Download as pdf or txt
Download as pdf or txt
You are on page 1of 17

Seminars in Cell & Developmental Biology 59 (2016) 10–26

Contents lists available at ScienceDirect

Seminars in Cell & Developmental Biology


journal homepage: www.elsevier.com/locate/semcdb

Review

Spermatogenesis in humans and its affecting factors


Filipe Tenorio Lira Neto, Phil Vu Bach, Bobby B. Najari, Philip S. Li, Marc Goldstein ∗
Center for Male Reproductive Medicine and Microsurgery, Department of Urology and Institute of Reproductive Medicine, Weill Cornell Medical College,
New York, NY, United States

a r t i c l e i n f o a b s t r a c t

Article history: Spermatogenesis is an extraordinary complex process. The differentiation of spermatogonia into sperma-
Received 11 February 2016 tozoa requires the participation of several cell types, hormones, paracrine factors, genes and epigenetic
Received in revised form 13 April 2016 regulators. Recent researches in animals and humans have furthered our understanding of the male
Accepted 15 April 2016
gamete differentiation, and led to clinical tools for the better management of male infertility. There is still
Available online 30 April 2016
much to be learned about this intricate process. In this review, the critical steps of human spermatogenesis
are discussed together with its main affecting factors.
Keywords:
© 2016 Elsevier Ltd. All rights reserved.
Human spermatogenesis
Sperm
Reproduction
Male infertility
Genetics
Testis
Testosterone
Varicocele
Germ cells

Contents

1. Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 11
2. Structures and cells . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 11
2.1. Seminiferous tubules . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 11
2.2. Peritubular myoid cells . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 11
2.3. Leydig cells . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 11
2.4. Sertoli cells . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 12
2.5. Germ cells . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 13
2.6. Blood testis barrier . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 15
3. Factors that directly affect human spermatogenesis . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 16
3.1. Obesity . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 16
3.2. Diabetes . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 16
3.3. Environmental chemicals . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 17
3.4. Varicocele . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 17
3.5. Genetic factors . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 17
3.5.1. Klinefelter syndrome . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 17
3.5.2. Y-chromosome microdeletions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 18
3.5.3. Micro RNAs . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 18
3.5.4. Epigenetic factors . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 19
4. Conclusion . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 20
References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 20

∗ Corresponding author.
E-mail address: mgoldst@med.cornell.edu (M. Goldstein).

http://dx.doi.org/10.1016/j.semcdb.2016.04.009
1084-9521/© 2016 Elsevier Ltd. All rights reserved.
F.T.L. Neto et al. / Seminars in Cell & Developmental Biology 59 (2016) 10–26 11

1. Introduction angiotensin II acting via type 1 angiotensin receptor (ATR1) and


oxytocin [10,16,17]. TNF-a strongly upregulates the production of
Spermatogenesis encompasses a complex network of processes proinflammatory interleukins by PTCs and might be related to the
that occur in the seminiferous tubules (STs) and culminates in the peritubular wall remodeling observed in some infertility patients
production of the mature male gamete. The processes are: pro- [16,18,19].
liferation of spermatogonia; spermatogonial differentiation into
spermatocytes; meiotic division of spermatocytes producing sper- 2.3. Leydig cells
matids; maturation of round spermatids; and the release of highly
specialized mature spermatozoa into the testicular tubule lumen Leydig cells (LCs) are located in clusters usually found in strate-
[1]. gic positions between blood vessels and ST [20,21]. They are
The entire spermatogenic process is thought to require approx- polygonal in shape and exhibit ultrastructural features that are con-
imately 74 days [2,3], but a more recent study in normal men sistent with their function as the primary source of testosterone (T)
concluded that the total time to produce ejaculated sperm might in males.
vary between 42–76 days [4]. The estimated daily sperm produc- LCs are thought to differentiate from fibroblast-like or mes-
tion per man ranges from 150 to 275 million spermatozoa [3,5]. enchymal cells in the testis interstitium [22]. In humans, their
Several testicular structures and cells play important roles dur- differentiation occurs in three waves, matching the also triphasic
ing spermatogenesis, while an ample array of factors can influence pattern of T production during development [23]. As a downstream
its quality and quantity. In this article, we review the current liter- event of sex-determining region (SRY) gene signaling, the first wave
ature regarding the main actors of human spermatogenesis. occurs between 8 and 18 weeks of gestation and is responsible for
the male secondary sexual differentiation [24].
The second wave occurs in the first 2–3 months after birth in
2. Structures and cells
response to a concurrent LH surge. This wave is responsible for the
hormonal imprinting of hypothalamus, liver, and prostate [1,25].
2.1. Seminiferous tubules
The third wave is triggered by the maturation of the hypothalamic
pituitary gonadal (HPG) axis during puberty. Originated from pre-
The average volume of an adult human testis is 30 mL and its
cursors cells and regressed neonatal LCs, adult LCs persist active
parenchyma is divided in 200–300 lobules by septations arising
throughout adulthood [23,26].
from the tunica albuginea. Each of these lobules contains one to
Testosterone is the main LCs product, and the testes secret
three loops of STs, each one measuring 70–80 cm in length when
between 3 and 10 mg/day of T, accounting for more than 95% of
stretched, and with a total length of 250 m per testis [1,6]. The STs
total circulating T in post-pubertal men. Testosterone is synthe-
are the functional unit of the testis and occupy two-thirds of the
sized from cholesterol in a complex process that involves multiple
organ’s volume. They consist of a basement membrane, Sertoli cells
cytoplasmic and mitochondrial enzymes and transporters [27,28].
(SCs), and germ cells at varying stages of maturation. So far, 13
The rate-determining step is the cholesterol transport from the
types of human germ cells (GCs) have been identified. Classified in
outer mitochondrial membrane to the inner mitochondrial mem-
ascending order of maturation, they are the dark type A spermato-
brane by two particular molecules, TSPO (translocator protein),
gonia, the pale type A spermatogonia, the type B spermatogonia, the
which acts as a channel [29], and the steroidogenic acute regu-
preleptotene, leptotene, zygotene, and pachytene primary sperma-
latory protein (StAR) whose mechanism of action is still unknown
tocytes, the secondary spermatocytes, and the Sa, Sb, Sc, Sd1, and
[30–32].
Sd2 spermatids [7].
Testosterone produced by LCs may be stored in the intrat-
A peritubular tissue surrounds the ST and is composed of layers
esticular compartment or released into the systemic circulation.
of myoid cells, fibrocyte-like adventitial cells, and collagen matrix.
In humans, intratesticular testosterone (ITT) levels are 100-fold
Together with SCs, this peritubular tissue forms the blood testis
higher than serum levels [33,34]. This may be due to local T pro-
barrier (BTB) [8].
duction by Leydig cells combined with a countercurrent exchange
mechanism in the pampiniform plexus [35]. Clinical studies in men
2.2. Peritubular myoid cells using a T-based contraceptive regimen revealed that high ITT lev-
els are necessary to support normal spermatogenesis [36] and that
Peritubular myoid cells (PTCs) are large, flat mesenchymal cells administration of hCG was able to bring ITT back to baseline levels
with some fibroblast-like and smooth muscle cell-like characteris- in those men [37]. In addition, ITT optimization through the use
tics [8]. They are arranged in discontinuous cell layers, providing of exogenous gonadotropins resulted in improved spermatogen-
support for the STs. One of their main functions is the propulsion of esis in men with non-obstructive azoospermia (NOA) [38]. There
the testicular fluid containing immotile spermatozoa towards the is no explanation for the need for such high levels of ITT in order
rete testis, using contractile filaments of actin and myosin [9,10]. to maintaining normal sperm production. In fact, much of the ITT
PTCs are also implicated in paracrine regulation of SCs functions. might not be available to bind AR, since it is bound to androgen-
PModS (peritubular modifies Sertoli) is a PTC-derived molecule binding protein, therefore, part of the ITT content might function
that modulates the secretion of transferrin, inhibin and androgen- just as a T reserve [39].
binding protein by SCs. Other factors produced by PTCs that LCs are also the predominant testicular source of estrogens after
regulate SCs function are IGF-1, bFGF and a number of interleukins. puberty [40]. Estradiol is the main testicular estrogen and is metab-
PTCs also take part in the production and maintenance of the blood olized from T by the microsomal P450 aromatase. This enzyme is
testis barrier trough the secretion of fibronectin, collagens, proteo- responsible for maintaining the testosterone/estradiol ratio, and its
glycans and entactin [11,12]. activity is altered in some men with spermatogenesis failure [41].
Several factors regulate PTCs activities. Studies using mice with Even though aromatase is mainly localized in LCs, it can be also
the androgen receptor (AR) gene deleted in PTCs showed that found in SCs and GCs [42], the latter may be an estrogen source
androgens are important modulators of PTCs functions [13,14]. In as important as LCs [43]. This widespread expression of aromatase
humans, AR are also found in PTCs, but to a lesser extent than in highlights the importance of estrogens for spermatogenesis. Very
SCs, and the clinical significance of androgen signaling in human high levels of intratesticular estrogens found in humans give even
PTCs is still unclear [15]. PTCs contractile function is regulated by more support to this idea [44].
12 F.T.L. Neto et al. / Seminars in Cell & Developmental Biology 59 (2016) 10–26

Other factors synthesized by LCs have been shown to influence Androgen-binding protein (ABP) is another secretory product of
human testicular function. Insulin-like factor 3 (INSL3) is a pep- SCs. It is a carrier protein that binds to T and dihydrotestosterone
tide produced almost exclusively by the LCs [45,46]. Acting via (DHT) with high affinity. ABP may regulate the bioavailability of
the relaxin-family peptide receptor 2 (RXFP2), INSL3 mediates the androgens in the extracellular space of the testis and epididymis,
migration of the testis to the scrotum [47,48]. Its role in the adult maintaining a high concentration of these molecules in these
human testis remains unknown, but a paracrine action on GCs is organs. Others advocate that ABP also controls spermatogenesis
plausible. Importantly, INSL3 concentrations may reflect the func- and sperm maturation, protecting androgens from metabolism
tional status of LCS, possibly better than T concentrations [49,50]. and facilitating androgen uptake in the male reproductive tract
Oxytocin is also produced by human LCs and acts on the PTCs, [80–83]. Transferrin, vitamin transporters, lactate, acetate, extra-
increasing STs contractions. Oxytocin may also act in an autocrine cellular matrix components, glial cell-derived neurotrophic factor
way, stimulating steroidogenesis [17,51]. (GNDF), TGF-a, TGF-b and interleukins are other SC-derived prod-
The steroidogenic activity of LCs is driven by a series of hor- ucts that seem to contribute to the microenvironment in which the
monal and non-hormonal factors. Luteinizing hormone (LH) is the GCs maturate [84–93].
most important of them and acts by binding to high-affinity LH/HCG Also produced by SCs, inhibin B is a glycoprotein hormone that
receptors on the plasma membrane. The binding of LH to LH/HCG regulates the production and secretion of FSH in the anterior pitu-
receptors starts a chain of acute and trophic events. The acute itary gland, in a classic negative feedback loop [94]. Inhibin B is
events are mediated by coupled G protein – cAMP mechanism composed of two subunits, ␣ and ␤. The ␣ subunit is located pre-
and include the synthesis of StAR and cytochrome P450 enzymes, dominantly in the SCs, while the ␤ subunit is located mainly in
leading to an immediate increase in the T and estradiol produc- the GCs, suggesting that an interaction between SCs and GCs is
tion within 24–72 h [52,53]. In contrast, the trophic effect is the necessary for inhibin B production, at least after puberty [95–97].
inhibition of LCs apoptosis mediated through the ERK1/2 and Akt Furthermore, inhibin B may be a paracrine and autocrine regulator
pathways [54,55]. In addition to the hypothalamic/pituitary nega- of LCs and SCs functions [98–100].
tive feedback, testosterone also regulates LCs function through an Sertoli cell also act as macrophages, using phagocytosis to clean
ultrashort loop negative feedback [56]. any degenerating GCs or residual bodies from spermatids. This is
Other factors, such as gonadotropin-releasing hormone, mela- a critical function, because a considerable proportion of GCs are
tonin, epidermal growth factor, insulin-like growth factor 1, atrial discarded during spermatogenesis, and the presence of these dead
natriuretic peptide, ghrelin, and neurotransmitters, such as the cells into the ST lumen could lead to the release of noxious contents
gamma-aminobutyric acid, have also been reported to affect that negatively impact sperm production [101–103].
human LCs activity, but with less robust data [57–65]. Regulation of SCs activities is achieved by hormonal, non-
hormonal and paracrine factors. The most important regulator of
SCs is the follicle-stimulating hormone (FSH). Produced by the
2.4. Sertoli cells pituitary gland under the stimulus of gonadotropin-releasing hor-
mone (GnRH), FSH mediates the connection between the brain and
Sertoli cells are considered the indispensable conductors of SCs. Acting via specific G-coupled receptors, FSH activates several
spermatogenesis and occupy 17–20% of the STs epithelium in an pathways, stimulating virtually all functions related to spermato-
adult man [66]. They exhibit irregular shape and polarized align- genesis. Despite the important role played by FSH, the hormone
ment, with the base resting over the basal membrane and the apex is not considered essential to spermatogenesis. Clinical data from
pointing to the STs lumen [67,68]. Another distinguished charac- men lacking FSH production or with FSH receptor mutation have
teristic is the prominent nucleoli, which are readily identified in revealed that spermatogenesis is heavily affected but not extin-
testis biopsy slides. guished in these cases; a finding that could be explained in part by
Their morphology is constantly changing, with vast cytoplas- the discovery that the FSH receptor has some low level of constitu-
mic ramifications resembling tree branches, which involve GCs tive activity [104–107].
and assist with their migration towards the tubule lumen dur- In contrast, androgen signaling is indispensable to normal sper-
ing spermatogenesis. A single SC can support up to 30–50 GCs at matogenesis, especially for meiotic progression and spermatid
different stages of development [69,70]. The framework for this maturation, and SCs are the main mediator of androgen action
intricate structure is provided by a complex cytoskeleton composed in spermatogenesis [15]. The AR is found in SCs and moves to
of actin filaments, tubulobulbar complexes, ectoplasmic specializa- the nucleus to stimulate transcription when combined to its lig-
tions, intermediate filaments and microtubules, each one with its ands (T or dihydrotestosterone). Unfortunately, the exact genetic
own ultrastructure and function. mechanism by which androgens influence spermatogenesis is still
Sertoli cells play a central role in the embryology of the testis. unclear because few consistent target genes have been identified
Until the sixth week of gestation, the human primordial gonad is [108–111]. Therefore, secondary signaling pathways, such as Ca2+-
fully bipotential, and the differentiation of SCs from progenitor cells dependent and MAP kinase pathways, have been proposed, but
of mesenchymal origin is the mark of male gonadal development there is also paucity of data about them [105,112,113].
[71,72]. The presence of the Y-chromosome, or, more specifically, Insulin is another factor that modulates SCs activity, specifically,
its SRY gene directs this unfinished organ towards testicular differ- the carbohydrate metabolism (i.e. production of lactate). Insulin-
entiation [73]. The product of SRY gene is a DNA-binding protein deprived SCs were shown to have decreased lactate production,
that upregulates a complex network of transcription factors, whose and, since lactate is the main energy source for some GCs, this effect
net effect is the differentiation of SCs. Other testicular cell types could explain part of the deleterious influence of type I diabetes
do not express SRY, and their male fate is directed by molecular over spermatogenesis. Other substances, such as DHT, estradiol,
signals arising from SCs [74,75]. Shortly after differentiation, SCs melatonin, bFGF and interleukins, have been implicated in the reg-
start to produce anti-Müllerian hormone (AMH), which induces ulation of SCs metabolism, but the their clinical significance is
regression of the Müllerian ducts and precludes the development still unknown [87,89,114,115]. As described earlier, PModS is a
of female internal genitalia [76,77]. SCs continue to secrete AMH molecule secreted by PTCs that modulates the secretion of trans-
throughout development and into adulthood, but the function of ferrin, inhibin B and androgen-binding protein by SCs. It has been
AMH after testicular differentiation remains unclear, with some
authors suggesting a role in non-gonadal development [78,79].
F.T.L. Neto et al. / Seminars in Cell & Developmental Biology 59 (2016) 10–26 13

Fig. 1. Hormonal and Paracrine Control of Spermatogenesis.

