony; Gray, Kristian; Jones, Tamsin E M; Seal, Ruth L; Yates, Bethan; Bruford, Elspeth A (2021). "Genenames.org: the HGNC and VGNC resources in 2021". Nucleic Acids Research 49 (D1): D939–D946. doi:10.1093/nar/gkaa980. ISSN 0305-1048. PMID 33152070. PMC 7779007. https://academic.oup.com/nar/article/49/D1/D939/5957168. 
  45. Swart, P.; Swart, A. C.; Waterman, M. R.; Estabrook, R. W.; Mason, J. I. (1993). "Progesterone 16 alpha-hydroxylase activity is catalyzed by human cytochrome P450 17 alpha-hydroxylase". J Clin Endocrinol Metab 77 (1): 98–102. doi:10.1210/jcem.77.1.8325965. PMID 8325965. 
  46. 46.0 46.1 46.2 Fukami, Maki; Homma, Keiko; Hasegawa, Tomonobu; Ogata, Tsutomu (2013). "Backdoor pathway for dihydrotestosterone biosynthesis: implications for normal and abnormal human sex development". Developmental Dynamics 242 (4): 320–9. doi:10.1002/dvdy.23892. PMID 23073980. 
  47. 47.0 47.1 47.2 47.3 47.4 47.5 Reisch, Nicole; Taylor, Angela E.; Nogueira, Edson F.; Asby, Daniel J.; Dhir, Vivek; Berry, Andrew; Krone, Nils; Auchus, Richard J. et al. (2019). "Alternative pathway androgen biosynthesis and human fetal female virilization". Proceedings of the National Academy of Sciences of the United States of America 116 (44): 22294–22299. doi:10.1073/pnas.1906623116. ISSN 1091-6490. PMID 31611378. PMC 6825302. //www.ncbi.nlm.nih.gov/pmc/articles/PMC6825302/. 
  48. 48.0 48.1 48.2 48.3 Flück, Christa E.; Meyer-Böni, Monika; Pandey, Amit V.; Kempná, Petra; Miller, Walter L.; Schoenle, Eugen J.; Biason-Lauber, Anna (2011). "Why boys will be boys: two pathways of fetal testicular androgen biosynthesis are needed for male sexual differentiation". American Journal of Human Genetics 89 (2): 201–218. doi:10.1016/j.ajhg.2011.06.009. ISSN 1537-6605. PMID 21802064. PMC 3155178. //www.ncbi.nlm.nih.gov/pmc/articles/PMC3155178/. 
  49. Biswas, Michael G.; Russell, David W. (1997). "Expression cloning and characterization of oxidative 17beta- and 3alpha-hydroxysteroid dehydrogenases from rat and human prostate". J Biol Chem 272 (25): 15959–66. doi:10.1074/jbc.272.25.15959. PMID 9188497. 
  50. Muthusamy, Selvaraj; Andersson, Stefan; Kim, Hyun-Jin; Butler, Ryan; Waage, Linda; Bergerheim, Ulf; Gustafsson, Jan-Åke (2011). "Estrogen receptor β and 17β-hydroxysteroid dehydrogenase type 6, a growth regulatory pathway that is lost in prostate cancer". Proc Natl Acad Sci U S A 108 (50): 20090–4. doi:10.1073/pnas.1117772108. PMID 22114194. PMC 3250130. //www.ncbi.nlm.nih.gov/pmc/articles/PMC3250130/. 
  51. Rižner, Tea Lanišnik; Lin, Hsueh K.; Penning, Trevor M. (2003). "Role of human type 3 3alpha-hydroxysteroid dehydrogenase (AKR1C2) in androgen metabolism of prostate cancer cells". Chem Biol Interact 143-144: 401–9. doi:10.1016/s0009-2797(02)00179-5. PMID 12604227. 
  52. Nishiyama, Tsutomu; Ishizaki, Fumio; Takizawa, Itsuhiro; Yamana, Kazutoshi; Hara, Noboru; Takahashi, Kota (2011). "5α-Androstane-3α 17β-diol Will Be a Potential Precursor of the Most Active Androgen 5α-Dihydrotestosterone in Prostate Cancer". Journal of Urology 185 (4S). doi:10.1016/j.juro.2011.02.378. 
  53. Snaterse, Gido; Mies, Rosinda; van Weerden, Wytske M.; French, Pim J.; Jonker, Johan W.; Houtsmuller, Adriaan B.; van Royen, Martin E.; Visser, Jenny A. et al. (2022-01-19). "Androgen receptor mutations modulate activation by 11-oxygenated androgens and glucocorticoids". Prostate Cancer and Prostatic Diseases. doi:10.1038/s41391-022-00491-z. ISSN 1476-5608. PMID 35046557. 
