Submitted Successfully!
To reward your contribution, here is a gift for you: A free trial for our video production service.
Thank you for your contribution! You can also upload a video entry or images related to this topic.
Version Summary Created by Modification Content Size Created at Operation
1 -- 1711 2022-06-30 03:36:44 |
2 format correct Meta information modification 1711 2022-06-30 04:25:08 |

Video Upload Options

Do you have a full video?

Confirm

Are you sure to Delete?
Cite
If you have any further questions, please contact Encyclopedia Editorial Office.
Barbagallo, F.;  Vignera, S.L.;  Cannarella, R.;  Aversa, A.;  Calogero, A.E.;  Condorelli, R.A. Sperm Mitochondrial Function. Encyclopedia. Available online: https://encyclopedia.pub/entry/24654 (accessed on 27 July 2024).
Barbagallo F,  Vignera SL,  Cannarella R,  Aversa A,  Calogero AE,  Condorelli RA. Sperm Mitochondrial Function. Encyclopedia. Available at: https://encyclopedia.pub/entry/24654. Accessed July 27, 2024.
Barbagallo, Federica, Sandro La Vignera, Rossella Cannarella, Antonio Aversa, Aldo E. Calogero, Rosita A. Condorelli. "Sperm Mitochondrial Function" Encyclopedia, https://encyclopedia.pub/entry/24654 (accessed July 27, 2024).
Barbagallo, F.,  Vignera, S.L.,  Cannarella, R.,  Aversa, A.,  Calogero, A.E., & Condorelli, R.A. (2022, June 30). Sperm Mitochondrial Function. In Encyclopedia. https://encyclopedia.pub/entry/24654
Barbagallo, Federica, et al. "Sperm Mitochondrial Function." Encyclopedia. Web. 30 June, 2022.
Sperm Mitochondrial Function
Edit

Sperm motility is a fundamental requirement to ensure male fertility. Studies and interest for sperm motility started in 1919 when Lillie Frank Rattray, an American zoologist, author of the book “Problems of fertilization”, for the first time, talked on the energetic metabolism of spermatozoa. He said: “Spermatozoa are probably incapable of receiving nourishment outside of the gonad after they are fully differentiated; certainly in the case of external insemination there is no opportunity for the restitution of substance …”. Since those years, several studies focused on the “power plant” of the cell, the mitochondrion, demonstrating the key role of this organelle on cellular homeostasis and sperm motility. Mitochondrial sperm dysfunction is also implicated in the pathogenesis of seminal oxidative stress, a key element responsible of many cases of “apparently” idiopathic male infertility.

sperm motility asthenozoospermia mitochondrial function antioxidants

1. Sperm Mitochondria: Anatomy

Mammalian spermatozoa typically have between 50 and 75 mitochondria [1]. Mitochondria of spermatozoa show peculiar characteristics. They are exclusively confined in the mid-piece, tightly wrapped around the axoneme. During spermiogenesis, the mitochondria line up end-to-end and wrap helically around the flagellum to form the thick mitochondrial sheath, just under the outer plasma membrane of the cell [2][3]. The mitochondrial capsule is thus shaped by multiple disulphide bridges formed between cysteine and proline-rich selenoproteins [4][5][6]. This intricate attachment to the fibrous sheath makes sperm mitochondria particularly difficult to isolate by conventional separation method [7]. Moreover, many proteins and enzymes such as subunit VIb of the cytochrome oxidase [8], E1-pyruvate decarboxylase and creatine kinase (CK) [9][10] are isoforms present only in sperm mitochondria.