shown to have more effect on SCs function than any other known In the most accepted model, the stem GCs are the A dark and
regulatory factor, including FSH [12,116] (Fig. 1). A pale SPG. The A dark SPG are the quiescent or reserve stem GCs,
while the A pale SPG are the active stem GCs. The A pale SPG pro-
liferate to either self-renew or to produce B SPG, which go through
2.5. Germ cells one mitotic division before initiating meiosis [132–134]. There is
some debate about the pattern of A pale SPG division, but, using any
Germ cells comprise a family of cells whose sole purpose is of the current models, 8 preleptotene spermatocytes are produced
to become spermatozoa, and thus, to transmit genetic and epi- from the original pair of A pale SPG [132,135].
genetic information across generations. They are the only human The decision whether the SPG will self-renew, differentiate or
cell type capable of meiosis. Located within the STs, GCs are dis- become apoptotic is mainly influenced by SC-derived factors. GNDF
tributed in a highly organized manner, with less matured cells has been demonstrated to promote spermatogonial self-renew.
occupying the basal compartment and progressing to the adlumi- In contrast, SCF, bone morphogenetic protein 4, retinoic acid and
nal compartment as they mature [3,7]. The primordial GCs arise Notch1/Jagged2 signaling system induce differentiation. Several
from extra-embryonic tissues surrounding the yolk sac. Between micro-RNAs are also implicated in the modulation of spermatogo-
3 and 5 weeks of development, they migrate to the gonadal ridge, nial fate [93,136–138]. Apoptosis is the mechanism used by SCs
where, directed by SCs factors, they differentiate into gonocytes. to control the number of GCs, keeping the proportion SCs/GCs
The gonocytes enter in arrest in the G0 phase and stay mitoti- constant. In fact, around 50–70% of developing GCs are discarded
cally inactive until after birth. Factors, such as the stem cell factor during spermatogenesis. The signal to GCs to enter into apoptosis is
(SCF), stromal cell derived factor 1 (SDF-1), activator protein- given by SCs trough the FaS/FaSL system, and modulated by Bcl-2,
2␥ (AP-2␥), growth and differentiation factor 3 (GDF3), estradiol, Bax and TNF-␣ [139–142]. The transcription factor nuclear factor
retinoic acid, and methylation of DNA, appear to have an important ␬B (NF-kB) is likewise important in this process, but instead of act-
role during the primordial GCs and gonocyte differentiations, and ing directly on GCs, it induces transcription of SCs factors involved
may be the source of some infertility causes and testicular tumors in the regulation of GCs death. Given their importance, any distur-
[117–127]. Between birth and 6 months, the gonocytes will further bance in theses mechanisms could affect the male fertility potential
differentiate into spermatogonia (SPG), which stay quiescent until [143,144].
the age of 5–7 years, when they increase in number via mitosis. Germ cells develop in consistent groupings called cellular asso-
Starting with puberty, and in parallel with the proliferation pro- ciations. In humans, a system with 6 cellular associations, based
cess, SPG begin the differentiation process towards spermatozoa on the morphological changes of the cellular nucleus after stain-
[128,129]. ing with hematoxylin-eosin, has been used since 1963 [7]. Each
Spermatogonia are the diploid progenitors of all other GCs types, cellular association is considered a stage in the spermatogenic
and have the dual responsibility of undergoing meiosis to produce cycle. Men with low sperm production might show atypical cellular
the male gamete, and mitosis to self-renew, maintaining the con- associations, due to missing GCs or intermingling of associations
tinuous production of spermatozoa throughout man’s life. Located [3,5,145]. Recently, a novel 12-stage cycle system was proposed
in the basal compartment of the STs, and in close contact with SCs, using changes in shape and size of the acrosome made visible by
they have an ovoid nucleus and a dense cytoplasm containing a immunohistochemical detection of the acrosin protein [146]. This
small Golgi apparatus, few mitochondria, and many free ribosomes. system is similar to other ones described in non-primate mammals.
SPG are divided in 3 subtypes based on their heterochromatin con- In addition, there is some debate about the spatial organization of
tent: A dark; A pale; and B SPG [130,131]. cellular associations. Animal models have shown that spermatoge-
14 F.T.L. Neto et al. / Seminars in Cell & Developmental Biology 59 (2016) 10–26

nesis is organized in waves throughout the STs. Early human studies DNA. During meiosis 2, a single secondary spermatocyte is turned
failed to find the same pattern, and described human spermatoge- into two haploid/n round Sa spermatids [161,162].
nesis as a chaotic process [3]. However, subsequent studies using Maturation arrest (MA) during meiosis (early maturation arrest)
computer modeling to map the distribution of primary sperma- is found in approximately 10% of the men with NOA. The
tocytes proposed that human spermatogenesis was organized in histopathologic features of early MA include reduced tubule diam-
helical waves [147,148]. Recently, 3D maps of cellular associations eter and number of GCs, degenerating spermatocytes, and cellular
showed that their arrangement in humans was a random event debris [163]. Since meiosis 1 is more complex and has more quality
[149]. control checkpoints than meiosis 2, MA at the stage of PSCs is more
Failure of GCs to develop further than the spermatogonial stage often found. Men with focal early MA may have some STs with nor-
leads to azoospermia and a histological pattern called Sertoli- mal spermatogenesis, and the overall surgical sperm retrieval rate
cell–only syndrome (SCO). SCO can be characterized by complete in these cases ranges from 20 to 40% with microsurgical testicular
absence of SPG (SCO type 1), or by presence of rare foci of GCs with sperm extraction (microTESE) [164–166].
residual spermatogenesis (SCO type2). SCO type 1 is caused by dis- As previously described, an important characteristic of sperma-
turbance of the migration of the primordial GCs from the yolk sac tocytes and spermatids is their reliance on lactate as their principal
to the seminiferous cord, and exhibits intact basal membrane with energy source [87]. Glucose is taken up by SCs via the specific
numerous SCs in a very good morphological shape, but no GCs can glucose transporter GLUT1 and then processed glycolytically into
be found. SCO type 2 is due to an insult occurring later on, allowing lactate, which is then transferred to GCs via monocarboxylate
the persistence of some SPG, but also damaging other components transporters. FSH, androgens, insulin and some paracrine factors
of the STs. Despite the great reduction of the number of SPG and stimulate this process, and disruptions can cause male infertility
the presence of alterations in the other components of STs (i.e. SCs [167,168].
and basal membrane), sperm may still be retrieved using testicular Spermiogenesis is the maturation process by which one round
microdissection techniques [150–152]. Sa spermatid becomes one spermatozoon. During this stage, no
New promising technologies are been developed to help the more cellular divisions occurs, and a series of cytoplasmic and
identification of STs with undisturbed spermatogenesis in men nuclear changes takes place. The acrosome is formed from the
with NOA. Multiphoton microscopy illuminates tissues with a near- Golgi apparatus with participation of a cytoskeletal structure called
infrared laser, which excites auto-fluorescence. This technique perinuclear theca [169,170]. Disturbances during acrosome devel-
enables real-time imaging of tissue in-vivo without labels, provid- opment could lead to dramatic alteration in the head shape, called
ing high-resolution images that can be obtained at clinically useful globozoospermia, a rare infertility condition that is difficult to
depths with minimal damage to the tissue. The images can be used manage. Globozoospermia is characterized by round-headed sper-
to differentiate between normal and abnormal spermatogenesis matozoa that are unable to penetrate and activate oocytes. Failure
in human testis [153]. Other technique, Raman spectroscopy (RS), of inner nuclear proteins, mainly the DPY19L2 protein, to anchor
is a noninvasive and label-free optical technique that utilizes the the acrosome to the nucleus is the principal mechnism behind
molecular fingerprint of materials and transforms their biochemi- this pathology [171,172]. DPY19L2 gene deletions are considered
cal information into a characteristic Raman spectrum. RS has been the most commom cause of globozoospermia, while SPATA16 and
tested in animal and human models and has shown capability to PICK1 mutations are also associated to this phenotype [173]. The
distinguish seminiferous tubules with spermatogenesis from SCO outcomes of conventional intracytoplasmic sperm injection (ICSI)
tubules [154]. These technologies deserve further investigation into are poor in these men, since their sperm cannot active the oocyte.
their clinical application to improve the sperm retrieval rates in ICSI coupled with assited oocyte activation has resulted in live
patients with NOA (Fig. 2). births [174].
Primary spermatocytes (PSCs) are produced by mitotic division The nucleus becomes more condensed and migrates to an eccen-
of type B SPG and mark the end of the proliferative portion of sper- tric position. Protamines replace 85% of the nuclear histone content
matogenesis. They are the first GCs to cross the BTB to the adluminal and DNA is neatly packed around them in supercoiled structures
compartment, where they become immunologically isolated and named toroids, increasing DNA protection and becoming, in most
proceed to meiosis. PSCs are subclassified according to the stages of part, transcriptionaly silent. In this setting of highly condensed
prophase 1 (i.e. preleptotene, leptotene, zygotene, and pachytene) DNA, protein synthesis relies on RNA stored in the cromatoid body.
[7]. Some of these RNAs have been identified as non-coding RNAs,
The DNA recombination process occurs during the pachytene which have a role in gene expression regulation. Poor DNA packing
stage and is crucial to human evolution, therefore, quality control results in DNA fragmentation and decreases the fertility potential,
of meiosis 1 is very strict, with checkpoints during the pachytene impacting negatively on IVF outcomes [175–179].
and the transition metaphase1/anaphase1. Ten percent of men with The sperm tail arises from the centriole, gradually elongating
NOA show a higher rate of recombination failure [155], while up the spermatid shape, and is formed of microtubules disposed in a
to 45% of men whose partner suffer from recurrent pregnancy 2 + 9 formation. In addition to the microtubules, other two struc-
loss, and who have normal sperm density, motility, and morphol- tures, the outer dense fibers and the fibrous sheath, are organized
ogy have increased sperm aneuploidy [156]. These findings could and provide additional support and flagellar motion regulation.
explain many cases of “idiopathic infertility”. Such infertile men Mitochondria increase in number and gather around the tail basis,
might be at a higher risk of producing chromosomally abnormal forming the midpiece [180–182]. In order to maximize sperma-
offspring and should have genetic counseling [155,157]. tozoa motility, the excess of cytoplasm containing the remnants
Another important player during meiotic recombination was of Golgi apparatus and other organelles is removed, forming the
recently identified. TEX11 is a germ cell-specific gene located in residual body, which is phagocytized by SCs [183].
the X chromosome and encodes a protein that regulates homolo- Spermiogenesis may be interrupted at any stage of spermatid
gous chromosome synapses and double-strand DNA break repair development. When the arrest occurs at the early spermatid phase
[158,159]. TEX11 mutations were found in 2.4% of men with “idio- (dark round nuclei), the condition is called late maturation arrest,
pathic infertility” and were associated to spermatocyte apoptosis, and when condensed, oval spermatids are present, it is called
maturation arrest and azoospermia [160]. hypospermatogenesis (HS) [163]. Late MA seems to be less com-
After meiosis 1, two secondary spermatocytes are formed, each mon than early MA, but carries a better chance of sperm retrieval
one with haploid number of chromosomes, but with 2n content of with microTESE (50–80%), while hypospermatogenesis is found in
F.T.L. Neto et al. / Seminars in Cell & Developmental Biology 59 (2016) 10–26 15

Fig. 2. Different seminiferous tubular histology patterns imaged with multiphoton microscopy at lower magnification (A, D, G) and high magnification (B, E, H) compared to
high magnification H&E stained tissue. Normal spermatogenesis (A, B, C). Seminiferous tubules with Sertoli cells only (D, E, F). Seminiferous tubules with Sertoli cells only
and moderate peritubular fibrosis (G, H, I). Scale bar represents 500 ␮m for low magnification and 80 ␮m for high magnification.
Reprinted from The Journal of Urology, 2012; Najari B.B., Ramasamy R., Sterling J., Aggarwal A., Sheth S., Li P.S., et al. Pilot study of the correlation of multiphoton tomography
of ex vivo human testis with histology. 188:538–543 with permission from Elsevier

up to 60% of men with NOA and has the best surgical sperm retrieval 2.6. Blood testis barrier
rate among all NOA histological patterns (80–90%) [163,184,185]
(Fig. 3). Due to their high immunogenicity and specific metabolic
Once the spermatid is fully metamorphosed, it is detached from demands, GCs need to be kept in a tightly regulated microenviron-
SCs and released into the tubule lumen in a process called spermia- ment. The BTB is an anatomical and functional barrier that divides
tion. The details of spermiation in humans have yet to be elucidated, the seminiferous epithelium in two separate spaces, the basal and
but it seems that the detachment of ectoplasmic specializations adluminal compartments, and restricts the paracellular transit of
progress from tail to head, and that small tubulobulbar complexes substances [188,189]. In mammalian testes, the most important
maintain the last point of contact. This step is FSH and testosterone- components of the BTB are the cellular junctions between adjacent
dependent. Although defects in spermiation alone have not been SCs, such as tight junctions, basal ectoplasmic specialization, gap
found to cause human male infertility, they likely do contribute in junction, and desmosomes. Peritubular myoid cells and endothelial
some cases [104,186,187]. cells play a secondary role [190,191].
The first cells to cross the BTB are the leptotene-PSCs, which
migrate to the adluminal compartment to finish meiosis. Thus,
16 F.T.L. Neto et al. / Seminars in Cell & Developmental Biology 59 (2016) 10–26

Fig. 3. Early (left) and late maturation arrest (right).


Reprinted from Fertility and sterility 2015. Bernie A.M., Shah K., Halpern J.A., Scovell J., Ramasamy R., Robinson B., et al. Outcomes of microdissection testicular sperm extraction
in men with nonobstructive azoospermia due to maturation arrest. 104:569–573 with permission from Elsevier.

meiosis initiation may not require the special microenvironment in obese men due to the increased peripheral conversion of testos-
[192], but its consummation should occur in the adluminal com- terone to estrogens by the aromatase enzyme found in the adipose
partment to avoid activation of the immune system by surface tissue [216,217]. The excess of estrogens results in inhibition of
antigens of haploid GCs. Since GCs in the basal compartment also LH and FSH by negative feedback on the hypothalamus and pitu-
contain autoantigens, the BTB immunologic shielding seems to be itary gland, decreasing T levels and the testosterone/estrogen ratio.
reinforced by immunosuppressive activity of leukocytes and SCs With low T and FSH levels, spermatogenesis becomes defective.
[85,193–195]. This mechanism is evidenced by the low levels of inhibin B in obese
The transport of PSCs to the adluminal compartment, as well as men [218–220]. Estrogen receptors (ER) have also been described
the transport of round spermatids to the luminal edge, appears to in human testis. ER␤ is found in almost all cell types of the inter-
be mediated by assembly and disassembly of cell junctions in an stitium and the STs, ER␣ is mainly found in LCs and SCs, and the
orderly manner to prevent the passage of unwanted substances G protein-coupled estrogen receptor (GPER) is present in PTCs, LCs
[85,192,196]. Furthermore, SCs transporters, such as efflux ATP and SCs. These findings imply a direct impact on testicular function
pumps, mediate the flow of molecules, so that SCs are able to [221–224].
provide the nutrients needed by GCs, and to remove harmful sub- Other mechanisms involved in the hypoandrogenism of obese
stances. This mechanism is implicated in the resistance shown by patients include diminished levels of sex hormone-binding glob-
some drugs to penetrate the BTB [197,198]. ulin (SHBG), insulin and leptin resistance, sleep apnea, and
Whereas several reports and animal models have suggested that adiponectin deficiency [225–230].
the BTB is regulated by hormones, cytokines and growth factors, Factors involved in the pathogenesis of obesity, such as high-
data gathered from the few human studies covering this topic are calorie diets, sedentarism, genetic and epigenetic disorders, may
not enough to draw definitive conclusions. For example, animal also influence sperm production [207,231,232]. Furthermore, some
studies have shown that intratesticular T acts as a trophic factor, bariatric surgeries have been shown to impair semen parameters,
however, in a human study, the suppression of intratesticular T probably due to malabsorption of vitamins and trace nutrients
did not disturb the BTB integrity. The study of BTB is an evolving [233–235].
field, but more human studies are needed. Advances in this area
will likely bring us clinical tools applicable to humans in the area
of contraception, male infertility and oncology [199–202].