  54. Handelsman, David J.; Cooper, Elliot R.; Heather, Alison K. (April 2022). "Bioactivity of 11 keto and hydroxy androgens in yeast and mammalian host cells". The Journal of Steroid Biochemistry and Molecular Biology 218: 106049. doi:10.1016/j.jsbmb.2021.106049. ISSN 1879-1220. PMID 34990809. 
  55. "Unpublished results communicated by Karl-Heinz Storbeck in review". 21 October 2022.
  56. Campana, Carmela; Rege, Juilee; Turcu, Adina F.; Pezzi, Vincenzo; Gomez-Sanchez, Celso E.; Robins, Diane M.; Rainey, William E. (2016). "Development of a novel cell based androgen screening model". J Steroid Biochem Mol Biol 156: 17–22. doi:10.1016/j.jsbmb.2015.11.005. PMID 26581480. PMC 4748855. //www.ncbi.nlm.nih.gov/pmc/articles/PMC4748855/. 
  57. Haru, Shibusawa; Yumiko, Sano; Shoichi, Okinaga; Kiyoshi, Arai (1980). "Studies on 11β-hydroxylase of the human fetal adrenal gland". Journal of Steroid Biochemistry (Elsevier BV) 13 (8): 881–887. doi:10.1016/0022-4731(80)90161-2. ISSN 0022-4731. PMID 6970302. 
  58. Schloms, Lindie; Storbeck, Karl-Heinz; Swart, Pieter; Gelderblom, Wentzel C.A.; Swart, Amanda C. (2012). "The influence of Aspalathus linearis (Rooibos) and dihydrochalcones on adrenal steroidogenesis: quantification of steroid intermediates and end products in H295R cells". J Steroid Biochem Mol Biol 128 (3–5): 128–38. doi:10.1016/j.jsbmb.2011.11.003. PMID 22101210. 
  59. Swart, Amanda C.; Schloms, Lindie; Storbeck, Karl-Heinz; Bloem, Liezl M.; Toit, Therina du; Quanson, Jonathan L.; Rainey, William E.; Swart, Pieter (2013). "11β-hydroxyandrostenedione, the product of androstenedione metabolism in the adrenal, is metabolized in LNCaP cells by 5α-reductase yielding 11β-hydroxy-5α-androstanedione". J Steroid Biochem Mol Biol 138: 132–42. doi:10.1016/j.jsbmb.2013.04.010. PMID 23685396. 
  60. Barnard, Monique; Quanson, Jonathan L.; Mostaghel, Elahe; Pretorius, Elzette; Snoep, Jacky L.; Storbeck, Karl-Heinz (2018). "11-Oxygenated androgen precursors are the preferred substrates for aldo-keto reductase 1C3 (AKR1C3): Implications for castration resistant prostate cancer". J Steroid Biochem Mol Biol 183: 192–201. doi:10.1016/j.jsbmb.2018.06.013. PMID 29936123. PMC 6283102. //www.ncbi.nlm.nih.gov/pmc/articles/PMC6283102/. 
  61. Ando, Takashi; Nishiyama, Tsutomu; Takizawa, Itsuhiro; Ishizaki, Fumio; Miyashiro, Yoshimichi; Takeda, Keisuke; Hara, Noboru; Tomita, Yoshihiko (2016). "Dihydrotestosterone synthesis pathways from inactive androgen 5α-androstane-3β,17β-diol in prostate cancer cells: Inhibition of intratumoural 3β-hydroxysteroid dehydrogenase activities by abiraterone". Scientific Reports 6 (1): 32198. doi:10.1038/srep32198. ISSN 2045-2322. PMID 27561382. PMC 4999866. //www.ncbi.nlm.nih.gov/pmc/articles/PMC4999866/. 
  62. Asiedu, Bernice; Anang, Yvonne; Nyarko, Adraina; Doku, Derek Amartey; Amoah, Brodrick Y.; Santa, Sheila; Ngala, Robert A.; Asare, George A. (2017). "The role of sex steroid hormones in benign prostatic hyperplasia". The Aging Male 20 (1): 17–22. doi:10.1080/13685538.2016.1272101. ISSN 1368-5538. PMID 28084142. https://www.tandfonline.com/doi/full/10.1080/13685538.2016.1272101. 
  63. Caron, Patrick; Turcotte, Véronique; Guillemette, Chantal (2021). "A quantitative analysis of total and free 11-oxygenated androgens and its application to human serum and plasma specimens using liquid-chromatography tandem mass spectrometry". Journal of Chromatography A 1650: 462228. doi:10.1016/j.chroma.2021.462228. https://linkinghub.elsevier.com/retrieve/pii/S0021967321003526. 
  64. 64.0 64.1 Skiba, Marina A.; Bell, Robin J.; Islam, Rakibul M.; Handelsman, David J.; Desai, Reena; Davis, Susan R. (2019). "Androgens During the Reproductive Years: What Is Normal for Women?". J Clin Endocrinol Metab 104 (11): 5382–5392. doi:10.1210/jc.2019-01357. PMID 31390028. 