2. Sperm Mitochondrial Metabolism

2.1. A Long Debate: Glycolysis or Oxidative Phosphorylation?

Many studies on sperm mitochondrial bioenergetics have concentrated on the following question: what is the main biochemical pathway that provides the energy for sperm motility, glycolysis or oxidative phosphorylation (OXPHOS)? Studies carried out in several species, including humans, have often provided different and/or conflicting results. Many authors have reached the conclusion that glycolysis has a primary role in energy production in human sperm motility [11][12][13][14]. Other authors have underlined the importance of mitochondrial OXPHOS for sperm motility [15][16][17][18][19][20][21][22]. Moreover, the experimental conditions varied significantly from one study to another and this complicated the interpretation of the results. Therefore, despite numerous studies, a clear conclusion cannot be drawn [6]. A reasonable concept that emerged from many studies is that these processes are not mutually exclusive and that spermatozoa exhibit a great versatility in their metabolism using glycolysis exclusively, mitochondrial OXPHOS exclusively or a combination of both pathways for energy production according to the substrates available in the female genital tracts [23][24]. Specifically, Zhu et al. had recently shown that mitochondrial oxidative phosphorylation is activated to produce ATP under low glucose condition. They incubated boar spermatozoa with different doses of glucose and they found that sperm progressive motility and straight-line velocity were significantly increased with decreasing glucose level in the incubation medium. They also showed that, in presence of the mitochondrial translation inhibitor d-chloramphenicol, mitochondrial protein synthesis, mitochondrial activity and ATP level were suppressed, and consequently, the linear motility speed decreased. Interestingly, despite the reduction of linear motility speed, total motility did not change [25]. These results suggest that sperm motility patterns depend on the substrates available and as a result on the biochemical pathway that is activated. While glycolysis is important for hyperactivated motility [26], the high-speed linear motility is induced via activating the mitochondrial activity in low glucose condition.

2.2. Reactive Oxygen Species and Sperm Mitochondria

Since mitochondria have a main role in sperm metabolism and energy production, they are the major reactive oxygen species (ROS) generator, as they convert approximately 1–2% of consumed oxygen into superoxide anions [27]. In spermatozoa, mitochondrial Complex I and Complex III are the major sites for ROS production [27]. An imbalance between impaired ROS production and antioxidants mechanisms may be extremely harmful for spermatozoa [28]. These cells are particularly susceptible to oxidative stress because they cannot restore damage caused by oxidative stress due to deficiency of cytoplasmic repair enzymes [29]. In contrast, at physiological concentrations, ROS are trigger for several reproductive mechanisms such as sperm capacitation, hyperactivation, acrosome reaction and oocyte fusion [30][31][32]. More recent research is changing the long-accepted dogma that ROS is ultimately a negative indicator of sperm function, thus indicating that ROS production can also reflect intense mitochondrial activity leading to increased sperm function [33].

3. Techniques to Study Sperm Mitochondria Function

Since sperm mitochondria have been related to sperm motility and fertilization, numerous tools have been developed to evaluate their function. Several parameters can be used to study sperm mitochondria. This include mitochondrial activity, MMP levels and mitochondrial calcium levels [34]. The use of fluorescent probes to detect changes in the MMP is the most popular tool used to evaluate mitochondrial function. These probes spread freely through the plasma membrane to the cell cytosol and accumulate electrophoretically in the mitochondrial matrix, due to the motive force of the proton and acting according to the ability of the mitochondria to pump the protons from the matrix to the intermembrane area [35][36][37]. Therefore, according to their properties, high concentrations of these fluorochromes accumulate in hyperpolarized mitochondria (high MMP), while lower concentrations are found in depolarized mitochondria (low MMP) and the intensity of the fluorescence correlates with MMP [38].
Many commercial fluorescent dyes are widely used, such as JC-1 (5,5,6,6-tetrachloro-1,1,3,3-tetraethylbenzimidazolyl carbocyanine iodide [36], Mito Tracker Green FM [39] and rhodamine 123 [40]. The JC-1 probe is one of the preferred fluorescent dyes for analyzing MMP and has been widely used for the analysis of spermatozoa in several species including humans [22][41], cattle [42], horse [43], ram [44], dog [45] and alpaca [46]. In 2004, Marchetti et al. compared the specificity of four fluorochromes in the evaluation of sperm MMP. They found that JC-1 is found only within the mitochondria and therefore provides the most accurate measurement of MMP [20]. Lugli et al. confirmed this finding, indicating that JC-1 is more reliable and specific for this type of evaluation than other probes [47]. Amaral and Ramalho-Santos found JC-1 more dynamic and suitable than MitoTracker Green and MitoTracker Red, as it is able to detect minimal changes in MMP [5][48]. Recently, Uribe et al. evaluated the usefulness of another fluorescent dye, tetramethyl rhodamine methyl ester perchlorate (TMRM) for measuring sperm MMP. They found that TMRM is able to accurately detect MMP variations comparable to the method widely used, JC-1 staining. In addition, TMRM was able to measure sperm MMP in the experimental conditions in which JC-1 had previously presented difficulties [49].
Finally, mitochondrial oxygen consumption is considered the central parameter of mitochondrial function. Studies on experimental animals have shown that mitochondrial oxygen consumption is positively correlated with traditional measures of sperm function including motility and vitality [50]. Therefore, measurement of oxygen consumption is another important biological endpoint for the study of sperm mitochondrial function and should be further investigated in future studies.