3. Factors that directly affect human spermatogenesis 3.2. Diabetes

3.1. Obesity Characterized by hyperglycemia due to failure of insulin produc-


tion and action, diabetes mellitus (DM) induces damage to various
Obesity has become one of the most studied health problems in organs and systems, including the testis. Epidemiologic studies
the recent years. Using a body mass index (BMI) >30 as the defini- have shown that men with type 1 DM have significantly fewer off-
tion for obesity, the WHO estimates that 13% of the world’s adult spring than their unaffected siblings [236], and that the estimated
population was obese in 2014 [203]. This scenario is even worse in infertility prevalence in diabetic males ranges from 35% to 51%
the US, where a recent study showed a prevalence of 35% in adults [237]. Despite the conflicting data about the impact of DM on clas-
[204]. Parallel to that, there is an ongoing debate about the pos- sical sperm parameters [238,239], there is some evidence pointing
sible relationship between the controversial worldwide decline in towards direct testicular damage. Diabetic men have increased
semen quality and the increasing number of obese men [205,206]. DNA fragmentation [240,241], which is probably due to oxidative
Several studies have shown a negative association between stress originated from increased levels of advanced glycation end
sperm concentration/total sperm count and increasing BMI products, specifically N␧-carboxymethyl-lysine [242–245].
[207–210]. A negative impact on sperm morphology, motility, and In addition, a recent study reported decreased lactate pro-
DNA fragmentation has also been reported [211–215]. duction by human SCs during insulin deprivation. This metabolic
Several mechanisms have been proposed to explain how obe- mechanism could directly affect spermatogenesis, since SCs-
sity could affect sperm production. The most studied mechanism is derived lactate has an anti-apoptotic effect and is the main energy
the hyperestrogenism. Levels of circulating estrogens are increased source for spermatocytes and spermatids [114].
F.T.L. Neto et al. / Seminars in Cell & Developmental Biology 59 (2016) 10–26 17

3.3. Environmental chemicals fore, intratesticular temperature should be kept 2–4 ◦ C lower than
the rectal temperature [275]. Heat causes an increased rate of
That exposure to environmental chemicals (ECs) affects sper- GCs apoptosis [276], likely mediated by decreased levels of cold-
matogenesis has been suspected since the ancient times. Some inducible RNA binding protein (Cirp) [277], as well as elevated
Roman emperors and patricians had reproductive problems likely levels of heat shock proteins [278]. High temperatures also impair
due to chronic exposure to lead from drinking contaminated wine testicular androgen production via a pathway involving increased
[246,247]. However, the definitive link between lead exposure and oxidative stress damage to the LCs [279].
male infertility was not established until 1975 [248]. Recently, studies of the seminal proteome of men with
Even though data from in vitro and animal studies corroborate varicoceles have highlighted the role of seminal components
the negative effects of ECs on the male reproductive tract, human in the condition’s pathophysiology. Adolescents with varicocele
epidemiological studies have been somewhat inconclusive. and abnormal semen parameters showed high expression of
There are many pathways by which ECs could disturb spermato- semenogelins I and II [280]. Semenogelins inhibit sperm motility
genesis, endocrine disruption is the most well known. Phthalates, and prevent premature sperm hyperactivation and capacitation,
chemicals used as plasticizers, have an anti-androgenic effect, thus, they may be responsible, at last in part, for the decreased
decreasing the T production in LCs by activation of peroxisome sperm motility found in these patients. The same group reported
proliferator-activated receptors (PPARs), as well as reduction of increased DNA-directed RNA polymerase III subunit (RPC2) levels
TPSO levels [249]. In addition, phthalates can also impair SCs in adolescents with varicoceles [281]. Since this protein is involved
and GCs functions [250]. Bisphenol A (BPA), a chemical used to in the oxidative cascade, this reinforces the role of oxidative stress
manufacture polycarbonate plastics and epoxy resins, is another in the pathophysiology of varicocele. It is important to note that
endocrine disruptor. BPA binds to AR, ER␣ and ER␤, exerting anti- seminal proteome analysis has been extensively studied in infer-
androgenic and anti-estrogenic effects, and exposure to this EC has tile men with the objective of clarifying the importance of seminal
been associated to poor semen parameters [251,252]. components in several male infertility conditions, as well as iden-
The BTB is likewise a target for ECs. Cadmium, a heavy metal tifying biomarkers that could be used as diagnostic tools. For good
used in the manufacture of batteries and pigments, has been shown reviews, refer to Ref. [282,283].
to cause damage to the BTB, disrupting tight junctions and caus- Another potential mechanism for testicular function impair-
ing spermiation failure [253,254]. Tetrachlorodibenzo-p-dioxin is ment is a decreased expression of E-cadherin and alpha-catenin
a chlorinated hydrocarbon formed as a side product of herbicides proteins in SCs, with subsequent damage to the BTB and autoim-
and pesticides synthesis. Dioxin increases the c-Src activity, which munity [284,285]. Impaired disposal of residual sperm cytoplasm is
alters the adhesive function of tight junctions and adherens junc- also proposed as possible mechanism, resulting in defective sperm
tion in the BTB [255,256]. Not surprisingly, men living in areas with function [286,287].
high dioxin contamination have impaired semen parameters [257]. The histological findings in infertile men with varicocele include
Alterations in the sperm chromatin structure is linked to some HS, MA, decreased STs diameter, LCs hyperplasia, and testicular
ECs, such as lead and ethylene dibromide. Lead is usually used as a atrophy [288–290]. The effects on spermatogenesis seem to be
pigment in paints and as a fuel component [258]. It binds to human bilateral, even in cases of unilateral varicocele, but the mechanism
protamines during spermiogenesis, altering sperm chromatin sta- by which the contralateral testis is affected is unclear [291].
bility and potentially affecting normal chromatin condensation
[259]. Ethylene dibromide (EDB) is a lead scavenger used in gasoline 3.5. Genetic factors
and was used as a fumigant. Studies with men chronically exposed
to EDB showed a negative impact on sperm count, morphology Genetic disorders account for 15–30% of male infertility cases
and motility [260,261]. Animal models suggests that EDB binds to and might be responsible for the majority of “idiopathic” cases
histones, disturbing DNA packing [262]. [292,293]. Since the first reports of successful pregnancies after
Increased oxidative stress is also associated to some ECs. 1,2- testicular sperm retrieval coupled with intracytoplasmic sperm
Dibromo-3-chloropropane (DBCP) is a fumigant once used to injection of eggs (ICSI) [294], the interest on this field has signifi-
control nematodes in field crops, but is still used in the synthesis cantly increased as specialists try to gather data that could be useful
of some fire retardants [263]. Men exposed to DBCP had decreased in the management and counseling of these couples. Notwith-
sperm count [264]. An in vitro study suggested that DBCP elevates standing the great advances in genetics over the last two decades,
ROS in GCs, inducing spermatogonial apoptosis [265] (Table 1). bench-to-bedside translation of this knowledge has been slow, hin-
dering the development of diagnostic and therapeutic tools for
3.4. Varicocele clinical practice. This section will focus on genetic factors affecting
human spermatogenesis with current, as well as potential future,
Varicocele is an abnormal dilation of internal spermatic veins clinical applicability.
caused by incompetent venous valves, leading to reflux and stasis
of venous blood. Varicocele can be found in 15% of all adult males, in 3.5.1. Klinefelter syndrome
35% of infertile men and in 70–80% of men with secondary infertility Klinefelter syndrome (KS) is the most common genetic cause of
[266]. Varicocele is considered a common etiology of male infertil- male infertility, with a prevalence of 5% in men with severe oligo-
ity, however, the cause-effect relationship between varicocele and zoospermia and 10% in men with NOA [295,296]. Eighty percent of
infertility has not been conclusively established yet. Although stud- the men are 47,XXY, while the remaining have mosaic patterns such
ies confirm a negative impact of varicocele on fertility, studies of as 46,XY/47,XXY, 48XXXY or 48 XXYY [297]. The extra X chromo-
unselected men reveal conflicting results [267–270]. Despite the some has been shown to be of paternal origin in 60% of the cases.
controversy, surgical treatment of infertile men with clinical varic- In this setting, X-Y non-disjunction during meiosis 1 is the most
oceles has been shown to improve semen parameters, pregnancy common error, and diminished X-Y recombination has also been
rates and testosterone levels [271–274]. described. For the maternal cases, non-disjunction due to errors in
Several hypotheses have been proposed to explain the potential meiosis is associated with increased maternal age [298].
negative effect of varicocele on spermatogenesis. Scrotal hyper- Despite the inactivation of an extra X chromosome in mam-
thermia caused by venous blood stasis is thought to be the primary mals, it is known that approximately 15% of the genes continue
mechanism. Testicular functions are temperature-sensitive, there- to be active in the silent chromosome. This process is skewed
18 F.T.L. Neto et al. / Seminars in Cell & Developmental Biology 59 (2016) 10–26

Table 1
Summary of some of the ECs that affects human spermatogenesis.

Chemical Agent Application Exposure route Effect

Phthalates Plastics, tubings Skin contact, inhalation Anti-androgenic


Bisphenol A Polycarbonate plastics and epoxy resins Ingestion Anti-androgenic, anti-estrogenic
Cadmiun Batteries and pigments Smoking, occupational BTB disruption, Decreased T levels
Tetrachlorodibenzo-p-dioxin Herbicides and pesticides Occupational, ingestion BTB disruption, Anti-androgenic
Lead Pigment in paints and fuel Inhalation Chromatin disruption, damage to mannose receptors
Ethylene dibromide Fuel component and fumigant Inhalation Chromatin disruption
1,2-Dibromo-3-chloropropane Fumigant Inhalation, skin contact Increased oxidative stress

in patients with KS, resulting in an excessive genetic output that removes 8 genic families, including the DAZ family. Smaller partial
impairs androgen production and spermatogenesis [298,299]. deletions also exist and may happen either via HR, such as “b1/b3”,
The main clinical characteristics of the KS are infertility, “b2/b3” and “gr/gr”, or via non-homologous recombination, such
hypergonadotrophic hypogonadism, and cognitive disorders. The as P3a, P3b, P3c and P3b [316,319].
phenotype spectrum is wide, depending on the parental origin of Due to the wide variability of these deletions, the clinical and
the X chromosomes, the extent of genetic inactivation, and the histological presentations are variable, but, in general, AZFc dele-
presence of mosaicism [298,300]. AR gene inactivation also seems tions are compatible with residual spermatogenesis. Patients may
to play an important role. Since the length of polymorphic stretch present with azoospermia or severe oligozoospermia, and his-
of CAG repeats contained in the exon 1 of the AR gene is inversely tological findings vary from SCO, to MA and HS. Viable sperm
related to the receptor’s activity, inactivation of the AR gene with a can be found in up to 70% of the azoospermic patients that
shorter or longer stretch of CAG repeats may be related to the sever- undergo microTESE [312]. Complete AZFc deletions could cause
ity of the syndrome [301]. Furthermore, a recent paper reported Y-chromosome loss and lead to 45X/46XY karyotype with Turner
that T production by LCs is normal or even increased in men with stigmata or sexual ambiguities. To avoid the transfer of 45, X0
KS [302]. These evidences point toward a lower release of T into the embryos, preimplantation diagnosis should be offered to these cou-
bloodstream associated to the lack of responsiveness of the AR. ples [320].
Regarding the impact of KS on spermatogenesis, a common Responsible for 15% of YCMD, the AZFb group is located from
feature is progressive degeneration of GCs and SCs, mainly after 18.1 to 24.7 Mb of the Y chromosome and contains the RBMY1
puberty [303,304]. The mechanisms leading to testicular degenera- and PRY genes. The first is a testis-specific splicing factor expressed
tion are still unknown, but overexpression of X chromosome genes, in the nucleus of spermatogonia, spermatocytes, and round sper-
such as the angiotensin type-II receptor and the TEX11 genes, mal- matids, and the second is involved in the regulation of GCs
function of FSH and androgen receptors, and increased aromatase apoptosis [316]. Complete deletions of this zone (proximalP5/P1)
activity may have a role [303,305,306]. are massive, perhaps the largest in human genome. Homologous
Although men with KS and azoospermia show dramatic testicu- and non-homologous recombinations participate in these events,
lar alterations, the success rate of microTESE in finding viable sperm but other unknown factors may be also involved [321]. Since AZFb
is 68%, and all the children born using IVF/ICSI with sperm from men overlaps AZFc by 1.5 Mb, combined AZFb + AZFc deletions occur and
with KS were healthy in one study [307]. This is probably due to are, indeed, more frequent than isolated AZFb deletion [312,322].
small niches of undisturbed spermatogenesis composed by either Patients with complete AZFb or AZFb + AZFc deletions show
a few GCs with normal karyotype, or by some 47,XXY GCs that are azoospermia, and testicular biopsy usually reveals SCO or diffuse
able to go through meiosis and produce sperm with normal kary- early MA, hence, these patients must use donor sperm or adoption
otype (23,XY or 23,XX) [308,309]. Despite that, the risk of using [312].
a hyperploid 24,XY sperm exists, and the conception of a 47,XXY The AZFa group is located closer to the centromere and con-
fetus has been reported, thus, preimplantation genetic diagnosis is tains three genes: DBY, USP9Y and UTY. The DBY gene acts as a
indicated [310]. spermatogenic regulator during the earliest stages (i.e. spermato-
gonia), while the other two apparently are not crucial for male
3.5.2. Y-chromosome microdeletions fertility [323,324]. AZFa deletions have been shown to occur via
Located in the euchromatin zone of the long arm of the Y intrachromosomal recombination between flanking repeats [325].
chromosome (Yq11), the AZF (azoospermia factor) region contains Complete deletion of AZFa is rare (3% of all YCMD) and car-
genes critical for spermatogenesis, among them, at least fourteen ries the worse prognosis among all YCMD. Invariably all patients
protein-encoding genes. These genes are divided in three groups show azoospermia and SCO, and no sperm is found on microTESE.
based on their location: AZFa, AZFb and AZFc [311]. Microdeletions Therefore, these patients should not be submitted to invasive
occurring in any of these zones have the potential to impair fertil- sperm retrieval procedures. In contrast, men with partial AZFa
ity and are found in 10% of men with non-obstructive azoospermia deletions often have HS and present with severe oligospermia or
and 5% of those with severe oligozoospermia, but the incidence, cryptozoospermia [312,322]. We expect that genetic engineering
and even the phenotypes, vary geographically and ethnically advances will allow us to induce meiosis progression in stem cells,
[312–314]. producing spermatozoa and helping men with complete AZFa, AZFb
The AZFc group is located in the distal aspect of Yq11 and and AZFb + c deletions to father offspring in the future (Fig. 4).
accounts for 60% of all Y-chromosome microdeletions (YCMD)
[315,316]. Several genes are located in the AZFc group, and the 3.5.3. Micro RNAs
DAZ (deleted in azoospermia), a family of four genes implicated Micro-RNAs (miRNAs) are a class of short (20–23 nucleotides)
in spermatogenesis, has been the most studied [317]. The relative single-stranded non-coding nucleotides, and constitute one of the
high incidence of de novo deletions via homologous recombina- most abundant ribonucleoprotein complexes in the cell. miRNAs
tion (HR) in this group is a product of the arrangement, similarity exert a regulatory function over the expression of several protein-
and the huge size of its amplicons, repetitive copies of nucleic acid coding genes, and, therefore, modulate a wide array of biological
sequences [318]. The most frequent deletion affecting the AZFc processes. Their mechanisms of action are still under debate,
group is the one involving the amplicons b2 and b4 (b2/b4), which but may include direct destruction of targets mRNAs, translation
F.T.L. Neto et al. / Seminars in Cell & Developmental Biology 59 (2016) 10–26 19

Fig. 4. Y-chromosome microdeletions [293,312,314,322].

Table 2
List of the most studied miRNAs in infertile men [329,335–338].

Name Function Expression Findings associated

miR-34 family p53 tumor suppressor network Down NOA


miR-122 Suppresses the transcription of transition protein 2 Down NOA
miR-19b Inhibition of apoptosis Up NOA
Let-7a Cell proliferation Up NOA
miR-181a Regulation of T cell sensitivity Down NOA
miR-146b Regulation of apoptosis Down NOA
miR-513a-5p Regulation of apoptosis Down NOA
miR-509-5p Regulation of apoptosis Down NOA
miR-374b Oncogene Down NOA
miR-141 Regulation of cell cycle Up NOA
miR-429 Unclear Up NOA
miR-202-5p Unclear Down NOA
miR-7-1-3p Regulation of cell cycle Up NOA

repression, and other indirect pathways to inhibit protein synthe- associated with gene suppression, while hypomethylation is linked
sis [326–328]. Due to the fact that the expression of miRNAs varies to gene expression [177,340].
among different developmental stages, tissues and diseases, spe- Histone methylation, acetylation, phosphorylation, ubiquity-
cific expression patterns could be linked to specific pathologies, lation and sumotylation also modulate gene expression. Amino
and, thus, be used as diagnostic and therapeutic tools [329]. acid residues in the N-termini of histone tails are the sites for
Regarding spermatogenesis, several animal studies showed that post-translational modifications, and several enzymes are involved
GCs miRNAs, many of them stored in the cromatoid body, are in the process. The final effect, gene activation or suppression,
implicated in the regulation of apoptosis, proliferation and differ- depends on the combination among different sites and radicals
entiation. This post-transcriptional regulation is essential because [339,341–343].
GCs are transcriptionally silent during certain stages of spermato- In addition, the manner by which DNA segments are packed
genesis. Alterations of miRNAs expression patterns could impair around histones determines whether or not they are available
spermatogenesis, and might explain a number of “idiopathic” male for transcription. Tightly packed DNA segments found in het-
infertility cases [330–332]. erochromatin are silent, while the loose DNA segments that
Recently, efforts have been directed to associate specific miRNAs constitute euchromatin are usually transcriptable, therefore, chro-
expression patterns in the seminal plasma with human testicular matin remodeling may activate or inhibit gene expression. The
histopathologic patterns and clinical findings, with the idea of cre- exact remodeling mechanisms are still unknown, but they appear to
ating new diagnostic tools to assess the human male fertility [333]. be regulated by ATP-dependent chromatin remodeling complexes
So far 1881 humans miRNAs have been described (www.mirbase. [344,345].
org) [334] (Table 2). The DNA content in the male GCs is packed in a very small
volume to fit into the sperm head. To accomplish that, 80% of
the histone content should be replaced by protamines via hyper-
3.5.4. Epigenetic factors acetylation of histone H4 during spermatid stages. Highlighting
Epigenetics is the study of several processes that alter gene the importance of this process, the degree of histone-protamine
expression without changing the DNA sequence. Some of these replacement has been correlated with the fertilizing capacity of
processes are DNA methylation, post-translational histone modifi- the sperm, and decreased levels of H4 hyperacetylation were
cations, and chromatin remodeling. They can be cell, tissue, organ, demonstrated in men with MA [346,347]. Furthermore, the resid-
sex and species-specific, can vary with different developmental ual histone-bond DNA content is thought to be crucial for sperm
stages, may be carried through generations and may be reversible function and early embryo development. The ratio between differ-
[339]. ent types of protamine also affects fertility, as shown by studies
DNA methylation occurs when a methyl radical is added to demonstrating an increased DNA fragmentation in men with low
cytosine-guanine dinucleotides (CpG) by DNA methyltransferases. P1/P2 ratio [339,348–350].
Areas of DNA with high content of CpG, called “CpG islands”, have
been found near promoters, and hypermethylation of CpG islands is
20 F.T.L. Neto et al. / Seminars in Cell & Developmental Biology 59 (2016) 10–26