  65. Nanba, Aya T.; Rege, Juilee; Ren, Jianwei; Auchus, Richard J.; Rainey, William E.; Turcu, Adina F. (2019). "11-Oxygenated C19 Steroids Do Not Decline With Age in Women". J Clin Endocrinol Metab 104 (7): 2615–2622. doi:10.1210/jc.2018-02527. PMID 30753518. PMC 6525564. //www.ncbi.nlm.nih.gov/pmc/articles/PMC6525564/. 
  66. Davio, Angela; Woolcock, Helen; Nanba, Aya T.; Rege, Juilee; o'Day, Patrick; Ren, Jianwei; Zhao, Lili; Ebina, Hiroki et al. (2020). "Sex Differences in 11-Oxygenated Androgen Patterns Across Adulthood". J Clin Endocrinol Metab 105 (8): e2921–e2929. doi:10.1210/clinem/dgaa343. PMID 32498089. PMC 7340191. //www.ncbi.nlm.nih.gov/pmc/articles/PMC7340191/. 
  67. Nagasaki, Keisuke; Takase, Kaoru; Numakura, Chikahiko; Homma, Keiko; Hasegawa, Tomonobu; Fukami, Maki (2020). "Foetal virilisation caused by overproduction of non-aromatisable 11-oxy C19 steroids in maternal adrenal tumour". Human Reproduction 35 (11): 2609–2612. doi:10.1093/humrep/deaa221. PMID 32862221. 
  68. Barnard, Lise; Schiffer, Lina; Louw Du-Toit, Renate; Tamblyn, Jennifer A.; Chen, Shiuan; Africander, Donita; Arlt, Wiebke; Foster, Paul A. et al. (2021). "11-Oxygenated Estrogens Are a Novel Class of Human Estrogens but Do not Contribute to the Circulating Estrogen Pool". Endocrinology 162 (3). doi:10.1210/endocr/bqaa231. PMID 33340399. PMC 7814299. //www.ncbi.nlm.nih.gov/pmc/articles/PMC7814299/. 
  69. Mahesh, Virendra; Herrmann, Walter (1963). "Isolation of estrone and 11β-hydroxy estrone from a feminizing adrenal carcinoma". Steroids 1 (1): 51–61. doi:10.1016/S0039-128X(63)80157-9. ISSN 0039-128X. https://www.sciencedirect.com/science/article/pii/S0039128X63801579. 
  70. 70.0 70.1 Sumińska, Marta; Bogusz-Górna, Klaudia; Wegner, Dominika; Fichna, Marta (2020). "Non-Classic Disorder of Adrenal Steroidogenesis and Clinical Dilemmas in 21-Hydroxylase Deficiency Combined with Backdoor Androgen Pathway. Mini-Review and Case Report". Int J Mol Sci 21 (13): 4622. doi:10.3390/ijms21134622. PMID 32610579. PMC 7369945. //www.ncbi.nlm.nih.gov/pmc/articles/PMC7369945/. 
  71. Yildiz, Bulent O. (2006). "Diagnosis of hyperandrogenism: clinical criteria". Best Practice & Research. Clinical Endocrinology & Metabolism (Elsevier BV) 20 (2): 167–176. doi:10.1016/j.beem.2006.02.004. ISSN 1521-690X. PMID 16772149. 
  72. 72.0 72.1 Turcu, Adina F.; Auchus, Richard J. (2017). "Clinical significance of 11-oxygenated androgens". Curr Opin Endocrinol Diabetes Obes 24 (3): 252–259. doi:10.1097/MED.0000000000000334. PMID 28234803. PMC 5819755. //www.ncbi.nlm.nih.gov/pmc/articles/PMC5819755/. 
  73. Yang, Yabo; Ouyang, Nengyong; Ye, Yang; Hu, Qin; Du, Tao; Di, Na; Xu, Wenming; Azziz, Ricardo et al. (2020). "The predictive value of total testosterone alone for clinical hyperandrogenism in polycystic ovary syndrome". Reprod Biomed Online 41 (4): 734–742. doi:10.1016/j.rbmo.2020.07.013. PMID 32912651. 
  74. Balsamo, Antonio; Baronio, Federico; Ortolano, Rita; Menabo, Soara; Baldazzi, Lilia; Di Natale, Valeria; Vissani, Sofia; Cassio, Alessandra (2020). "Congenital Adrenal Hyperplasias Presenting in the Newborn and Young Infant". Frontiers in Pediatrics (Frontiers Media SA) 8: 593315. doi:10.3389/fped.2020.593315. ISSN 2296-2360. PMID 33415088. PMC 7783414. //www.ncbi.nlm.nih.gov/pmc/articles/PMC7783414/. 