4. Asthenozoospermia and Sperm Mitochondrial Dysfunction

Almost forty years have passed since the first study relating mitochondrial function to sperm motility. Everson et al. reported a good correlation between sperm motility and MMP, comparing ejaculate from fertile men with those from patients whose spermatozoa showed reduced sperm motility [51]. Ruiz-Pesini et al. compared mitochondrial respiratory complex activities of 86 asthenozoospermic patients with those of 33 controls. Their results showed that semen samples of controls had substantially higher activities of complexes I, II and IV compared with those of patients with asthenozoospermia. Moreover, a direct and positive correlation was found in the whole population studied between spermatozoa motility and all the mitochondrial respiratory complex activities assayed (I, II, I+III, II+III, and IV). They suggested that more specific mitochondrial dysfunctions could be the underlying cause of idiopathic asthenozoospermia [18]. In a following study, these authors found that mitochondrial enzyme activities not only correlate with sperm motility but also with vitality and cell concentration [19]. As mitochondria are the major source of pro-oxidative agents, it is suggested that dysfunction of this organelle would have a fundamental role in the oxidative imbalance affecting sperm function [27]. Wang et al. [52] identified low MMP and high ROS production in spermatozoa from infertile patients, probably as a consequence of such mitochondrial injury. Other researchers have observed changes in mitochondrial function in sperm derived from infertile patients [53]. However, spermatozoa with high MMP have been identified in fertile men [20][21][54][55]. Paoli et al. [41], correlated MMP with increasing motility to establish MMP values corresponding to a precise sequence of gradually increasing motility. They evaluated 185 semen samples, divided into 13 motility classes (from 0 to 60%) with an increment of 5% between classes and a second group of 28 semen samples showing nonlinear motility only, divided into five classes. A positive correlation was found between sperm motility and FL2 (percentage of sperm with high and low MMP) in all samples of both groups. In group A, the first 0 motility class, showed a mean FL2 of 11.5. This value increased gradually with increasing motility, reaching its highest mean value of 75.7% in the 60% motility class. They found both immotility and severe asthenozoospermia to be characterized by an extremely low MMP. These results diverge from those of Piasecka, whose study on 32 subjects with normal motility and 25 patients with asthenozoospermia identified two asthenozoospermic subpopulations, one with a low and the other with a high MMP, suggesting that the reduced motility might be caused not only by functional alteration of mitochondria but also by abnormal morphology of the axoneme, the dense fibers and the fibrous sheath [56]. Pelliccione et al. demonstrated the importance of structural defects in the mitochondrial membrane in asthenozoospermic semen samples, indicating a close correlation between forward motility and the percentage of the intermediate tract having a normal membrane [57].
Over the last years, proteomic studies have tried to identify dysfunctional proteins responsible for asthenozoospermia [58][59]. Recently, Nowicka-Bauer et al. compared proteomic profiles in spermatozoa of normozoospermic men and in patients with isolated asthenozoospermia. The results of this study were further supported by two additional mitochondrial tests (JC-1 and MitoSox Red) to establish a possible direct connection between the identified proteins and functional status of mitochondria [59]. In accordance with a previous study by Amaral et al. [58], they found that most of the identified dysfunctional proteins in low motility spermatozoa were of mitochondrial origin and a great proportion of them were engaged in cell metabolism and energy production being involved in tricarboxylic acid cycle (TCA), mitochondrial OXPHOS and the metabolism of butanoates, propanoates and pyruvates. These results strongly support the emerging idea that several bioenergetics metabolic pathways contribute to sperm motility (de)regulation [58][59].