A key event during spermatogenesis is the epigenetic repro- receptor in testis peritubular myoid cells, Proc. Natl. Acad. Sci. U. S. A. 103
gramming of GCs by widespread erasure of DNA methylation (2006) 17718–17723.
[14] M. Welsh, P.T. Saunders, N. Atanassova, R.M. Sharpe, L.B. Smith, Androgen
followed by de novo methylation. The reprogramming takes place action via testicular peritubular myoid cells is essential for male fertility,
during GCs differentiation during gonadal development, and dur- FASEB J. 23 (2009) 4218–4230.
ing spermatogenesis, establishing a male germ line pattern of [15] Y. Kato, K. Shiraishi, H. Matsuyama, Expression of testicular androgen
receptor in non-obstructive azoospermia and its change after hormonal
DNA hypomethylation. Not surprisingly, elevated DNA methyla- therapy, Andrology 2 (2014) 734–740.
tion at numerous sequences has been associated with poor quality [16] H. Welter, A. Huber, S. Lauf, D. Einwang, C. Mayer, J.U. Schwarzer, et al.,
human sperm. However, the relationship between widespread Angiotensin II regulates testicular peritubular cell function via AT1 receptor:
a specific situation in male infertility, Mol. Cell. Endocrinol. 393 (2014)
DNA methylation and fertility is still a matter of debate [351–354].
171–178.
Since some epigenetic alterations can be transmitted through [17] J. Frayne, H.D. Nicholson, Localization of oxytocin receptors in the human
generations, attention has been drawn to the association between and macaque monkey male reproductive tracts: evidence for a physiological
role of oxytocin in the male, Mol. Hum. Reprod. 4 (1998) 527–532.
assisted reproductive technologies and pathologies related to
[18] C. Schell, M. Albrecht, C. Mayer, J.U. Schwarzer, M.B. Frungieri, A.
genomic imprinting, such as Prader-Willi, Beckwith-Wiedemann Mayerhofer, Exploring human testicular peritubular cells: identification of
and Angelman syndromes. This may occur due to the use of secretory products and regulation by tumor necrosis factor-alpha,
defective sperm with incomplete reprogramming, or epigenetically Endocrinology 149 (2008) 1678–1686.
[19] S.G. Haider, J. Talati, G. Servos, Ultrastructure of peritubular tissue in
imperfect oocytes arising from super-ovulation. Other cause may association with tubular hyalinization in human testis, Tissue Cell 31 (1999)
be ART procedures performed at the time of epigenetic reprogram- 90–98.
ming [177,339,350,355]. [20] R. Paniagua, P. Amat, M. Nistal, A. Martin, Ultrastructure of Leydig cells in
human ageing testes, J. Anat. 146 (1986) 173–183.
Epigenetic mechanisms may also be the way by which several [21] S.G. Haider, Cell biology of Leydig cells in the testis, Int. Rev. Cytol. 233
diseases and conditions, such as obesity and environmental expo- (2004) 181–241.
sure, affect spermatogenesis and influence the offspring [356–358]. [22] R.E. Mancini, O. Vilar, J.C. Lavieri, J.A. Andrada, J.J. Heinrich, Cytological and
cytochemical study of the development of the Leydig cells in the normal
However, further studies are needed to clarify these associations. human tasticle, Rev. Soc. Argent. Biol. 36 (1960) 443.
[23] F.P. Prince, The triphasic nature of Leydig cell development in humans, and
comments on nomenclature, J. Endocrinol. 168 (2001) 213–216.
4. Conclusion
[24] K. Svechnikov, L. Landreh, J. Weisser, G. Izzo, E. Colon, I. Svechnikova, et al.,
Origin, development and regulation of human Leydig cells, Horm. Res.
Advances in Andrology over the last three decades have paved Paediatr. 73 (2010) 93–101.
the way for the elucidation of the molecular mechanisms that con- [25] J. Codesal, J. Regadera, M. Nistal, J. Regadera-Sejas, R. Paniagua, Involution of
human fetal Leydig cells. An immunohistochemical, ultrastructural and
trol human spermatogenesis. Rapid development of translational quantitative study, J. Anat. 172 (1990) 103–114.
medicine and bench-to-beside collaborations are indispensable to [26] S.M. Mendis-Handagama, H.B. Ariyaratne, Differentiation of the adult Leydig
further scientific knowledge. Much of the data reported here was cell population in the postnatal testis, Biol. Reprod. 65 (2001) 660–671.
[27] A.H. Payne, D.B. Hales, Overview of steroidogenic enzymes in the pathway
first hypothesized in the clinical scenario, then elaborated using from cholesterol to active steroid hormones, Endocr. Rev. 25 (2004)
animal models, confirmed in human descriptive studies, and finally 947–970.
transported back to clinical practice via clinical trials. We still are [28] W.L. Miller, R.J. Auchus, The molecular biology, biochemistry, and physiology
of human steroidogenesis and its disorders, Endocr. Rev. 32 (2011) 81–151.
far from completely comprehending the origin and function of the [29] V. Papadopoulos, M. Baraldi, T.R. Guilarte, T.B. Knudsen, J.J. Lacapere, P.
male gamete, but now we can have glimpses of this enormous Lindemann, et al., Translocator protein (18 kDa): new nomenclature for the
universe. peripheral-type benzodiazepine receptor based on its structure and
molecular function, Trends Pharmacol. Sci. 27 (2006) 402–409.
[30] C.B. Kallen, J.T. Billheimer, S.A. Summers, S.E. Stayrook, M. Lewis, J.F. Strauss,
References Steroidogenic acute regulatory protein (StAR) is a sterol transfer protein, J.
Biol. Chem. 273 (1998) 26285–26288.
[1] P. Turek, Male reproductive physiology, in: A.J. Wein, L.R. Kavoussi, A.W. [31] L.K. Christenson, J.F. Strauss 3rd, Steroidogenic acute regulatory protein: an
Partin, C.A. Peters (Eds.), Campbell-Walsh Urology, 11th ed., Elsevier update on its regulation and mechanism of action, Arch. Med. Res. 32 (2001)
Saunders, Philadelphia, PA, 2016, pp. 516–537. 576–586.
[2] C.G. Heller, Y. Clermont, Spermatogenesis in man: an estimate of its [32] E. Anuka, M. Gal, D.M. Stocco, J. Orly, Expression and roles of steroidogenic
duration, Science 140 (1963) 184–186. acute regulatory (StAR) protein in ‘non-classical’, extra-adrenal and
[3] C.H. Heller, Y. Clermont, Kinetics of the germinal epithelium in man, Recent extra-gonadal cells and tissues, Mol. Cell. Endocrinol. 371 (2013) 47–61.
Prog. Horm. Res. 20 (1964) 545–575. [33] R.S. Swerdloff, P.C. Walsh, Pituitary and gonadal hormones in patients with
[4] L.M. Misell, D. Holochwost, D. Boban, N. Santi, S. Shefi, M.K. Hellerstein, varicocele, Fertil. Steril. 26 (1975) 1006–1012.
et al., A stable isotope-mass spectrometric method for measuring human [34] J.P. Jarow, H. Chen, T.W. Rosner, S. Trentacoste, B.R. Zirkin, Assessment of the
spermatogenesis kinetics in vivo, J. Urol. 175 (2006) 242–246 (discussion 6). androgen environment within the human testis: minimally invasive method
[5] R.P. Amann, The cycle of the seminiferous epithelium in humans: a need to to obtain intratesticular fluid, J. Androl. 22 (2001) 640–645.
revisit, J. Androl. 29 (2008) 469–487. [35] J.P. Jarow, W.W. Wright, T.R. Brown, X. Yan, B.R. Zirkin, Bioactivity of
[6] E.N.J. Kobil, Cytology of testis and intrinsic control mechanisms, in: E. androgens within the testes and serum of normal men, J. Androl. 26 (2005)
Knobil, J.D. Neill, T.M. Plant (Eds.), Knobil and Neill’s Physiology of 343–348.
Reproduction, 3rd ed., Elsevier, Amsterdam; Boston, 2006, pp. 827–947. [36] A.D. Coviello, W.J. Bremner, A.M. Matsumoto, K.L. Herbst, J.K. Amory, B.D.
[7] Y. Clermont, The cycle of the seminiferous epithelium in man, Am. J. Anat. Anawalt, et al., Intratesticular testosterone concentrations comparable with
112 (1963) 35–51. serum levels are not sufficient to maintain normal sperm production in men
[8] L. Hermo, M. Lalli, Y. Clermont, Arrangement of connective tissue receiving a hormonal contraceptive regimen, J. Androl. 25 (2004)
components in the walls of seminiferous tubules of man and monkey, Am. J. 931–938.
Anat. 148 (1977) 433–445. [37] A.D. Coviello, A.M. Matsumoto, W.J. Bremner, K.L. Herbst, J.K. Amory, B.D.
[9] F. Rossi, A. Ferraresi, P. Romagni, L. Silvestroni, V. Santiemma, I.I. Anawalt, et al., Low-dose human chorionic gonadotropin maintains
Angiotensin, stimulates contraction and growth of testicular peritubular intratesticular testosterone in normal men with testosterone-induced
myoid cells in vitro, Endocrinology 143 (2002) 3096–3104. gonadotropin suppression, J. Clin. Endocrinol. Metabol. 90 (2005)
[10] C. Schell, M. Albrecht, S. Spillner, C. Mayer, L. Kunz, F.M. Kohn, et al., 2595–2602.
15-Deoxy-delta 12-14-prostaglandin-J2 induces hypertrophy and loss of [38] E. Shinjo, K. Shiraishi, H. Matsuyama, The effect of human chorionic
contractility in human testicular peritubular cells: implications for human gonadotropin-based hormonal therapy on intratesticular testosterone levels
male fertility, Endocrinology 151 (2010) 1257–1268. and spermatogonial DNA synthesis in men with non-obstructive
[11] S.B. Cigorraga, H. Chemes, E. Pellizzari, Steroidogenic and morphogenic azoospermia, Andrology 1 (2013) 929–935.
characteristics of human peritubular cells in culture, Biol. Reprod. 51 (1994) [39] J.P. Jarow, B.R. Zirkin, The androgen microenvironment of the human testis
1193–1205. and hormonal control of spermatogenesis, Ann. N. Y. Acad. Sci. 1061 (2005)
[12] G. Verhoeven, E. Hoeben, K. De Gendt, Peritubular cell-Sertoli cell 208–220.
interactions: factors involved in PmodS activity, Andrologia 32 (2000) [40] A.H. Payne, R.P. Kelch, S.S. Musich, M.E. Halpern, Intratesticular site of
42–45. aromatization in the human, J. Clin. Endocrinol. Metabol. 42 (1976)
[13] C. Zhang, S. Yeh, Y.T. Chen, C.C. Wu, K.H. Chuang, H.Y. Lin, et al., 1081–1087.
Oligozoospermia with normal fertility in male mice lacking the androgen
F.T.L. Neto et al. / Seminars in Cell & Developmental Biology 59 (2016) 10–26 21