  75. Kamrath, Clemens; Wettstaedt, Lisa; Boettcher, Claudia; Hartmann, Michaela F.; Wudy, Stefan A. (2018). "Androgen excess is due to elevated 11-oxygenated androgens in treated children with congenital adrenal hyperplasia". The Journal of Steroid Biochemistry and Molecular Biology (Elsevier BV) 178: 221–228. doi:10.1016/j.jsbmb.2017.12.016. ISSN 0960-0760. PMID 29277706. 
  76. Carmina, E.; Stanczyk, F. Z.; Chang, L.; Miles, R. A.; Lobo, R. A. (1992). "The ratio of androstenedione:11 beta-hydroxyandrostenedione is an important marker of adrenal androgen excess in women". Fertil Steril 58 (1): 148–52. doi:10.1016/s0015-0282(16)55152-8. PMID 1623996. 
  77. 77.0 77.1 Nowotny, Hanna F.; Braun, Leah; Vogel, Frederick; Bidlingmaier, Martin; Reincke, Martin; Tschaidse, Lea; Auer, Matthias K.; Lottspeich, Christian et al. (2022). "11-oxygenated C19 steroids are the predominant androgens responsible for hyperandrogenemia in Cushing's disease". Eur J Endocrinol. doi:10.1530/EJE-22-0320. PMID 36074938. 
  78. Lipsett, Mortimer B.; Riter, Barbara (1960). "Urinary ketosteroids and pregnanetriol in hirsutism". J Clin Endocrinol Metab 20 (2): 180–6. doi:10.1210/jcem-20-2-180. PMID 14417423. 
  79. Polson, D. W.; Reed, M. J.; Franks, S.; Scanlon, M. J.; James, V. H. T. (1988). "Serum 11 beta-hydroxyandrostenedione as an indicator of the source of excess androgen production in women with polycystic ovaries". J Clin Endocrinol Metab 66 (5): 946–50. doi:10.1210/jcem-66-5-946. PMID 3129451. 
  80. El-Maouche, Diala; Arlt, Wiebke; Merke, Deborah P. (2017). "Congenital adrenal hyperplasia". Lancet 390 (10108): 2194–2210. doi:10.1016/S0140-6736(17)31431-9. PMID 28576284. https://research.birmingham.ac.uk/portal/en/publications/congenital-adrenal-hyperplasia-the-lancet-seminar-series(d1a71692-d5a2-4ef3-b2d6-cef4e1318fbb).html. 
  81. Speiser, Phyllis W.; White, Perrin C. (2003). "Congenital adrenal hyperplasia". N Engl J Med 349 (8): 776–88. doi:10.1056/NEJMra021561. PMID 12930931. 
  82. Speiser, Phyllis W.; Arlt, Wiebke; Auchus, Richard J.; Baskin, Laurence S.; Conway, Gerard S.; Merke, Deborah P.; Meyer-Bahlburg, Heino F L.; Miller, Walter L. et al. (2018). "Congenital Adrenal Hyperplasia Due to Steroid 21-Hydroxylase Deficiency: An Endocrine Society Clinical Practice Guideline". The Journal of Clinical Endocrinology and Metabolism 103 (11): 4043–4088. doi:10.1210/jc.2018-01865. PMID 30272171. PMC 6456929. //www.ncbi.nlm.nih.gov/pmc/articles/PMC6456929/. 
  83. 83.0 83.1 Lee, Hyun Gyung; Kim, Chan Jong (2022). "Classic and backdoor pathways of androgen biosynthesis in human sexual development". Ann Pediatr Endocrinol Metab 27 (2): 83–89. doi:10.6065/apem.2244124.062. PMID 35793998. PMC 9260366. //www.ncbi.nlm.nih.gov/pmc/articles/PMC9260366/. 
  84. Turcu, Adina F.; Auchus, Richard J. (2015). "Adrenal Steroidogenesis and Congenital Adrenal Hyperplasia". Endocrinology and Metabolism Clinics of North America (Elsevier BV) 44 (2): 275–296. doi:10.1016/j.ecl.2015.02.002. ISSN 0889-8529. PMID 26038201. PMC 4506691703046. //www.ncbi.nlm.nih.gov/pmc/articles/PMC4506691703046/. 
  85. Turcu, A. F.; Rege, J.; Chomic, R.; Liu, J.; Nishimoto, H. K.; Else, T.; Moraitis, A. G.; Palapattu, G. S. et al. (2015). "Profiles of 21-Carbon Steroids in 21-hydroxylase Deficiency". The Journal of Clinical Endocrinology and Metabolism 100 (6): 2283–2290. doi:10.1210/jc.2015-1023. PMID 25850025. PMC 4454804. //www.ncbi.nlm.nih.gov/pmc/articles/PMC4454804/. 