References

  1. Ankel-Simons, F.; Cummins, J.M. Misconceptions about mitochondria and mammalian fertilization: Implications for theories on human evolution. Proc. Natl. Acad. Sci. USA 1996, 93, 13859–13863.
  2. Lindemann, C.B.; Lesich, K.A. Functional anatomy of the mammalian sperm flagellum. Cytoskeleton 2016, 73, 652–669.
  3. Ho, H.C.; Wey, S. Three dimensional rendering of the mitochondrial sheath morphogenesis during mouse spermiogenesis. Microsc. Res. Tech. 2007, 70, 719–723.
  4. Ursini, F.; Heim, S.; Kiess, M.; Maiorino, M.; Roveri, A.; Wissing, J.; Flohé, L. Dual function of the selenoprotein PHGPx during sperm maturation. Science 1999, 285, 1393–1396.
  5. Ramalho-Santos, J.; Varum, S.; Amaral, S.; Mota, P.C.; Sousa, A.P.; Amaral, A. Mitochondrial functionality in reproduction: From gonads and gametes to embryos and embryonic stem cells. Hum. Reprod. Update 2009, 15, 553–572.
  6. Piomboni, P.; Focarelli, R.; Stendardi, A.; Ferramosca, A.; Zara, V. The role of mitochondria in energy production for human sperm motility. Int. J. Androl. 2012, 35, 109–124.
  7. Ferramosca, A.; Focarelli, R.; Piomboni, P.; Coppola, L.; Zara, V. Oxygen uptake by mitochondria in demembranated human spermatozoa: A reliable tool for the evaluation of sperm respiratory efficiency. Int. J. Androl. 2008, 31, 337–345.
  8. Hüttemann, M.; Jaradat, S.; Grossman, L.I. Cytochrome c oxidase of mammals contains a testes-specific isoform of subunit VIb—The counterpart to testes-specific cytochrome c? Mol. Reprod. Dev. 2003, 66, 8–16.
  9. Gerez de Burgos, N.M.; Gallina, F.; Burgos, C.; Blanco, A. Effect of L-malate on pyruvate dehydrogenase activity of spermatozoa. Arch. Biochem. Biophys. 1994, 308, 520–524.
  10. Huszar, G.; Stone, K.; Dix, D.; Vigue, L. Putative creatine kinase M-isoform in human sperm is identifiedas the 70-kilodalton heat shock protein HspA2. Biol. Reprod. 2000, 63, 925–932.
  11. Peterson, R.N.; Freund, M. Profile of glycolytic enzyme activities in human spermatozoa. Fertil. Steril. 1970, 21, 151–158.
  12. Peterson, R.N.; Freund, M. Citrate formation from exogenous substrates by washed human spermatozoa. J. Reprod. Fertil. 1974, 38, 73–79.
  13. Storey, B.T.; Kayne, F.J. Energy metabolism of spermatozoa. VII. Interactions between lactate, pyruvate and malate as oxidative substrates for rabbit sperm mitochondria. Biol. Reprod. 1978, 18, 527–536.
  14. Williams, A.C.; Ford, W.C. The role of glucose in supporting motility and capacitation in human spermatozoa. J. Androl. 2001, 22, 680–695.
  15. Auger, J.; Ronot, X.; Dadoune, J.P. Human sperm mitochondrial function related to motility: A flow and image cytometric assessment. J. Androl. 1989, 10, 439–448.
  16. Auger, J.; Leonce, S.; Jouannet, P.; Ronot, X. Flow cytometric sorting of living, highly motile human spermatozoa based on evaluation of their mitochondrial activity. J. Histochem. Cytochem. 1993, 41, 1247–1251.
  17. Kramer, R.Y.; Garner, D.L.; Bruns, E.S.; Ericsson, S.A.; Prins, G.S. Comparison of motility and flow cytometric assessments of seminal quality in fresh, 24-h extended and cryopreserved human spermatozoa. J. Androl. 1993, 14, 374–384.
  18. Ruiz-Pesini, E.; Diez, C.; Lapeña, A.C.; Pérez-Martos, A.; Montoya, J.; Alvarez, E.; Arenas, J.; López-Pérez, M.J. Correlation of sperm motility with mitochondrial enzymatic activities. Clin. Chem. 1998, 44, 1616–1620.