[41] M.C. Lardone, P. Castillo, R. Valdevenito, M. Ebensperger, A.M. Ronco, R. The Sertoli Cell, 1st ed., Cache River Press, Clearwater, FL, 1993,
Pommer, et al., P450-aromatase activity and expression in human testicular pp. 305–329.
tissues with severe spermatogenic failure, Int. J. Androl. 33 (2010) 650–660. [70] C.Y. Cheng, E.W. Wong, H.H. Yan, D.D. Mruk, Regulation of spermatogenesis
[42] S. Carreau, S. Lambard, C. Delalande, I. Denis-Galeraud, B. Bilinska, S. in the microenvironment of the seminiferous epithelium: new insights and
Bourguiba, Aromatase expression and role of estrogens in male gonad: a advances, Mol. Cell. Endocrinol. 315 (2010) 49–56.
review, Reprod. Biol. Endocrinol. 1 (2003) 35. [71] A.G. Byskov, Differentiation of mammalian embryonic gonad, Physiol. Rev.
[43] R.A. Hess, Estrogen in the adult male reproductive tract: a review, Reprod. 66 (1986) 71–117.
Biol. Endocrinol. 1 (2003) 52. [72] S.C. Munger, A. Natarajan, L.L. Looger, U. Ohler, B. Capel, Fine time course
[44] K.L. Matthiesson, P.G. Stanton, L. O’Donnell, S.J. Meachem, J.K. Amory, R. expression analysis identifies cascades of activation and repression and
Berger, et al., Effects of testosterone and levonorgestrel combined with a maps a putative regulator of mammalian sex determination, PLoS Genet. 9
5alpha-reductase inhibitor or gonadotropin-releasing hormone antagonist (2013) e1003630.
on spermatogenesis and intratesticular steroid levels in normal men, J. Clin. [73] A.H. Sinclair, P. Berta, M.S. Palmer, J.R. Hawkins, B.L. Griffiths, M.J. Smith,
Endocrinol. Metabol. 90 (2005) 5647–5655. et al., A gene from the human sex-determining region encodes a protein
[45] J.I. Park, J. Semyonov, C.L. Chang, W. Yi, W. Warren, S.Y. Hsu, Origin of with homology to a conserved DNA-binding motif, Nature 346 (1990)
INSL3-mediated testicular descent in therian mammals, Genome Res. 18 240–244.
(2008) 974–985. [74] C. Larney, T.L. Bailey, P. Koopman, Switching on sex: transcriptional
[46] K. Bay, S. Hartung, R. Ivell, M. Schumacher, D. Jurgensen, N. Jorgensen, et al., regulation of the testis-determining gene Sry, Development 141 (2014)
Insulin-like factor 3 serum levels in 135 normal men and 85 men with 2195–2205.
testicular disorders: relationship to the luteinizing hormone-testosterone [75] N. Warr, A. Greenfield, The molecular and cellular basis of gonadal sex
axis, J. Clin. Endocrinol. Metabol. 90 (2005) 3410–3418. reversal in mice and humans, Wiley Interdiscipl. Rev. Dev. Biol. 1 (2012)
[47] K. Bay, R. Anand-Ivell, Human testicular insulin-like factor 3 and endocrine 559–577.
disrupters, Vitam. Horm. 94 (2014) 327–348. [76] R.R. Behringer, M.J. Finegold, R.L. Cate, Mullerian-inhibiting substance
[48] C. Foresta, D. Zuccarello, A. Garolla, A. Ferlin, Role of hormones, genes, and function during mammalian sexual development, Cell 79 (1994) 415–425.
environment in human cryptorchidism, Endocr. Rev. 29 (2008) 560–580. [77] H.A. Ingraham, Y. Hirokawa, L.M. Roberts, S.H. Mellon, E. McGee, M.W.
[49] K. Kawamura, J. Kumagai, S. Sudo, S.Y. Chun, M. Pisarska, H. Morita, et al., Nachtigal, et al., Autocrine and paracrine Mullerian inhibiting substance
Paracrine regulation of mammalian oocyte maturation and male germ cell hormone signaling in reproduction, Recent Prog. Horm. Res. 55 (2000)
survival, Proc. Natl. Acad. Sci. U.S.A. 101 (2004) 7323–7328. 53–67 (discussion—8).
[50] C. Foresta, A. Bettella, C. Vinanzi, P. Dabrilli, M.C. Meriggiola, A. Garolla, et al., [78] M.M. Lee, P.K. Donahoe, T. Hasegawa, B. Silverman, G.B. Crist, S. Best, et al.,
A novel circulating hormone of testis origin in humans, J. Clin. Endocrinol. Mullerian inhibiting substance in humans: normal levels from infancy to
Metabol. 89 (2004) 5952–5958. adulthood, J. Clin. Endocrinol. Metabol. 81 (1996) 571–576.
[51] H. Thackare, H.D. Nicholson, K. Whittington, Oxytocin—its role in male [79] K. Morgan, N.A. Dennis, T. Ruffman, D.K. Bilkey, I.S. McLennan, The stature of
reproduction and new potential therapeutic uses, Hum. Reprod. Update 12 boys is inversely correlated to the levels of their sertoli cell hormones: do
(2006) 437–448. the testes restrain the maturation of boys? PLoS One 6 (2011) e20533.
[52] J.A. Mahoudeau, J.C. Valcke, H. Bricaire, Dissociated responses of plasma [80] Z. Herbert, S. Weigel, E. Sendemir, A. Marshall, J.D. Caldwell, P. Petrusz, et al.,
testosterone and estradiol to human chorionic gonadotropin in adult men, J. Androgen-binding protein is co-expressed with oxytocin in the male
Clin. Endocrinol. Metabol. 41 (1975) 13–20. reproductive tract, Anat. Histol. Embryol. J. Vet. Med. Series C 34 (2005)
[53] S.M. Eacker, N. Agrawal, K. Qian, H.L. Dichek, E.Y. Gong, K. Lee, et al., 286–293.
Hormonal regulation of testicular steroid and cholesterol homeostasis, Mol. [81] V. Hansson, O. Djøseland, E. Reusch, S. Aasvik, Preliminary characterization
Endocrinol. 22 (2008) 623–635. of the 5␣-dihydrotestosterone binding protein in the epididymal cytosol
[54] P. Tai, K. Shiraishi, M. Ascoli, Activation of the lutropin/choriogonadotropin fraction. In vivo studies, Acta Endocrinol. 71 (3) (1972) 614–624.
receptor inhibits apoptosis of immature Leydig cells in primary culture, [82] V. Santiemma, P. Rosati, C. Guerzoni, S. Mariani, F. Beligotti, M. Magnanti,
Endocrinology 150 (2009) 3766–3773. et al., Human Sertoli cells in vitro: morphological features and
[55] A.K. Christensen, K.C. Peacock, Increase in Leydig cell number in testes of androgen-binding protein secretion, J. Steroid Biochem. Mol. Biol. 43 (1992)
adult rats treated chronically with an excess of human chorionic 423–429.
gonadotropin, Biol. Reprod. 22 (1980) 383–391. [83] J. Della-Maria, A. Gerard, P. Franck, H. Gerard, Effects of androgen-binding
[56] K. Purvis, O.P. Clausen, V. Hansson, Androgen effects on rat Leydig cells, Biol. protein (ABP) on spermatid Tnp1 gene expression in vitro, Mol. Cell.
Reprod. 20 (1979) 304–309. Endocrinol. 198 (2002) 131–141.
[57] R.B. White, J.A. Eisen, T.L. Kasten, R.D. Fernald, Second gene for [84] R. Robinson, I.B. Fritz, Metabolism of glucose by Sertoli cells in culture, Biol.
gonadotropin-releasing hormone in humans, Proc. Natl. Acad. Sci. U.S.A. 95 Reprod. 24 (1981) 1032–1041.
(1998) 305–309. [85] D.D. Mruk, C.Y. Cheng, Sertoli–Sertoli and Sertoli-germ cell interactions and
[58] R.M. Sharpe, H.M. Fraser, I. Cooper, F.F. Rommerts, Sertoli-Leydig cell their significance in germ cell movement in the seminiferous epithelium
communication via an LHRH-like factor, Nature 290 (1981) 785–787. during spermatogenesis, Endocr. Rev. 25 (2004) 747–806.
[59] S.P. Rossi, S. Windschuettl, M.E. Matzkin, C. Terradas, R. Ponzio, E. [86] C.A. Hogarth, M.D. Griswold, The key role of vitamin A in spermatogenesis, J.
Puigdomenech, et al., Melatonin in testes of infertile men: evidence for Clin. Invest. 120 (2010) 956–962.
anti-proliferative and anti-oxidant effects on local macrophage and mast [87] F. Boussouar, M. Benahmed, Lactate and energy metabolism in male germ
cell populations, Andrology 2 (2014) 436–449. cells, Trends Endocrinol. Metab. 15 (2004) 345–350.
[60] S.P. Rossi, M.E. Matzkin, C. Terradas, R. Ponzio, E. Puigdomenech, O. Levalle, [88] L. Rato, M.G. Alves, S. Socorro, A.I. Duarte, J.E. Cavaco, P.F. Oliveira, Metabolic
et al., New insights into melatonin/CRH signaling in hamster Leydig cells, regulation is important for spermatogenesis, Nat. Rev. Urol. 9 (2012)
Gen. Comp. Endocrinol. 178 (2012) 153–163. 330–338.
[61] V. Syed, S.A. Khan, E. Nieschlag, Epidermal growth factor stimulates [89] M.G. Alves, S. Socorro, J. Silva, A. Barros, M. Sousa, J.E. Cavaco, et al., In vitro
testosterone production of human Leydig cells in vitro, J. Endocrinol. Invest. cultured human Sertoli cells secrete high amounts of acetate that is
14 (1991) 93–97. stimulated by 17beta-estradiol and suppressed by insulin deprivation,
[62] G. Lucio, F. Andrea, S. Giovanni, Gonadal peptides as mediators of Biochim. Biophys. Acta 1823 (2012) 1389–1394.
development and functional control of the testis: an integrated system with [90] A. Meinhardt, M.P. Hedger, Immunological, paracrine and endocrine aspects
hormones and local environment, Endocr. Rev. 18 (1997) 541–609. of testicular immune privilege, Mol. Cell. Endocrinol. 335 (2011) 60–68.
[63] T. Ishikawa, H. Fujioka, T. Ishimura, A. Takenaka, M. Fujisawa, Ghrelin [91] C. Cudicini, H. Lejeune, E. Gomez, E. Bosmans, F. Ballet, J. Saez, et al., Human
expression in human testis and serum testosterone level, J. Androl. 28 Leydig cells and Sertoli cells are producers of interleukins-1 and -6, J. Clin.
(2007) 320–324. Endocrinol. Metabol. 82 (1997) 1426–1433.
[64] C. Geigerseder, R. Doepner, A. Thalhammer, M.B. Frungieri, K. [92] G.F. Weinbauer, J. Wessels, ‘Paracrine’ control of spermatogenesis,
Gamel-Didelon, R.S. Calandra, et al., Evidence for a GABAergic system in Andrologia 31 (1999) 249–262.
rodent and human testis: local GABA production and GABA receptors, [93] Y. Hai, J. Hou, Y. Liu, Y. Liu, H. Yang, Z. Li, et al., The roles and regulation of
Neuroendocrinology 77 (2003) 314–323. Sertoli cells in fate determinations of spermatogonial stem cells and
[65] C. Geigerseder, R.F.G. Doepner, A. Thalhammer, A. Krieger, A. Mayerhofer, spermatogenesis, Semin. Cell Dev. Biol. 29 (2014) 66–75.
Stimulation of TM3 Leydig cell proliferation via GABA(A)receptors: a new [94] H.G. Burger, Inhibin: definition and nomenclature, including related
role for testicular GABA, Reprod. Biol. Endocrinol. 2 (2004) 13. substances, J. Endocrinol. 117 (1988) 159–160.
[66] L.D. Russell, H.P. Ren, I. Sinha Hikim, W. Schulze, A.P. Sinha Hikim, A [95] A.M. Andersson, Inhibin B in the assessment of seminiferous tubular
comparative study in twelve mammalian species of volume densities, function, Baillieres Best Pract. Res. Clin. Endocrinol. Metab. 14 (2000)
volumes, and numerical densities of selected testis components, 389–397.
emphasizing those related to the Sertoli cell, Am. J. Anat. 188 (1990) 21–30. [96] C. Marchetti, M. Hamdane, V. Mitchell, K. Mayo, L. Devisme, J.M. Rigot, et al.,
[67] C. Schulze, A.F. Holstein, C. Schirren, F. Korner, On the morphology of the Immunolocalization of inhibin and activin alpha and betaB subunits and
human Sertoli cells under normal conditions and in patients with impaired expression of corresponding messenger RNAs in the human adult testis,
fertility, Andrologia 8 (1976) 167–178. Biol. Reprod. 68 (2003) 230–235.
[68] L.L. Ewing, J.C. Davis, B.R. Zirkin, Regulation of testicular function: a spatial [97] F.H. Pierik, J.T.M. Vreeburg, T. Stijnen, F.H. de Jong, R.F.A. Weber, Serum
and temporal view, Int. Rev. Physiol. 22 (1980) 41–115. inhibin B as a marker of spermatogenesis, J. Clin. Endocrinol. Metabol. 83
[69] C. Morales, Y. Clermont, Structural changes of the Sertoli cell during the (1998) 3110–3114.
cycle of the seminiferous epithelium, in: L.D. Russell, M.D. Griswold (Eds.),
22 F.T.L. Neto et al. / Seminars in Cell & Developmental Biology 59 (2016) 10–26

[98] K.A. Lewis, P.C. Gray, A.L. Blount, L.A. MacConnell, E. Wiater, L.M. Bilezikjian, [126] R. Lambrot, H. Coffigny, C. Pairault, A.C. Donnadieu, R. Frydman, R. Habert,
et al., Betaglycan binds inhibin and can mediate functional antagonism of et al., Use of organ culture to study the human fetal testis development:
activin signalling, Nature 404 (2000) 411–414. effect of retinoic acid, J. Clin. Endocrinol. Metabol. 91 (2006) 2696–2703.
[99] M.M. Matzuk, M.J. Finegold, Y. Mishina, A. Bradley, R.R. Behringer, [127] Culty M. Gonocytes, the forgotten cells of the germ cell lineage Birth defects
Synergistic effects of inhibins and mullerian-inhibiting substance on research Part C, Embryo Today: Rev. 87 (2009) 1–26.
testicular tumorigenesis, Mol. Endocrinol. 9 (1995) 1337–1345. [128] C.G. Print, K.L. Loveland, Germ cell suicide: new insights into apoptosis
[100] S.J. Meachem, E. Nieschlag, M. Simoni, Inhibin B in male reproduction: during spermatogenesis, BioEssays 22 (2000) 423–430.
pathophysiology and clinical relevance, Eur. J. Endocrinol./Eur. Fed. Endocr. [129] R. Paniagua, M. Nistal, Morphological and histometric study of human
Soc. 145 (2001) 561–571. spermatogonia from birth to the onset of puberty, J. Anat. 139 (Pt 3) (1984)
[101] A. Nakagawa, A. Shiratsuchi, K. Tsuda, Y. Nakanishi, In vivo analysis of 535–552.
phagocytosis of apoptotic cells by testicular Sertoli cells, Mol. Reprod. Dev. [130] Y. Clermont, Renewal of spermatogonia in man, Am. J. Anat. 118 (1966)
71 (2005) 166–177. 509–524.
[102] I. Carr, E.J. Clegg, G.A. Meek, Sertoli cells as phagocytes: an electron [131] E. Goossens, H. Tournaye, Adult stem cells in the human testis, Semin.
microscopic study, J. Anat. 102 (1968) 501–509. Reprod. Med. 31 (2013) 39–48.
[103] Y. Nakanishi, A. Shiratsuchi, Phagocytic removal of apoptotic spermatogenic [132] Y. Clermont, Spermatogenesis in man: a study of the spermatogonial
cells by Sertoli cells: mechanisms and consequences, Biol. Pharm. Bull. 27 population, Fertil. Steril. 17 (1966) 705–721.
(2004) 13–16. [133] M.M. van Alphen, H.J. van de Kant, D.G. de Rooij, Repopulation of the
[104] R.I. McLachlan, L. O’Donnell, S.J. Meachem, P.G. Stanton, D.M. de Kretser, K. seminiferous epithelium of the rhesus monkey after X irradiation, Radiat.
Pratis, et al., Identification of specific sites of hormonal regulation in Res. 113 (1988) 487–500.
spermatogenesis in rats, monkeys, and man, Recent Prog. Horm. Res. 57 [134] D.G. de Rooij, L.D. Russell, All you wanted to know about spermatogonia but
(2002) 149–179. were afraid to ask, J. Androl. 21 (2000) 776–798.
[105] W.H. Walker, J. Cheng, FSH and testosterone signaling in Sertoli cells, [135] J. Ehmcke, S. Schlatt, A revised model for spermatogonial expansion in man:
Reproduction 130 (2005) 15–28. lessons from non-human primates, Reproduction 132 (2006) 673–680.
[106] T.M. Plant, G.R. Marshall, The functional significance of FSH in [136] L. Grisanti, I. Falciatori, M. Grasso, L. Dovere, S. Fera, B. Muciaccia, et al.,
spermatogenesis and the control of its secretion in male primates, Endocr. Identification of spermatogonial stem cell subsets by morphological analysis
Rev. 22 (2001) 764–786. and prospective isolation, Stem cells (Dayton, Ohio) 27 (2009) 3043–3052.
[107] P.J. O’Shaughnessy, H. Johnston, L. Willerton, P.J. Baker, Failure of normal [137] T. Hayashi, T. Yamada, Y. Kageyama, T. Negishi, K. Kihara, Expression failure
adult Leydig cell development in androgen-receptor-deficient mice, J. Cell of the Notch signaling system is associated with the pathogenesis of
Sci. 115 (2002) 3491–3496. maturation arrest in male infertility patients, Fertil. Steril. 81 (2004)
[108] M.J. Tsai, B.W. O’Malley, Molecular mechanisms of action of steroid/thyroid 697–699.
receptor superfamily members, Annu. Rev. Biochem 63 (1994) 451–486. [138] S. van den Driesche, R.M. Sharpe, P.T. Saunders, R.T. Mitchell, Regulation of
[109] S. RM, Regulation of spermatogenesis, in: E. Knobil, J.D. Neill (Eds.), The the germ stem cell niche as the foundation for adult spermatogenesis: a role
Physiology of Reproduction, 2nd ed., Raven Press, New York, 1994, pp. for miRNAs? Semin. Cell Dev. Biol. 29 (2014) 76–83.
1363–1434. [139] L. Johnson, C.S. Petty, J.C. Porter, W.B. Neaves, Germ cell degeneration during
[110] P.I. Sadate-Ngatchou, D.J. Pouchnik, M.D. Griswold, Identification of postprophase of meiosis and serum concentrations of gonadotropins in
testosterone-regulated genes in testes of hypogonadal mice using young adult and older adult men, Biol. Reprod. 31 (1984) 779–784.
oligonucleotide microarray, Mol. Endocrinol. 18 (2004) 422–433. [140] S.K. Kim, Y.D. Yoon, Y.S. Park, J.T. Seo, J.H. Kim, Involvement of the Fas–Fas
[111] G. Verhoeven, E. Denolet, J.V. Swinnen, A. Willems, F. Claessens, P.T.K. ligand system and active caspase-3 in abnormal apoptosis in human testes
Saunders, R.M. Sharpe, K. De Gendt, Contribution of recent transgenic with maturation arrest and Sertoli cell-only syndrome, Fertil. Steril. 87
models and transcriptional profiling studies to our understanding of the (2007) 547–553.
mechanisms by which androgens control spermatogenesis immunology, [141] D.A. O’Neill, C.M. McVicar, N. McClure, P. Maxwell, I. Cooke, K.M. Pogue,
Endocr. Metabol. Agents Med. Chem. 8 (2008) 2–13. et al., Reduced sperm yield from testicular biopsies of vasectomized men is
[112] F.R. Silva, L.D. Leite, G.F. Wassermann, Rapid signal transduction in Sertoli due to increased apoptosis, Fertil. Steril. 87 (2007) 834–841.
cells, Eur. J. Endocrinol./Eur. Fed. Endocr. Soc. 147 (2002) 425–433. [142] L. Hermo, R.M. Pelletier, D.G. Cyr, C.E. Smith, Surfing the wave, cycle, life
[113] A.C. Cato, A. Nestl, S. Mink, Rapid actions of steroid receptors in cellular history, and genes/proteins expressed by testicular germ cells. Part 4:
signaling pathways, Sci. STKE: Signal Transduct. Knowl. Environ. 2002 intercellular bridges, mitochondria, nuclear envelope, apoptosis,
(2002) re9. ubiquitination, membrane/voltage-gated channels, methylation/acetylation,
[114] P.F. Oliveira, M.G. Alves, L. Rato, S. Laurentino, J. Silva, R. Sa, et al., Effect of and transcription factors, Microsc. Res. Tech. 73 (2010) 364–408.
insulin deprivation on metabolism and metabolism-associated gene [143] S.M. Ruwanpura, R.I. McLachlan, K.L. Matthiesson, S.J. Meachem,
transcript levels of in vitro cultured human Sertoli cells, Biochim. Biophys. Gonadotrophins regulate germ cell survival, not proliferation, in normal
Acta 1820 (2012) 84–89. adult men, Hum. Reprod. 23 (2008) 403–411.
[115] C.S. Rocha, A.D. Martins, L. Rato, B.M. Silva, P.F. Oliveira, M.G. Alves, [144] V. Pentikainen, L. Suomalainen, K. Erkkila, E. Martelin, M. Parvinen, M.O.
Melatonin alters the glycolytic profile of Sertoli cells: implications for male Pentikainen, et al., Nuclear factor-kappa B activation in human testicular
fertility, Mol. Hum. Reprod. 20 (2014) 1067–1076. apoptosis, Am. J. Pathol. 160 (2002) 205–218.
[116] M. Rosselli, M.K. Skinner, Developmental regulation of Sertoli cell aromatase [145] L. Johnson, P.K. Chaturvedi, J.D. Williams, Missing generations of
activity and plasminogen activator production by hormones, retinoids and spermatocytes and spermatids in seminiferous epithelium contribute to low
the testicular paracrine factor, PModS, Biol. Reprod. 46 (1992) 586–594. efficiency of spermatogenesis in humans, Biol. Reprod. 47 (1992)
[117] de Rooij D. vD-E.F, Regulation of proliferation and differentiation of stem 1091–1098.
cells in the male germ line, in: C.S. Potten (Ed.), Stem Cells (Dayton, Ohio), [146] B. Muciaccia, C. Boitani, B.P. Berloco, F. Nudo, G. Spadetta, M. Stefanini, et al.,
Academic Press, London; San Diego, 1997, pp. 283–313. Novel stage classification of human spermatogenesis based on acrosome
[118] S.F. Gilbert, Developmental Biology. PubMed Bookshelf, 6th ed., Sinauer development, Biol. Reprod. 89 (2013) 60.
Associates, Sunderland, Mass. Bethesda, MD, 2016 (NCBI,; 2000. p. xviii, 749 [147] W. Schulze, U. Rehder, Organization and morphogenesis of the human
pages). seminiferous epithelium, Cell Tissue Res. 237 (1984) 395–407.
[119] L.L. Robinson, T.L. Gaskell, P.T. Saunders, R.A. Anderson, Germ cell specific [148] W. Schulze, M. Riemer, U. Rehder, K.H. Hohne, Computer-aided
expression of c-kit in the human fetal gonad, Mol. Hum. Reprod. 7 (2001) three-dimensional reconstructions of the arrangement of primary
845–852. spermatocytes in human seminiferous tubules, Cell Tissue Res. 244 (1986)
[120] J. del Mazo, J. Garcia-Lopez, M. Weber, Epigenetic traits of testicular cancer: 1–7.
from primordial germ cells to germ cell tumors, Epigenomics 6 (2014) [149] L. Johnson, K.S. McKenzie, J.R. Snell, Partial wave in human seminiferous
253–255. tubules appears to be a random occurrence, Tissue Cell 28 (1996) 127–136.
[121] M.I. Figueira, H.J. Cardoso, S. Correia, C.J. Maia, S. Socorro, Hormonal [150] E.B. Del Castillo, A. Trabucco, F.A. D.ElB, Syndrome produced by absence of
regulation of c-KIT receptor and its ligand: implications for human the germinal epithelium without impairment of the Sertoli or Leydig cells, J.
infertility, Prog. Histochem. Cytochem. 49 (2014) 1–19. Clin. Endocrinol. Metabol. 7 (1947) 493–502.
[122] D.C. Gilbert, I. Chandler, A. McIntyre, N.C. Goddard, R. Gabe, R.A. Huddart, [151] R. Anniballo, F. Ubaldi, L. Cobellis, M. Sorrentino, L. Rienzi, E. Greco, et al.,
et al., Clinical and biological significance of CXCL12 and CXCR4 expression in Criteria predicting the absence of spermatozoa in the Sertoli cell-only
adult testes and germ cell tumours of adults and adolescents, J. Pathol. 217 syndrome can be used to improve success rates of sperm retrieval, Hum.
(2009) 94–102. Reprod. 15 (2000) 2269–2277.
[123] K. Pauls, R. Jager, S. Weber, E. Wardelmann, A. Koch, R. Buttner, et al., [152] H. Tournaye, J. Liu, P.Z. Nagy, M. Camus, A. Goossens, S. Silber, et al.,
Transcription factor AP-2gamma, a novel marker of gonocytes and Correlation between testicular histology and outcome after
seminomatous germ cell tumors, Int. J. Cancer 115 (2005) 470–477. intracytoplasmic sperm injection using testicular spermatozoa, Hum.
[124] A.J. Levine, A.H. Brivanlou, GDF3, a BMP inhibitor, regulates cell fate in stem Reprod. 11 (1996) 127–132.
cells and early embryos, Development 133 (2006) 209–216. [153] B.B. Najari, R. Ramasamy, J. Sterling, A. Aggarwal, S. Sheth, P.S. Li, et al., Pilot
[125] V. Pentikainen, K. Erkkila, L. Suomalainen, M. Parvinen, L. Dunkel, Estradiol study of the correlation of multiphoton tomography of ex vivo human testis
acts as a germ cell survival factor in the human testis in vitro, J. Clin. with histology, J. Urol. 188 (2012) 538–543.
Endocrinol. Metabol. 85 (2000) 2057–2067. [154] Y. Liu, Y. Zhu, L. Di, E.C. Osterberg, F. Liu, L. He, et al., Raman spectroscopy as
an ex vivo noninvasive approach to distinguish complete and incomplete
F.T.L. Neto et al. / Seminars in Cell & Developmental Biology 59 (2016) 10–26 23