  86. Nguyen, Lee S.; Rouas-Freiss, Nathalie; Funck-Brentano, Christian; Leban, Monique; Carosella, Edgardo D.; Touraine, Philippe; Varnous, Shaida; Bachelot, Anne et al. (2019). "Influence of hormones on the immunotolerogenic molecule HLA-G: a cross-sectional study in patients with congenital adrenal hyperplasia". Eur J Endocrinol 181 (5): 481–488. doi:10.1530/EJE-19-0379. PMID 31505456. 
  87. Kawarai, Yoshimasa; Ishikawa, Hiroshi; Segawa, Tomoya; Teramoto, Shokichi; Tanaka, Tomoaki; Shozu, Makio (2017). "High serum progesterone associated with infertility in a woman with nonclassic congenital adrenal hyperplasia". J Obstet Gynaecol Res 43 (5): 946–950. doi:10.1111/jog.13288. PMID 28188961. 
  88. Fiet, J.; Gueux, B.; Raux-DeMay, M. C.; Kuttenn, F.; Vexiau, P.; Brerault, J. L.; Couillin, P.; Galons, H. et al. (1989-03). "Increased plasma 21-deoxycorticosterone (21-DB) levels in late-onset adrenal 21-hydroxylase deficiency suggest a mild defect of the mineralocorticoid pathway". The Journal of Clinical Endocrinology and Metabolism 68 (3): 542–547. doi:10.1210/jcem-68-3-542. ISSN 0021-972X. PMID 2537337. https://pubmed.ncbi.nlm.nih.gov/2537337. 
  89. Fiet, J.; Gueux, B.; Raux-Demay, M. C.; Kuttenn, F.; Vexiau, P.; Gourmelen, M.; Couillin, P.; Mornet, E. et al. (1989-12-02). "[21-deoxycortisol. A new marker of virilizing adrenal hyperplasia caused by 21-hydroxylase deficiency"]. Presse Medicale (Paris, France: 1983) 18 (40): 1965–1969. ISSN 0755-4982. PMID 2531882. https://pubmed.ncbi.nlm.nih.gov/2531882. 
  90. Fiet, Jean; Le Bouc, Yves; Guéchot, Jérôme; Hélin, Nicolas; Maubert, Marie-Anne; Farabos, Dominique; Lamazière, Antonin (2017-03-01). "A Liquid Chromatography/Tandem Mass Spectometry Profile of 16 Serum Steroids, Including 21-Deoxycortisol and 21-Deoxycorticosterone, for Management of Congenital Adrenal Hyperplasia". Journal of the Endocrine Society 1 (3): 186–201. doi:10.1210/js.2016-1048. ISSN 2472-1972. PMID 29264476. PMC 5686660. https://pubmed.ncbi.nlm.nih.gov/29264476. 
  91. Turcu, Adina F.; Nanba, Aya T.; Chomic, Robert; Upadhyay, Sunil K.; Giordano, Thomas J.; Shields, James J.; Merke, Deborah P.; Rainey, William E. et al. (2016). "Adrenal-derived 11-oxygenated 19-carbon steroids are the dominant androgens in classic 21-hydroxylase deficiency". Eur J Endocrinol 174 (5): 601–9. doi:10.1530/EJE-15-1181. PMID 26865584. PMC 4874183. //www.ncbi.nlm.nih.gov/pmc/articles/PMC4874183/. 
  92. Auchus, Richard J.; Buschur, Elizabeth O.; Chang, Alice Y.; Hammer, Gary D.; Ramm, Carole; Madrigal, David; Wang, George; Gonzalez, Martha et al. (2014). "Abiraterone acetate to lower androgens in women with classic 21-hydroxylase deficiency". J Clin Endocrinol Metab 99 (8): 2763–70. doi:10.1210/jc.2014-1258. PMID 24780050. PMC 4121028. //www.ncbi.nlm.nih.gov/pmc/articles/PMC4121028/. 
  93. 93.0 93.1 Turcu, Adina F.; Mallappa, Ashwini; Nella, Aikaterini A.; Chen, Xuan; Zhao, Lili; Nanba, Aya T.; Byrd, James Brian; Auchus, Richard J. et al. (2021). "24-Hour Profiles of 11-Oxygenated C19 Steroids and Δ5-Steroid Sulfates during Oral and Continuous Subcutaneous Glucocorticoids in 21-Hydroxylase Deficiency". Front Endocrinol (Lausanne) 12: 751191. doi:10.3389/fendo.2021.751191. PMID 34867794. PMC 8636728. //www.ncbi.nlm.nih.gov/pmc/articles/PMC8636728/. 
  94. 94.0 94.1 94.2 Turcu, Adina F; Mallappa, Ashwini; Elman, Meredith S; Avila, Nilo A; Marko, Jamie; Rao, Hamsini; Tsodikov, Alexander; Auchus, Richard J et al. (2017). "11-Oxygenated Androgens Are Biomarkers of Adrenal Volume and Testicular Adrenal Rest Tumors in 21-Hydroxylase Deficiency". The Journal of Clinical Endocrinology and Metabolism 102 (8): 2701–2710. doi:10.1210/jc.2016-3989. PMID 28472487. PMC 5546849. //www.ncbi.nlm.nih.gov/pmc/articles/PMC5546849/. 