spermatogenesis within human seminiferous tubules, Fertil. Steril. 102 [181] E.M. Eddy, K. Toshimori, D.A. O’Brien, Fibrous sheath of mammalian
(2014), 54 e2–60 e2. spermatozoa, Microsc. Res. Tech. 61 (2003) 103–115.
[155] J. Gonsalves, F. Sun, P.N. Schlegel, P.J. Turek, C.V. Hopps, C. Greene, et al., [182] A. Amaral, J. Castillo, J.M. Estanyol, J.L. Ballesca, J. Ramalho-Santos, R. Oliva,
Defective recombination in infertile men, Hum. Mol. Genet. 13 (2004) Human sperm tail proteome suggests new endogenous metabolic pathways,
2875–2883. Mol. Cell. Proteom.: MCP 12 (2013) 330–342.
[156] R. Ramasamy, J.M. Scovell, J.R. Kovac, P.J. Cook, D.J. Lamb, L.I. Lipshultz, [183] H. Breucker, E. Schafer, A.F. Holstein, Morphogenesis and fate of the residual
Fluorescence in situ hybridization detects increased sperm aneuploidy in body in human spermiogenesis, Cell Tissue Res. 240 (1985) 303–309.
men with recurrent pregnancy loss, Fertil. Steril. 103 (2015), 906 e1–909 e1. [184] A. Abdel Raheem, G. Garaffa, N. Rushwan, F. De Luca, E. Zacharakis, T. Abdel
[157] F. Sun, M. Mikhaail-Philips, M. Oliver-Bonet, E. Ko, A. Rademaker, P. Turek, Raheem, et al., Testicular histopathology as a predictor of a positive sperm
et al., Reduced meiotic recombination on the XY bivalent is correlated with retrieval in men with non-obstructive azoospermia, BJU Int. 111 (2013)
an increased incidence of sex chromosome aneuploidy in men with 492–499.
non-obstructive azoospermia, Mol. Hum. Reprod. 14 (2008) 399–404. [185] R. Ramasamy, N. Yagan, P.N. Schlegel, Structural and functional changes to
[158] F. Yang, K. Gell, G.W. van der Heijden, S. Eckardt, N.A. Leu, D.C. Page, et al., the testis after conventional versus microdissection testicular sperm
Meiotic failure in male mice lacking an X-linked factor, Genes Dev. 22 extraction, Urology 65 (2005) 1190–1194.
(2008) 682–691. [186] X. Qian, D.D. Mruk, Y.H. Cheng, E.I. Tang, D. Han, W.M. Lee, et al., Actin
[159] C.A. Adelman, J.H. Petrini, ZIP4H (TEX11) deficiency in the mouse impairs binding proteins, spermatid transport and spermiation, Semin. Cell Dev.
meiotic double strand break repair and the regulation of crossing over, PLoS Biol. 30 (2014) 75–85.
Genet. 4 (2008) e1000042. [187] L. O’Donnell, P.K. Nicholls, M.K. O’Bryan, R.I. McLachlan, P.G. Stanton,
[160] A.N. Yatsenko, A.P. Georgiadis, A. Ropke, A.J. Berman, T. Jaffe, M. Olszewska, Spermiation: the process of sperm release, Spermatogenesis 1 (2011) 14–35.
et al., X-linked TEX11 mutations, meiotic arrest, and azoospermia in infertile [188] C.Y. Cheng, D.D. Mruk, The blood-testis barrier and its implications for male
men, N. Engl. J. Med. 372 (2015) 2097–2107. contraception, Pharmacol. Rev. 64 (2012) 16–64.
[161] J. Tesarik, B. Balaban, A. Isiklar, C. Alatas, B. Urman, S. Aksoy, et al., In-vitro [189] L.L. Mitic, J.M. Anderson, Molecular architecture of tight junctions, Annu.
spermatogenesis resumption in men with maturation arrest: relationship Rev. Physiol. 60 (1998) 121–142.
with in-vivo blocking stage and serum FSH, Hum. Reprod. 15 (2000) [190] M. Dym, The fine structure of the monkey (Macaca) Sertoli cell and its role
1350–1354. in maintaining the blood-testis barrier, Anat. Record 175 (1973) 639–656.
[162] N. Sofikitis, T. Mantzavinos, D. Loutradis, Y. Yamamoto, V. Tarlatzis, I. [191] B.P. Setchell, Blood-testis barrier, junctional and transport proteins and
Miyagawa, Ooplasmic injections of secondary spermatocytes for spermatogenesis, Adv. Exp. Med. Biol. 636 (2008) 212–233.
non-obstructive azoospermia, Lancet (Lond. Engl.) 351 (1998) 1177–1178. [192] M. Dym, J.C. Cavicchia, Further observations on the blood-testis barrier in
[163] R.I. McLachlan, E. Rajpert-De Meyts, C.E. Hoei-Hansen, D.M. de Kretser, N.E. monkeys, Biol. Reprod. 17 (1977) 390–403.
Skakkebaek, Histological evaluation of the human testis–approaches to [193] T.D. Yule, G.D. Montoya, L.D. Russell, T.M. Williams, K.S. Tung, Autoantigenic
optimizing the clinical value of the assessment: mini review, Hum. Reprod. germ cells exist outside the blood testis barrier, J. Immunol. 141 (1988)
22 (2007) 2–16. 1161–1167 (Baltimore, Md: 1950).
[164] A.M. Bernie, K. Shah, J.A. Halpern, J. Scovell, R. Ramasamy, B. Robinson, et al., [194] C.A. Mahi-Brown, T.D. Yule, K.S. Tung, Evidence for active immunological
Outcomes of microdissection testicular sperm extraction in men with regulation in prevention of testicular autoimmune disease independent of
nonobstructive azoospermia due to maturation arrest, Fertil. Steril. 104 the blood-testis barrier, Am. J. Reprod. Immunol. Microbiol.: AJRIM 16
(2015), 569 e1–573 e1. (1988) 165–170.
[165] J.W. Weedin, R.C. Bennett, D.M. Fenig, D.J. Lamb, L.I. Lipshultz, Early versus [195] G. Kaur, L.A. Thompson, J.M. Dufour, Sertoli cells–immunological sentinels of
late maturation arrest: reproductive outcomes of testicular failure, J. Urol. spermatogenesis, Semin. Cell Dev. Biol. 30 (2014) 36–44.
186 (2011) 621–626. [196] X. Xiao, D.D. Mruk, F.L. Cheng, C.Y. Cheng, C-Src and c-Yes are two unlikely
[166] M.C. Tsai, Y.S. Cheng, T.Y. Lin, W.H. Yang, Y.M. Lin, Clinical characteristics partners of spermatogenesis and their roles in blood-testis barrier
and reproductive outcomes in infertile men with testicular early and late dynamics, Adv. Exp. Med. Biol. 763 (2012) 295–317.
maturation arrest, Urology 80 (2012) 826–832. [197] J. Bart, H. Hollema, H.J. Groen, E.G. de Vries, N.H. Hendrikse, D.T. Sleijfer,
[167] T.R. Dias, M.G. Alves, B.M. Silva, P.F. Oliveira, Sperm glucose transport and et al., The distribution of drug-efflux pumps, P-gp, BCRP, MRP1 and MRP2, in
metabolism in diabetic individuals, Mol. Cell. Endocrinol. 396 (2014) 37–45. the normal blood-testis barrier and in primary testicular tumours, Eur. J.
[168] P.F. Oliveira, M.G. Alves, L. Rato, J. Silva, R. Sa, A. Barros, et al., Influence of Cancer 40 (2004) 2064–2070 (Oxford, Engl.: 1990).
5alpha-dihydrotestosterone and 17beta-estradiol on human Sertoli cells [198] D.D. Mruk, L. Su, C.Y. Cheng, Emerging role for drug transporters at the
metabolism, Int. J. Androl. 34 (2011) e612–20. blood-testis barrier, Trends Pharmacol. Sci. 32 (2011) 99–106.
[169] Y. Clermont, C.P. Leblond, Spermiogenesis of man, monkey, ram and other [199] N. Ilani, N. Armanious, Y.H. Lue, R.S. Swerdloff, S. Baravarian, A. Adler, et al.,
mammals as shown by the periodic acid-Schiff technique, Am. J. Anat. 96 Integrity of the blood-testis barrier in healthy men after suppression of
(1955) 229–253. spermatogenesis with testosterone and levonorgestrel, Hum. Reprod. 27
[170] R. Oko, P. Sutovsky, Biogenesis of sperm perinuclear theca and its role in (2012) 3403–3411.
sperm functional competence and fertilization, J. Reprod. Immunol. 83 [200] W.H. Nah, J.E. Lee, H.J. Park, N.C. Park, M.C. Gye, Claudin-11 expression
(2009) 2–7. increased in spermatogenic defect in human testes, Fertil. Steril. 95 (2011)
[171] C. Alvarez Sedo, V.Y. Rawe, H.E. Chemes, Acrosomal biogenesis in human 385–388.
globozoospermia: immunocytochemical, ultrastructural and proteomic [201] G.V. Landon, J.P. Pryor, The blood-testis barrier in men of diverse fertility
studies, Hum. Reprod. 27 (2012) 1912–1921. status: an ultrastructural study, Virchows Archiv. A Pathol. Anat. Histol. 392
[172] V. Pierre, G. Martinez, C. Coutton, J. Delaroche, S. Yassine, C. Novella, et al., (1981) 355–364.
Absence of Dpy19l2, a new inner nuclear membrane protein, causes [202] P.P. Lie, C.Y. Cheng, D.D. Mruk, Signalling pathways regulating the
globozoospermia in mice by preventing the anchoring of the acrosome to blood-testis barrier, Int. J. Biochem. Cell Biol. 45 (2013) 621–625.
the nucleus, Development 139 (2012) 2955–2965. [203] WHO, Obesity and Overwight, WHO, 2015.
[173] I. Koscinski, E. Elinati, C. Fossard, C. Redin, J. Muller, J. Velez de la Calle, et al., [204] C.L. Ogden, M.D. Carroll, B.K. Kit, K.M. Flegal, Prevalence of childhood and
DPY19L2 deletion as a major cause of globozoospermia, Am. J. Hum. Genet. adult obesity in the United States, 2011–2012, JAMA 311 (2014) 806–814.
88 (2011) 344–350. [205] S.H. Swan, E.P. Elkin, L. Fenster, The question of declining sperm density
[174] P. Kuentz, F. Vanden Meerschaut, E. Elinati, M.H. Nasr-Esfahani, T. Gurgan, N. revisited: an analysis of 101 studies published 1934–1996, Environ. Health
Iqbal, et al., Assisted oocyte activation overcomes fertilization failure in Perspect. 108 (2000) 961–966.
globozoospermic patients regardless of the DPY19L2 status, Hum. Reprod. [206] M. Cocuzza, S.C. Esteves, Shedding light on the controversy surrounding the
28 (2013) 1054–1061. temporal decline in human sperm counts: a systematic review, Sci. World J.
[175] N. Tanphaichitr, P. Sobhon, N. Taluppeth, P. Chalermisarachai, Basic nuclear 2014 (2014) 365691.
proteins in testicular cells and ejaculated spermatozoa in man, Exp. Cell Res. [207] A.O. Hammoud, M. Gibson, C.M. Peterson, B.D. Hamilton, D.T. Carrell,
117 (1978) 347–356. Obesity and male reproductive potential, J. Androl. 27 (2006) 619–626.
[176] C.C. Conwell, I.D. Vilfan, N.V. Hud, Controlling the size of nanoscale toroidal [208] E.V. Magnusdottir, T. Thorsteinsson, S. Thorsteinsdottir, M. Heimisdottir, K.
DNA condensates with static curvature and ionic strength, Proc. Natl. Acad. Olafsdottir, Persistent organochlorines, sedentary occupation, obesity and
Sci. U.S.A. 100 (2003) 9296–9301. human male subfertility, Hum. Reprod. 20 (2005) 208–215.
[177] K. Biermann, K. Steger, Epigenetics in male germ cells, J. Androl. 28 (2007) [209] S. Koloszar, I. Fejes, Z. Zavaczki, J. Daru, J. Szollosi, A. Pal, Effect of body
466–480. weight on sperm concentration in normozoospermic males, Arch. Androl.
[178] K.C. Kleene, Patterns, mechanisms, and functions of translation regulation in 51 (2005) 299–304.
mammalian spermatogenic cells, Cytogenet. Genome Res. 103 (2003) [210] T.K. Jensen, A.M. Andersson, N. Jorgensen, A.G. Andersen, E. Carlsen, J.H.
217–224. Petersen, et al., Body mass index in relation to semen quality and
[179] J.G. Sun, A. Jurisicova, R.F. Casper, Detection of deoxyribonucleic acid reproductive hormones among 1,558 Danish men, Fertil. Steril. 82 (2004)
fragmentation in human sperm: correlation with fertilization in vitro, Biol. 863–870.
Reprod. 56 (1997) 602–607. [211] H.I. Kort, J.B. Massey, C.W. Elsner, D. Mitchell-Leef, D.B. Shapiro, M.A. Witt,
[180] P.M.G. Sutovsky, Mammalian spermatogenesis and sperm structure: et al., Impact of body mass index values on sperm quantity and quality, J.
anatomical and compartmental analysis, in: C.J.D. Jonge, C.L.R. Barratt (Eds.), Androl. 27 (2006) 450–452.
The Sperm Cell: Production, Maturation, Fertilization, Regeneration, [212] S. La Vignera, R.A. Condorelli, E. Vicari, A.E. Calogero, Negative effect of
Cambridge University Press, Cambridge, UK; New York, 2006, pp. 1–30. increased body weight on sperm conventional and nonconventional flow
cytometric sperm parameters, J. Androl. 33 (2012) 53–58.
24 F.T.L. Neto et al. / Seminars in Cell & Developmental Biology 59 (2016) 10–26

[213] E.R. Hofny, M.E. Ali, H.Z. Abdel-Hafez, D. Kamal Eel, E.E. Mohamed, H.G. Abd [242] L. Piconi, L. Quagliaro, A. Ceriello, Oxidative stress in diabetes, Clin. Chem.
El-Azeem, et al., Semen parameters and hormonal profile in obese fertile Lab. Med.: CCLM/FESCC 41 (2003) 1144–1149.
and infertile males, Fertil. Steril. 94 (2010) 581–584. [243] C. Mallidis, I.M. Agbaje, D.A. Rogers, J.V. Glenn, R. Pringle, A.B. Atkinson,
[214] A.O. Hammoud, N. Wilde, M. Gibson, A. Parks, D.T. Carrell, A.W. Meikle, Male et al., Advanced glycation end products accumulate in the reproductive tract
obesity and alteration in sperm parameters, Fertil. Steril. 90 (2008) of men with diabetes, Int. J. Androl. 32 (2009) 295–305.
2222–2225. [244] J. Karimi, M.T. Goodarzi, H. Tavilani, I. Khodadadi, I. Amiri, Relationship
[215] J.E. Chavarro, T.L. Toth, D.L. Wright, J.D. Meeker, R. Hauser, Body mass index between advanced glycation end products and increased lipid peroxidation
in relation to semen quality, sperm DNA integrity, and serum reproductive in semen of diabetic men, Diabetes Res. Clin. Pract. 91 (2011) 61–66.
hormone levels among men attending an infertility clinic, Fertil. Steril. 93 [245] C. Mallidis, I. Agbaje, J. O’Neill, N. McClure, The influence of type 1 diabetes
(2010) 2222–2231. mellitus on spermatogenic gene expression, Fertil. Steril. 92 (2009)
[216] H. de Boer, L. Verschoor, J. Ruinemans-Koerts, M. Jansen, Letrozole 2085–2087.
normalizes serum testosterone in severely obese men with [246] J.O. Nriagu, Saturnine gout among Roman aristocrats. Did lead poisoning
hypogonadotropic hypogonadism, Diabetes Obesity Metabol. 7 (2005) contribute to the fall of the Empire? N. Engl. J. Med. 308 (1983) 660–663.
211–215. [247] M.A. Riva, A. Lafranconi, M.I. D’Orso, G. Cesana, Lead poisoning: historical
[217] G. Schneider, M.A. Kirschner, R. Berkowitz, N.H. Ertel, Increased estrogen aspects of a paradigmatic occupational and environmental disease, Saf.
production in obese men, J. Clin. Endocrinol. Metabol. 48 (1979) 633–638. Health Work 3 (2012) 11–16.
[218] H.G. Burger, Physiological principles of endocrine replacement: estrogen, [248] I. Lancranjan, H.I. Popescu, O. GAvănescu, I. Klepsch, M. Serbanescu,
Horm. Res. 56 (Suppl. 1) (2001) 82–85. Reproductive ability of workmen occupationally exposed to lead, Arch.
[219] V.A. Giagulli, J.M. Kaufman, A. Vermeulen, Pathogenesis of the decreased Environ. Health 30 (1975) 396–401.
androgen levels in obese men, J. Clin. Endocrinol. Metabol. 79 (1994) [249] R.S. Ge, G.R. Chen, C. Tanrikut, M.P. Hardy, Phthalate ester toxicity in Leydig
997–1000. cells: developmental timing and dosage considerations, Reprod. Toxicol.
[220] N.V. Gutorova, M.A. Kleshchyov, E.V. Tipisova, L.V. Osadchuk, Effects of (Elmsford, NY) 23 (2007) 366–373.
overweight and obesity on the spermogram values and levels of [250] B. Lucas, C. Fields, M.C. Hofmann, Signaling pathways in spermatogonial
reproductive hormones in the male population of the European north of stem cells and their disruption by toxicants birth defects research part C,
Russia, Bull. Exp. Biol. Med. 157 (2014) 95–98. Embryo Today: Rev. 87 (2009) 35–42.
[221] J.E. Cavaco, S.S. Laurentino, A. Barros, M. Sousa, S. Socorro, Estrogen [251] J.D. Meeker, S. Ehrlich, T.L. Toth, D.L. Wright, A.M. Calafat, A.T. Trisini, et al.,
receptors alpha and beta in human testis: both isoforms are expressed, Syst. Semen quality and sperm DNA damage in relation to urinary bisphenol A
Biol. Reprod. Med. 55 (2009) 137–144. among men from an infertility clinic, Reprod. Toxicol. (Elmsford, NY) 30
[222] E. Filipiak, D. Suliborska, M. Laszczynska, R. Walczak-Jedrzejowska, E. (2010) 532–539.
Oszukowska, K. Marchlewska, et al., Estrogen receptor alpha localization in [252] T.H. Lassen, H. Frederiksen, T.K. Jensen, J.H. Petersen, U.N. Joensen, K.M.
the testes of men with normal spermatogenesis, Folia Histochem. Cytobiol. Main, et al., Urinary bisphenol A levels in young men: association with
50 (2012) 340–345. reproductive hormones and semen quality, Environ. Health Perspect. 122
[223] V. Rago, F. Romeo, F. Giordano, M. Maggiolini, A. Carpino, Identification of (2014) 478–484.
the estrogen receptor GPER in neoplastic and non-neoplastic human testes, [253] N.P. Chung, C.Y. Cheng, Is cadmium chloride-induced inter-sertoli tight
Reprod. Biol. Endocrinol. 9 (2011) 135. junction permeability barrier disruption a suitable in vitro model to study
[224] F. Sandner, H. Welter, J.U. Schwarzer, F.M. Kohn, H.F. Urbanski, A. the events of junction disassembly during spermatogenesis in the rat testis,
Mayerhofer, Expression of the oestrogen receptor GPER by testicular Endocrinology 142 (2001) 1878–1888.
peritubular cells is linked to sexual maturation and male fertility, Andrology [254] E.R. Siu, E.W. Wong, D.D. Mruk, K.L. Sze, C.S. Porto, C.Y. Cheng, An
2 (2014) 695–701. occludin-focal adhesion kinase protein complex at the blood-testis barrier:
[225] T.L. Nielsen, C. Hagen, K. Wraae, K. Brixen, P.H. Petersen, E. Haug, et al., a study using the cadmium model, Endocrinology 150 (2009) 3336–3344.
Visceral and subcutaneous adipose tissue assessed by magnetic resonance [255] F. El-Sabeawy, S. Wang, J. Overstreet, M. Miller, B. Lasley, E. Enan, Treatment
imaging in relation to circulating androgens, sex hormone-binding globulin, of rats during pubertal development with
and luteinizing hormone in young men, J. Clin. Endocrinol. Metabol. 92 2,3,7,8-tetrachlorodibenzo-p-dioxin alters both signaling kinase activities
(2007) 2696–2705. and epidermal growth factor receptor binding in the testis and the motility
[226] E.C. Tsai, A.M. Matsumoto, W.Y. Fujimoto, E.J. Boyko, Association of and acrosomal reaction of sperm, Toxicol. Appl. Pharmacol. 150 (1998)
bioavailable, free, and total testosterone with insulin resistance: influence of 427–442.
sex hormone-binding globulin and body fat, Diabetes Care 27 (2004) [256] E.W. Wong, C.Y. Cheng, Impacts of environmental toxicants on male
861–868. reproductive dysfunction, Trends Pharmacol. Sci. 32 (2011) 290–299.
[227] D. Landry, F. Cloutier, L.J. Martin, Implications of leptin in neuroendocrine [257] K. Van Waeleghem, N. De Clercq, L. Vermeulen, F. Schoonjans, F. Comhaire,
regulation of male reproduction, Reprod. Biol. 13 (2013) 1–14. Deterioration of sperm quality in young healthy Belgian men, Hum. Reprod.
[228] R. Luboshitzky, L. Lavie, Z. Shen-Orr, P. Herer, Altered luteinizing hormone (1996) 325–329.
and testosterone secretion in middle-aged obese men with obstructive [258] ATSDR. Toxicological Profile for Lead 2007. http://www.atsdr.cdc.gov/
sleep apnea, Obes. Res. 13 (2005) 780–786. toxprofiles/tp.asp?id=96&tid=22.
[229] L.J. Martin, Implications of adiponectin in linking metabolism to testicular [259] B. Quintanilla-Vega, D.J. Hoover, W. Bal, E.K. Silbergeld, M.P. Waalkes, L.D.
function, Endocrine 46 (2014) 16–28. Anderson, Lead interaction with human protamine (HP2) as a mechanism of
[230] P.M. Mah, G.A. Wittert, Obesity and testicular function, Mol. Cell. Endocrinol. male reproductive toxicity, Chem. Res. Toxicol. 13 (2000) 594–600.
316 (2010) 180–186. [260] J.M. Ratcliffe, S.M. Schrader, K. Steenland, D.E. Clapp, T. Turner, R.W.
[231] L. Rato, M.G. Alves, J.E. Cavaco, P.F. Oliveira, High-energy diets: a threat for Hornung, Semen quality in papaya workers with long term exposure to
male fertility? Obesity Rev. 15 (2014) 996–1007. ethylene dibromide, Br. J. Ind. Med. 44 (1987) 317–326.
[232] Y. Okada, K. Tateishi, Y. Zhang, Histone demethylase JHDM2A is involved in [261] S.M. Schrader, T.W. Turner, J.M. Ratcliffe, The effects of ethylene dibromide
male infertility and obesity, J. Androl. 31 (2010) 75–78. on semen quality: a comparison of short-term and chronic exposure,
[233] A.S. di Frega, B. Dale, L. Di Matteo, M. Wilding, Secondary male factor Reprod. Toxicol. (Elmsford, NY) 2 (1988) 191–198.
infertility after Roux-en-Y gastric bypass for morbid obesity: case report, [262] D. Amir, The spermicidal effect of ethylene dibromide in bulls and rams,
Hum. Reprod. 20 (2005) 997–998. Mol. Reprod. Dev. 28 (1991) 99–109.
[234] N. Sermondade, N. Massin, F. Boitrelle, J. Pfeffer, F. Eustache, C. Sifer, et al., [263] U.S. Dept. of Health and Human Services PHS, National Toxicology Program.
Sperm parameters and male fertility after bariatric surgery: three case Report on carcinogens: 1,2-dibromo-3-chloropropane. Report on
series, Reprod. Biomed. Online 24 (2012) 206–210. carcinogens: carcinogen profiles/US Dept of Health and Human Services,
[235] L. Lazaros, E. Hatzi, S. Markoula, A. Takenaka, N. Sofikitis, K. Zikopoulos, Public Health Service, National Toxicology Program 2011;12:134–5.
et al., Dramatic reduction in sperm parameters following bariatric surgery: [264] G. Potashnik, A. Porath, Dibromochloropropane (DBCP): a 17-year
report of two cases, Andrologia 44 (2012) 428–432. reassessment of testicular function and reproductive performance, J. Occup.
[236] J.C. Wiebe, A. Santana, N. Medina-Rodriguez, M. Hernandez, J. Novoa, D. Environ. Med./Am. College Occup. Environ. Med. 37 (1995) 1287–1292.
Mauricio, et al., Fertility is reduced in women and in men with type 1 [265] C.A. Easley, J.M. Bradner, A. Moser, C.A. Rickman, Z.T. McEachin, M.M.
diabetes: results from the Type 1 Diabetes Genetics Consortium (T1DGC), Merritt, et al., Assessing reproductive toxicity of two environmental
Diabetologia 57 (2014) 2501–2504. toxicants with a novel in vitro human spermatogenic model, Stem Cell Res.
[237] S. La Vignera, A.E. Calogero, R. Condorelli, F. Lanzafame, B. Giammusso, E. 14 (2015) 347–355.
Vicari, Andrological characterization of the patient with diabetes mellitus, [266] J.I. Gorelick, M. Goldstein, Loss of fertility in men with varicocele, Fertil.
Minerva Endocrinol. 34 (2009) 1–9. Steril. 59 (1993) 613–616.
[238] S. La Vignera, R. Condorelli, E. Vicari, R. D’Agata, A.E. Calogero, Diabetes [267] World Health Organization, The influence of varicocele on parameters of
mellitus and sperm parameters, J. Androl. 33 (2012) 145–153. fertility in a large group of men presenting to infertility clinics. World
[239] S.M. Bhattacharya, M. Ghosh, N. Nandi, Diabetes mellitus and abnormalities Health Organization. Fertility and sterility. 57 (1992) 1289–1293.
in semen analysis, J. Obstet. Gynaecol. Res. 40 (2014) 167–171. [268] F.F. Pasqualotto, A.M. Lucon, P.M. de Goes, B.P. Sobreiro, J. Hallak, E.B.
[240] V. Bartak, M. Josifko, M. Horackova, Juvenile diabetes and human sperm Pasqualotto, et al., Semen profile, testicular volume, and hormonal levels in
quality, Int. J. Fertil. 20 (1975) 30–32. infertile patients with varicoceles compared with fertile men with and
[241] I.M. Agbaje, D.A. Rogers, C.M. McVicar, N. McClure, A.B. Atkinson, C. Mallidis, without varicoceles, Fertil. Steril. 83 (2005) 74–77.
et al., Insulin dependant diabetes mellitus: implications for male [269] D.E. Johnson, D.R. Pohl, H. Rivera-Correa, Varicocele: an innocuous
reproductive function, Hum. Reprod. 22 (2007) 1871–1877. condition, South. Med. J. 63 (1970) 34–36.
F.T.L. Neto et al. / Seminars in Cell & Developmental Biology 59 (2016) 10–26 25