  95. Schröder, Mariska A M.; Turcu, Adina F.; o'Day, Patrick; Van Herwaarden, Antonius E.; Span, Paul N.; Auchus, Richard J.; Sweep, Fred C G J.; Claahsen-Van Der Grinten, Hedi L. (2022). "Production of 11-Oxygenated Androgens by Testicular Adrenal Rest Tumors". J Clin Endocrinol Metab 107 (1): e272–e280. doi:10.1210/clinem/dgab598. PMID 34390337. PMC 8684463. //www.ncbi.nlm.nih.gov/pmc/articles/PMC8684463/. 
  96. Miller, Walter L.; Auchus, Richard J. (2019). "The "backdoor pathway" of androgen synthesis in human male sexual development". PLOS Biology 17 (4): e3000198. doi:10.1371/journal.pbio.3000198. PMID 30943210. PMC 6464227. //www.ncbi.nlm.nih.gov/pmc/articles/PMC6464227/. 
  97. Flück, Christa E.; Pandey, Amit V. (2014). "Steroidogenesis of the testis -- new genes and pathways". Ann Endocrinol (Paris) 75 (2): 40–7. doi:10.1016/j.ando.2014.03.002. PMID 24793988. 
  98. Zachmann, M. (1996). "Prismatic cases: 17,20-desmolase (17,20-lyase) deficiency". J Clin Endocrinol Metab 81 (2): 457–9. doi:10.1210/jcem.81.2.8636249. PMID 8636249. 
  99. 99.0 99.1 Boettcher, Claudia; Flück, Christa E. (2022). "Rare forms of genetic steroidogenic defects affecting the gonads and adrenals". Best Pract Res Clin Endocrinol Metab 36 (1): 101593. doi:10.1016/j.beem.2021.101593. PMID 34711511. 
  100. 100.0 100.1 Du Toit, Therina; Swart, Amanda C. (2021). "Turning the spotlight on the C11-oxy androgens in human fetal development". J Steroid Biochem Mol Biol 212: 105946. doi:10.1016/j.jsbmb.2021.105946. PMID 34171490. 
  101. Finkielstain, Gabriela P.; Vieites, Ana; Bergadá, Ignacio; Rey, Rodolfo A. (2021). "Disorders of Sex Development of Adrenal Origin". Front Endocrinol (Lausanne) 12: 770782. doi:10.3389/fendo.2021.770782. PMID 34987475. PMC 8720965. //www.ncbi.nlm.nih.gov/pmc/articles/PMC8720965/. 
  102. Dumesic, Daniel A.; Oberfield, Sharon E.; Stener-Victorin, Elisabet; Marshall, John C.; Laven, Joop S.; Legro, Richard S. (2015-10-01). "Scientific Statement on the Diagnostic Criteria, Epidemiology, Pathophysiology, and Molecular Genetics of Polycystic Ovary Syndrome". Endocrine Reviews 36 (5): 487–525. doi:10.1210/er.2015-1018. ISSN 0163-769X. PMID 26426951. PMC PMC4591526. https://academic.oup.com/edrv/article/36/5/487/2354671. 
  103. Stewart, P. M.; Shackleton, C. H.; Beastall, G. H.; Edwards, C. R. (1990). "5 alpha-reductase activity in polycystic ovary syndrome". Lancet (London, England) 335 (8687): 431–433. doi:10.1016/0140-6736(90)90664-q. ISSN 0140-6736. PMID 1968168. https://pubmed.ncbi.nlm.nih.gov/1968168. 
  104. Vassiliadi, Dimitra A.; Barber, Thomas M.; Hughes, Beverly A.; McCarthy, Mark I.; Wass, John A. H.; Franks, Stephen; Nightingale, Peter; Tomlinson, Jeremy W. et al. (2009). "Increased 5 alpha-reductase activity and adrenocortical drive in women with polycystic ovary syndrome". J Clin Endocrinol Metab 94 (9): 3558–66. doi:10.1210/jc.2009-0837. PMID 19567518. 
  105. Marti, Nesa; Galván, José A.; Pandey, Amit V.; Trippel, Mafalda; Tapia, Coya; Müller, Michel; Perren, Aurel; Flück, Christa E. (2017-02). "Genes and proteins of the alternative steroid backdoor pathway for dihydrotestosterone synthesis are expressed in the human ovary and seem enhanced in the polycystic ovary syndrome". Molecular and Cellular Endocrinology 441: 116–123. doi:10.1016/j.mce.2016.07.029. https://linkinghub.elsevier.com/retrieve/pii/S0303720716302696. 
  106. Janner, Marco; Sommer, Grit; Groessl, Michael; Flück, Christa E (2020). "Premature Adrenarche in Girls Characterized By Enhanced 17,20-Lyase and 17β-Hydroxysteroid Dehydrogenase Activities". The Journal of Clinical Endocrinology & Metabolism 105 (12): e4439–e4451. doi:10.1210/clinem/dgaa598. ISSN 0021-972X. PMID 32865200. https://academic.oup.com/jcem/article/105/12/e4439/5899560. 
  107. Rege, Juilee; Turcu, Adina; Kasa-Vubu, Josephine Z; Lerario, Antonio M; Auchus, Gabriela C; Auchus, Richard J; Smith, Joshua M; White, Perrin C et al. (2018). "11-ketotestosterone is the dominant circulating bioactive androgen during normal and premature adrenarche". The Journal of Clinical Endocrinology & Metabolism. doi:10.1210/jc.2018-00736. ISSN 0021-972X. PMID 30137510. PMC 6226603. //www.ncbi.nlm.nih.gov/pmc/articles/PMC6226603/. 
  108. Du Toit, Therina; Swart, Amanda C. (2018). "Inefficient UGT-conjugation of adrenal 11β-hydroxyandrostenedione metabolites highlights C11-oxy C19 steroids as the predominant androgens in prostate cancer". Mol Cell Endocrinol 461: 265–276. doi:10.1016/j.mce.2017.09.026. PMID 28939401. 
  109. Ventura-Bahena, Arianna; Hernández-Pérez, Jesús Gibran; Torres-Sánchez, Luisa; Sierra-Santoyo, Adolfo; Escobar-Wilches, Derly Constanza; Escamilla-Núñez, Consuelo; Gómez, Rocio; Rodríguez-Covarrubias, Francisco et al. (2021). "Urinary androgens excretion patterns and prostate cancer in Mexican men". Endocr Relat Cancer 28 (12): 745–756. doi:10.1530/ERC-21-0160. PMID 34520388. 
  110. Snaterse, G.; Van Dessel, L. F.; Van Riet, J.; Taylor, A. E.; Van Der Vlugt-Daane, M.; Hamberg, P.; De Wit, R.; Visser, J. A. et al. (2021). "11-Ketotestosterone is the predominant active androgen in prostate cancer patients after castration". JCI Insight 6 (11). doi:10.1172/jci.insight.148507. PMID 33974560. PMC 8262344. //www.ncbi.nlm.nih.gov/pmc/articles/PMC8262344/. 
  111. 111.0 111.1 Dimitrakov, Jordan; Joffe, Hylton V.; Soldin, Steven J.; Bolus, Roger; Buffington, C.A. Tony; Nickel, J. Curtis (2008). "Adrenocortical hormone abnormalities in men with chronic prostatitis/chronic pelvic pain syndrome". Urology 71 (2): 261–6. doi:10.1016/j.urology.2007.09.025. PMID 18308097. PMC 2390769. //www.ncbi.nlm.nih.gov/pmc/articles/PMC2390769/. 
  112. Du Toit, Therina; Swart, Amanda C. (2020). "The 11β-hydroxyandrostenedione pathway and C11-oxy C21 backdoor pathway are active in benign prostatic hyperplasia yielding 11keto-testosterone and 11keto-progesterone". The Journal of Steroid Biochemistry and Molecular Biology 196: 105497. doi:10.1016/j.jsbmb.2019.105497. PMID 31626910. 
  113. Carmina, Enrico; Dewailly, Didier; Escobar-Morreale, Héctor F.; Kelestimur, Fahrettin; Moran, Carlos; Oberfield, Sharon; Witchel, Selma F.; Azziz, Ricardo (2017). "Non-classic congenital adrenal hyperplasia due to 21-hydroxylase deficiency revisited: an update with a special focus on adolescent and adult women". Hum Reprod Update 23 (5): 580–599. doi:10.1093/humupd/dmx014. PMID 28582566. 
  114. Witchel, Selma Feldman; Azziz, Ricardo (2010). "Nonclassic congenital adrenal hyperplasia". Int J Pediatr Endocrinol 2010: 625105. doi:10.1155/2010/625105. PMID 20671993. PMC 2910408. //www.ncbi.nlm.nih.gov/pmc/articles/PMC2910408/. 
  115. Masiutin, Maxim G.; Yadav, Maneesh K. (2022). "Letter to the editor regarding the article "Adrenocortical hormone abnormalities in men with chronic prostatitis/chronic pelvic pain syndrome"". Urology. doi:10.1016/j.urology.2022.07.051. ISSN 0090-4295. PMID 35987379. https://www.goldjournal.net/article/S0090-4295(22)00690-2/abstract. 