[270] J. Zargooshi, Sperm count and sperm motility in incidental high-grade testosterone therapy in Klinefelter patients, J. Clin. Endocrinol. Metabol. 89
varicocele, Fertil. Steril. 88 (2007) 1470–1473. (2004) 6208–6217.
[271] L. Dubin, R.D. Amelar, Varicocelectomy: 986 cases in a twelve-year study, [302] F. Tuttelmann, O.S. Damm, C.M. Luetjens, M. Baldi, M. Zitzmann, S. Kliesch,
Urology 10 (1977) 446–449. et al., Intratesticular testosterone is increased in men with Klinefelter
[272] M. Goldstein, B.R. Gilbert, A.P. Dicker, J. Dwosh, C. Gnecco, Microsurgical syndrome and may not be released into the bloodstream owing to altered
inguinal varicocelectomy with delivery of the testis: an artery and testicular vascularization—a preliminary report, Andrology 2 (2014)
lymphatic sparing technique, J. Urol. 148 (1992) 1808–1811. 275–281.
[273] J.L. Marmar, A. Agarwal, S. Prabakaran, R. Agarwal, R.A. Short, S. Benoff, et al., [303] L. Aksglaede, A.M. Wikstrom, E. Rajpert-De Meyts, L. Dunkel, N.E.
Reassessing the value of varicocelectomy as a treatment for male Skakkebaek, A. Juul, Natural history of seminiferous tubule degeneration in
subfertility with a new meta-analysis, Fertil. Steril. 88 (2007) 639–648. Klinefelter syndrome, Hum. Reprod. Update 12 (2006) 39–48.
[274] L.M. Su, M. Goldstein, P.N. Schlegel, The effect of varicocelectomy on serum [304] M.S. Wosnitzer, D.A. Paduch, Endocrinological issues and hormonal
testosterone levels in infertile men with varicoceles, J. Urol. 154 (1995) manipulation in children and men with Klinefelter syndrome, Am. J. Med.
1752–1755. Genet. Part C Semin. Med. Genet. 163C (2013) 16–26.
[275] M. Goldstein, J.F. Eid, Elevation of intratesticular and scrotal skin surface [305] M. Kotula-Balak, L. Bablok, S. Fracki, A. Jankowska, B. Bilinska,
temperature in men with varicocele, J. Urol. 142 (1989) 743–745. Immunoexpression of androgen receptors and aromatase in testes of patient
[276] C. Wang, Y.G. Cui, X.H. Wang, Y. Jia, A. Sinha Hikim, Y.H. Lue, et al., Transient with Klinefelter’s syndrome, Folia Histochem. Cytobiol. 42 (2004) 215–220.
scrotal hyperthermia and levonorgestrel enhance testosterone-induced [306] Y.H. Yu, F.P. Siao, L.C. Hsu, P.H. Yen, TEX11 modulates germ cell proliferation
spermatogenesis suppression in men through increased germ cell apoptosis, by competing with estrogen receptor beta for the binding to HPIP, Mol.
J. Clin. Endocrinol. Metabol. 92 (2007) 3292–3304. Endocrinol. 26 (2012) 630–642.
[277] H. Nishiyama, S. Danno, Y. Kaneko, K. Itoh, H. Yokoi, M. Fukumoto, et al., [307] R. Ramasamy, J.A. Ricci, G.D. Palermo, L.V. Gosden, Z. Rosenwaks, P.N.
Decreased expression of cold-inducible RNA-binding protein (CIRP) in male Schlegel, Successful fertility treatment for Klinefelter’s syndrome, J. Urol.
germ cells at elevated temperature, Am. J. Pathol. 152 (1998) 289–296. 182 (2009) 1108–1113.
[278] S. Ogi, N. Tanji, T. Iseda, M. Yokoyama, Expression of heat shock proteins in [308] P. Lenz, C.M. Luetjens, A. Kamischke, B. Kuhnert, I. Kennerknecht, E.
developing and degenerating rat testes, Arch. Androl. 43 (1999) 163–171. Nieschlag, Mosaic status in lymphocytes of infertile men with or without
[279] K. Shiraishi, H. Takihara, H. Matsuyama, Elevated scrotal temperature, but Klinefelter syndrome, Hum. Reprod. 20 (2005) 1248–1255.
not varicocele grade, reflects testicular oxidative stress-mediated apoptosis, [309] C. Foresta, C. Galeazzi, A. Bettella, P. Marin, M. Rossato, A. Garolla, et al.,
World J. Urol. 28 (2010) 359–364. Analysis of meiosis in intratesticular germ cells from subjects affected by
[280] D.S. Zylbersztejn, C. Andreoni, P.T. Del Giudice, D.M. Spaine, L. Borsari, G.H. classic Klinefelter’s syndrome, J. Clin. Endocrinol. Metabol. 84 (1999)
Souza, et al., Proteomic analysis of seminal plasma in adolescents with and 3807–3810.
without varicocele, Fertil. Steril. 99 (2013) 92–98. [310] F. Lanfranco, A. Kamischke, M. Zitzmann, E. Nieschlag, Klinefelter’s
[281] P.T. Del Giudice, B.F. da Silva, E.G. Lo Turco, R. Fraietta, D.M. Spaine, L.F. syndrome, Lancet (Lond., Engl.) 364 (2004) 273–283.
Santos, et al., Changes in the seminal plasma proteome of adolescents before [311] P.H. Vogt, A. Edelmann, S. Kirsch, O. Henegariu, P. Hirschmann, F.
and after varicocelectomy, Fertil. Steril. 100 (2013) 667–672. Kiesewetter, et al., Human Y chromosome azoospermia factors (AZF)
[282] M. Camargo, P. Intasqui, R.P. Bertolla, Proteomic profile of seminal plasma in mapped to different subregions in Yq11, Hum. Mol. Genet. 5 (1996)
adolescents and adults with treated and untreated varicocele, Asian J. 933–943.
Androl. 18 (2016) 194–201. [312] P.J. Stahl, P. Masson, A. Mielnik, M.B. Marean, P.N. Schlegel, D.A. Paduch, A
[283] A.P. Drabovich, P. Saraon, K. Jarvi, E.P. Diamandis, Seminal plasma as a decade of experience emphasizes that testing for Y microdeletions is
diagnostic fluid for male reproductive system disorders, Nat. Rev. Urol. 11 essential in American men with azoospermia and severe oligozoospermia,
(2014) 278–288. Fertil. Steril. 94 (2010) 1753–1756.
[284] I.T. Koksal, Y. Ishak, M. Usta, A. Danisman, E. Guntekin, I.C. Bassorgun, et al., [313] P. Navarro-Costa, Sex, rebellion and decadence: the scandalous evolutionary
Varicocele-induced testicular dysfunction may be associated with history of the human Y chromosome, Biochim. Biophys. Acta 1822 (2012)
disruption of blood-testis barrier, Arch. Androl. 53 (2007) 43–48. 1851–1863.
[285] C.K. Naughton, A.K. Nangia, A. Agarwal, Pathophysiology of varicoceles in [314] S. Sen, A.R. Pasi, R. Dada, M.B. Shamsi, D.Y. Modi, Chromosome
male infertility, Hum. Reprod. Update 7 (2001) 473–481. microdeletions in infertile men: prevalence, phenotypes and screening
[286] M. Cocuzza, K.S. Athayde, A. Agarwal, R. Pagani, S.C. Sikka, A.M. Lucon, et al., markers for the Indian population, J. Assist. Reprod. Genet. 30 (2013)
Impact of clinical varicocele and testis size on seminal reactive oxygen 413–422.
species levels in a fertile population: a prospective controlled study, Fertil. [315] C.V. Hopps, A. Mielnik, M. Goldstein, G.D. Palermo, Z. Rosenwaks, P.N.
Steril. 90 (2008) 1103–1108. Schlegel, Detection of sperm in men with Y chromosome microdeletions of
[287] A. Zini, G. Defreitas, M. Freeman, S. Hechter, K. Jarvi, Varicocele is associated the AZFa, AZFb and AZFc regions, Hum. Reprod. 18 (2003) 1660–1665.
with abnormal retention of cytoplasmic droplets by human spermatozoa, [316] P. Asero, A.E. Calogero, R.A. Condorelli, L. Mongioi, E. Vicari, F. Lanzafame,
Fertil. Steril. 74 (2000) 461–464. et al., Relevance of genetic investigation in male infertility, J. Endocrinol.
[288] A.A. Ibrahim, H.A. Awad, S. El-Haggar, B.A. Mitawi, Bilateral testicular biopsy Invest. 37 (2014) 415–427.
in men with varicocele, Fertil. Steril. 28 (1977) 663–667. [317] R. Saxena, J.W. de Vries, S. Repping, R.K. Alagappan, H. Skaletsky, L.G. Brown,
[289] J.J. Sirvent, R. Bernat, M.A. Navarro, J. Rodriguez Tolra, R. Guspi, R. Bosch, et al., Four DAZ genes in two clusters found in the AZFc region of the human
Leydig cell in idiopathic varicocele, Eur. Urol. 17 (1990) 257–261. Y chromosome, Genomics 67 (2000) 256–267.
[290] P. Agger, S.G. Johnsen, Quantitative evaluation of testicular biopsies in [318] T. Kuroda-Kawaguchi, H. Skaletsky, L.G. Brown, P.J. Minx, H.S. Cordum, R.H.
varicocele, Fertil. Steril. 29 (1978) 52–57. Waterston, et al., The AZFc region of the Y chromosome features massive
[291] G.S. Hurt, S.S. Howards, T.T. Turner, The effects of unilateral, experimental palindromes and uniform recurrent deletions in infertile men, Nat. Genet.
varicocele are not mediated through the ipsilateral testis, J. Androl. 8 (1987) 29 (2001) 279–286.
403–408. [319] M.J. Noordam, S.K. van Daalen, S.E. Hovingh, C.M. Korver, F. van der Veen,
[292] T.J. Walsh, R.R. Pera, P.J. Turek, The genetics of male infertility, Semin. S.A. Repping, novel partial deletion of the Y chromosome azoospermia
Reprod. Med. 27 (2009) 124–136. factor c region is caused by non-homologous recombination between
[293] K.L. O’Flynn O’Brien, A.C. Varghese, A. Agarwal, The genetic causes of male palindromes and may be associated with increased sperm counts, Hum.
factor infertility: a review, Fertil. Steril. 93 (2010) 1–12. Reprod. 26 (2011) 713–723.
[294] P.N. Schlegel, G.D. Palermo, M. Goldstein, S. Menendez, N. Zaninovic, L.L. [320] P.C. Patsalis, C. Sismani, L. Quintana-Murci, F. Taleb-Bekkouche, C. Krausz, K.
Veeck, et al., Testicular sperm extraction with intracytoplasmic sperm McElreavey, Effects of transmission of Y chromosome AZFc deletions, Lancet
injection for nonobstructive azoospermia, Urology 49 (1997) 435–440. (Lond., Engl.) 360 (2002) 1222–1224.
[295] L. Abramsky, J. Chapple, 47,XXY (Klinefelter syndrome) and 47,XYY: [321] S. Repping, H. Skaletsky, J. Lange, S. Silber, F. Van Der Veen, R.D. Oates, et al.,
estimated rates of and indication for postnatal diagnosis with implications Recombination between palindromes P5 and P1 on the human Y
for prenatal counselling, Prenat. Diagn. 17 (1997) 363–368. chromosome causes massive deletions and spermatogenic failure, Am. J.
[296] P.J. Stahl, P.N. Schlegel, Genetic evaluation of the azoospermic or severely Hum. Genet. 71 (2002) 906–922.
oligozoospermic male, Curr. Opin. Obstet. Gynecol. 24 (2012) 221–228. [322] A. Ferlin, B. Arredi, E. Speltra, C. Cazzadore, R. Selice, A. Garolla, et al.,
[297] C. Foresta, C. Galeazzi, A. Bettella, M. Stella, C. Scandellari, High incidence of Molecular and clinical characterization of Y chromosome microdeletions in
sperm sex chromosomes aneuploidies in two patients with Klinefelter’s infertile men: a 10-year experience in Italy, J. Clin. Endocrinol. Metabol. 92
syndrome, J. Clin. Endocrinol. Metabol. 83 (1998) 203–205. (2007) 762–770.
[298] Y. Iitsuka, A. Bock, D.D. Nguyen, C.A. Samango-Sprouse, J.L. Simpson, F.Z. [323] C. Ramathal, B. Angulo, M. Sukhwani, J. Cui, J. Durruthy-Durruthy, F. Fang,
Bischoff, Evidence of skewed X-chromosome inactivation in 47,XXY and et al., DDX3Y gene rescue of a Y chromosome AZFa deletion restores germ
48,XXYY Klinefelter patients, Am. J. Med. Genet. 98 (2001) 25–31. cell formation and transcriptional programs, Sci. Rep. 5 (2015) 15041.
[299] L. Carrel, H.F. Willard, X-inactivation profile reveals extensive variability in [324] C. Foresta, E. Moro, A. Rossi, M. Rossato, A. Garolla, A. Ferlin, Role of the AZFa
X-linked gene expression in females, Nature 434 (2005) 400–404. candidate genes in male infertility, J. Endocrinol. Invest. 23 (2000) 646–651.
[300] M.K. Samplaski, K.C. Lo, E.D. Grober, A. Millar, A. Dimitromanolakis, K.A. [325] P. Blanco, M. Shlumukova, C.A. Sargent, M.A. Jobling, N. Affara, M.E. Hurles,
Jarvi, Phenotypic differences in mosaic Klinefelter patients as compared Divergent outcomes of intrachromosomal recombination on the human Y
with non-mosaic Klinefelter patients, Fertil. Steril. 101 (2014) 950–955. chromosome: male infertility and recurrent polymorphism, J. Med. Genet.
[301] M. Zitzmann, M. Depenbusch, J. Gromoll, E. Nieschlag, X-chromosome 37 (2000) 752–758.
inactivation patterns and androgen receptor functionality influence [326] V. Ambros, The functions of animal microRNAs, Nature 431 (2004) 350–355.
phenotype and social characteristics as well as pharmacogenetics of
26 F.T.L. Neto et al. / Seminars in Cell & Developmental Biology 59 (2016) 10–26

[327] R.C. Lee, R.L. Feinbaum, V. Ambros, The C. elegans heterochronic gene lin-4 [343] M. Vigodner, T. Ishikawa, P.N. Schlegel, P.L. Morris, SUMO-1, human male
encodes small RNAs with antisense complementarity to lin-14, Cell 75 germ cell development, and the androgen receptor in the testis of men with
(1993) 843–854. normal and abnormal spermatogenesis, Am. J. Physiol. Endocrinol. Metab.
[328] D.P. Bartel, MicroRNAs: genomics, biogenesis, mechanism, and function, Cell 290 (2006) E1022–E1033.
116 (2004) 281–297. [344] K. Havas, A. Flaus, M. Phelan, R. Kingston, P.A. Wade, D.M. Lilley, et al.,
[329] W. Wu, Z. Hu, Y. Qin, J. Dong, J. Dai, C. Lu, et al., Seminal plasma microRNAs: Generation of superhelical torsion by ATP-dependent chromatin remodeling
potential biomarkers for spermatogenesis status, Mol. Hum. Reprod. 18 activities, Cell 103 (2000) 1133–1142.
(2012) 489–497. [345] R.E. Kingston, G.J. Narlikar, ATP-dependent remodeling and acetylation as
[330] J. Lian, X. Zhang, H. Tian, N. Liang, Y. Wang, C. Liang, et al., Altered microRNA regulators of chromatin fluidity, Genes Dev. 13 (1999) 2339–2352.
expression in patients with non-obstructive azoospermia, Reprod. Biol. [346] V. Sonnack, K. Failing, M. Bergmann, K. Steger, Expression of
Endocrinol. 7 (2009) 13. hyperacetylated histone H4 during normal and impaired human
[331] M.D. Papaioannou, S. Nef, microRNAs in the testis: building up male fertility, spermatogenesis, Andrologia 34 (2002) 384–390.
J. Androl. 31 (2010) 26–33. [347] K. Steger, K. Failing, T. Klonisch, H.M. Behre, M. Manning, W. Weidner, et al.,
[332] S. de Mateo, P. Sassone-Corsi, Regulation of spermatogenesis by small Round spermatids from infertile men exhibit decreased protamine-1 and -2
non-coding RNAs: role of the germ granule, Semin. Cell Dev. Biol. 29 (2014) mRNA, Hum. Reprod. 16 (2001) 709–716.
84–92. [348] J.M. Gatewood, G.R. Cook, R. Balhorn, E.M. Bradbury, C.W. Schmid,
[333] A. Salas-Huetos, J. Blanco, F. Vidal, J.M. Mercader, N. Garrido, E. Anton, New Sequence-specific packaging of DNA in human sperm chromatin, Science
insights into the expression profile and function of micro-ribonucleic acid in 236 (1987) 962–964.
human spermatozoa, Fertil. Steril. 102 (2014) 213–222, e4. [349] M. Gardiner-Garden, M. Ballesteros, M. Gordon, Tam PP. Histone- and
[334] A. Kozomara, S. Griffiths-Jones, miRBase: annotating high confidence protamine-DNA association: conservation of different patterns within the
microRNAs using deep sequencing data, Nucleic Acids Res. 42 (2014) beta-globin domain in human sperm, Mol. Cell. Biol. 18 (1998) 3350–3356.
D68–73. [350] D.T. Carrell, Epigenetics of the male gamete, Fertil. Steril. 97 (2012) 267–274.
[335] M. Abu-Halima, M. Hammadeh, C. Backes, U. Fischer, P. Leidinger, A.M. [351] E. Li, Chromatin modification and epigenetic reprogramming in mammalian
Lubbad, et al., Panel of five microRNAs as potential biomarkers for the development, Nat. Rev. Genet. 3 (2002) 662–673.
diagnosis and assessment of male infertility, Fertil. Steril. 102 (2014) [352] S. Houshdaran, V.K. Cortessis, K. Siegmund, A. Yang, P.W. Laird, R.Z. Sokol,
989–997, e1. Widespread epigenetic abnormalities suggest a broad DNA methylation
[336] C. Wang, C. Yang, X. Chen, B. Yao, C. Yang, C. Zhu, et al., Altered profile of erasure defect in abnormal human sperm, PLoS One 2 (2007) e1289.
seminal plasma microRNAs in the molecular diagnosis of male infertility, [353] M. Benchaib, V. Braun, D. Ressnikof, J. Lornage, P. Durand, A. Niveleau, et al.,
Clin. Chem. 57 (2011) 1722–1731. Influence of global sperm DNA methylation on IVF results, Hum. Reprod. 20
[337] W. Wu, Y. Qin, Z. Li, J. Dong, J. Dai, C. Lu, et al., Genome-wide microRNA (2005) 768–773.
expression profiling in idiopathic non-obstructive azoospermia: significant [354] D. Montjean, A. Zini, C. Ravel, S. Belloc, A. Dalleac, H. Copin, et al., Sperm
up-regulation of miR-141, miR-429 and miR-7-1-3p, Hum. Reprod. 28 global DNA methylation level: association with semen parameters and
(2013) 1827–1836. genome integrity, Andrology 3 (2015) 235–240.
[338] A.A. Dabaja, A. Mielnik, B.D. Robinson, M.S. Wosnitzer, P.N. Schlegel, D.A. [355] L.N. Odom, J. Segars, Imprinting disorders and assisted reproductive
Paduch, Possible germ cell-Sertoli cell interactions are critical for technology, Curr. Opin. Endocrinol. Diabetes Obes. 17 (2010) 517–522.
establishing appropriate expression levels for the Sertoli cell-specific [356] M.D. Anway, A.S. Cupp, M. Uzumcu, M.K. Skinner, Epigenetic
MicroRNA, miR-202-5p, in human testis, Basic Clin. Androl. 25 (2015) 2. transgenerational actions of endocrine disruptors and male fertility, Science
[339] R. Dada, M. Kumar, R. Jesudasan, J.L. Fernandez, J. Gosalvez, A. Agarwal, 308 (2005) 1466–1469.
Epigenetics and its role in male infertility, J. Assist. Reprod. Genet. 29 (2012) [357] I.V.S. Donkin, L.R. Ingerslev, K. Qian, M. Mechta, L. Nordkap, B. Mortensen,
213–223. E.V. Appel, N. Jørgensen, V.B. Kristiansen, T. Hansen, C.T. Workman, J.R.
[340] D. Takai, P.A. Jones, Comprehensive analysis of CpG islands in human Zierath, R. Romain Barre, Obesity and bariatric surgery drive epigenetic
chromosomes 21 and 22, Proc. Natl. Acad. Sci. U.S.A. 99 (2002) 3740–3745. variation of spermatozoa in humans, Cell Metab. 23 (2016) 1–10.
[341] W. Fischle, Y. Wang, S.A. Jacobs, Y. Kim, C.D. Allis, S. Khorasanizadeh, [358] L. Stuppia, M. Franzago, P. Ballerini, V. Gatta, I. Antonucci, Epigenetics and
Molecular basis for the discrimination of repressive methyl-lysine marks in male reproduction: the consequences of paternal lifestyle on fertility,
histone H3 by polycomb and HP1 chromodomains, Genes Dev. 17 (2003) embryo development, and children lifetime health, Clin. Epigenet. 7 (2015)
1870–1881. 120.
[342] B.D. Strahl, C.D. Allis, The language of covalent histone modifications, Nature
403 (2000) 41–45.

You might also like