  116. Dimitrakoff, Jordan; Nickel, J. Curtis (2022). "AUTHOR REPLY". Urology. doi:10.1016/j.urology.2022.07.049. ISSN 0090-4295. PMID 35985522. https://www.goldjournal.net/article/S0090-4295(22)00691-4/abstract. 
  117. Miller, Walter L. (2019). "Congenital Adrenal Hyperplasia: Time to Replace 17OHP with 21-Deoxycortisol". Hormone Research in Paediatrics 91 (6): 416–420. doi:10.1159/000501396. ISSN 1663-2826. PMID 31450227. https://pubmed.ncbi.nlm.nih.gov/31450227. 
  118. Frederiksen, D. W.; Wilson, J. D. (1971). "Partial characterization of the nuclear reduced nicotinamide adenine dinucleotide phosphate: delta 4-3-ketosteroid 5 alpha-oxidoreductase of rat prostate". J Biol Chem 246 (8): 2584–93. doi:10.1016/S0021-9258(18)62328-2. PMID 4396507. 
  119. Ellsworth, K.; Harris, G. (1995). "Expression of the type 1 and 2 steroid 5 alpha-reductases in human fetal tissues". Biochem Biophys Res Commun 215 (2): 774–80. doi:10.1006/bbrc.1995.2530. PMID 7488021. 
  120. Lunacek, A.; Schwentner, C.; Oswald, J.; Fritsch, H.; Sergi, C.; Thomas, L.N.; Rittmaster, R.S.; Klocker, H. et al. (2007). "Fetal distribution of 5alpha-reductase 1 and 5alpha-reductase 2, and their input on human prostate development". J Urol 178 (2): 716–21. doi:10.1016/j.juro.2007.03.089. PMID 17574609. 
  121. Kovats, Susan; Carreras, Esther; Agrawal, Hemant (2010). Klein, Sabra L.. ed. Sex Steroid Receptors in Immune Cells (in en). Berlin, Heidelberg: Springer Berlin Heidelberg. pp. 53–91. doi:10.1007/978-3-642-02155-8_3. ISBN 978-3-642-02154-1. http://link.springer.com/10.1007/978-3-642-02155-8_3. 
  122. Sellau, Julie; Groneberg, Marie; Lotter, Hannelore (2019). "Androgen-dependent immune modulation in parasitic infection". Seminars in Immunopathology 41 (2): 213–224. doi:10.1007/s00281-018-0722-9. ISSN 1863-2297. http://link.springer.com/10.1007/s00281-018-0722-9. 
  123. Trigunaite, Abhishek; Dimo, Joana; Jørgensen, Trine N. (2015-04). "Suppressive effects of androgens on the immune system". Cellular Immunology 294 (2): 87–94. doi:10.1016/j.cellimm.2015.02.004. https://linkinghub.elsevier.com/retrieve/pii/S0008874915000313. 
  124. Moradi, Fatemeh; Enjezab, Behnaz; Ghadiri-Anari, Akram (2020). "The role of androgens in COVID-19". Diabetes & Metabolic Syndrome: Clinical Research & Reviews 14 (6): 2003–2006. doi:10.1016/j.dsx.2020.10.014. ISSN 1871-4021. https://www.sciencedirect.com/science/article/pii/S1871402120304021. 
  125. Wambier, Carlos Gustavo; Goren, Andy; Vaño‐Galván, Sergio; Ramos, Paulo Müller; Ossimetha, Angelina; Nau, Gerard; Herrera, Sabina; McCoy, John (2020). "Androgen sensitivity gateway to COVID ‐19 disease severity". Drug Development Research 81 (7): 771–776. doi:10.1002/ddr.21688. ISSN 0272-4391. PMID 32412125. PMC PMC7273095. https://onlinelibrary.wiley.com/doi/10.1002/ddr.21688. 
  126. Meng, Jialin; Ge, Qintao; Li, Jiawei; Lu, Xiaofan; Chen, Yonghao; Wang, Haitao; Zhang, Meng; Du, Juan et al. (2022). "Protective trend of anti-androgen therapy during the COVID-19 pandemic: A meta-analysis". Journal of Infection 84 (6): 834–872. doi:10.1016/j.jinf.2022.03.020. ISSN 0163-4453. PMID 35341831. https://www.journalofinfection.com/article/S0163-4453(22)00175-X/abstract. 
  127. Schiffer, Lina; Bossey, Alicia; Kempegowda, Punith; Taylor, Angela E.; Akerman, Ildem; Scheel-Toellner, Dagmar; Storbeck, Karl-Heinz; Arlt, Wiebke (2021). "Peripheral blood mononuclear cells preferentially activate 11-oxygenated androgens". European Journal of Endocrinology 184 (3): 353–363. doi:10.1530/EJE-20-1077. ISSN 0804-4643. https://eje.bioscientifica.com/view/journals/eje/184/3/EJE-20-1077.xml.