  19. Ruiz-Pesini, E.; Lapeña, A.C.; Díez, C.; Alvarez, E.; Enríquez, J.A.; López-Pérez, M.J. Seminal quality correlates with mitochondrial functionality. Clin. Chim. Acta 2000, 300, 97–105.
  20. Marchetti, C.; Obert, G.; Deffosez, A.; Formstecher, P.; Marchetti, P. Study of mitochondrial membrane potential, reactive oxygen species, DNA fragmentation and cell viability by flow cytometry in human sperm. Hum. Reprod. 2002, 17, 1257–1265.
  21. Gallon, F.; Marchetti, C.; Jouy, N.; Marchetti, P. The functionality of mitochondria differentiates human spermatozoa with high and low fertilizing capability. Fertil. Steril. 2006, 86, 1526–1530.
  22. Espinoza, J.A.; Schulz, M.A.; Sánchez, R.; Villegas, J.V. Integrity of mitochondrial membrane potential reflects human sperm quality. Andrologia 2009, 41, 51–54.
  23. Ruiz-Pesini, E.; Díez-Sánchez, C.; López-Pérez, M.J.; Enríquez, J.A. The role of the mitochondrion in sperm function: Is there a place for oxidative phosphorylation or is this a purely glycolytic process? Curr. Top Dev. Biol. 2007, 77, 3–19.
  24. Storey, B.T. Mammalian sperm metabolism: Oxygen and sugar, friend and foe. Int. J. Dev. Biol. 2008, 52, 427–437.
  25. Zhu, Z.; Umehara, T.; Okazaki, T.; Goto, M.; Fujita, Y.; Hoque, M.; Kawai, T.; Zeng, W.; Shimada, M. Gene Expression and Protein Synthesis in Mitochondria Enhance the Duration of High-Speed Linear Motility in Boar Sperm. Front. Physiol. 2019, 10, 252.
  26. Odet, F.; Duan, C.; Willis, W.D.; Goulding, E.H.; Kung, A.; Eddy, E.M.; Goldberg, E. Expression of the gene for mouse lactate dehydrogenase C (Ldhc) is required for male fertility. Biol. Reprod. 2008, 79, 26–34.
  27. Koppers, A.J.; De Iuliis, G.N.; Finnie, J.M.; McLaughlin, E.A.; Aitken, R.J. Significance of mitochondrial reactive oxygen species in the generation of oxidative stress in spermatozoa. J. Clin. Endocrinol. Metab. 2008, 93, 3199–3207.
  28. Halliwell, B. Antioxidant defence mechanisms: From the beginning to the end (of the beginning). Free Radic. Res. 1999, 31, 261–272.
  29. Agarwal, A.; Virk, G.; Ong, C.; du Plessis, S.S. Effect of oxidative stress on male reproduction. World J. Mens Health 2014, 32, 1–17.
  30. Griveau, J.E.; Renard, P.; Le Lannou, D. An in vitro promoting role for hydrogen peroxide in human sperm capacitation. Int. J. Androl. 1994.
  31. Aitken, R.J. Free radicals, lipid peroxidation and sperm function. Reprod. Fertil. Dev. 1995, 7, 659–668.
  32. Kodama, H.; Yamaguchi, R.; Fukuda, J.; Kasai, H.; Tanaka, T. Increased oxidative deoxyribonucleic acid damage in the spermatozoa of infertile male patients. Fertil. Steril. 1997, 68, 519–524.
  33. Moraes, C.R.; Meyers, S. The sperm mitochondrion: Organelle of many functions. Anim. Reprod. Sci. 2018, 194, 71–80.
  34. Losano, J.D.A.; Angrimani, D.S.R.; Ferreira Leite, R.; Simões da Silva, B.D.C.; Barnabe, V.H.; Nichi, M. Spermatic mitochondria: Role in oxidative homeostasis, sperm function and possible tools for their assessment. Zygote 2018, 26, 251–260.
  35. Chen, L.B. Mitochondrial membrane potential in living cells. Annu. Rev. Cell Biol. 1988, 4, 155–181.
  36. Garner, D.L.; Thomas, C.A.; Joerg, H.W.; DeJarnette, J.M.; Marshall, C.E. Fluorometric assessments of mitochondrial function and viability in cryopreserved bovine spermatozoa. Biol. Reprod. 1997, 57, 1401–1406.
  37. Piccoli, C.; Scrima, R.; D’Aprile, A.; Ripoli, M.; Lecce, L.; Boffoli, D.; Capitanio, N. Mitochondrial dysfunction in hepatitis C virus infection. Biochim. Biophys. Acta 2006, 1757, 1429–1437.
  38. Perry, S.W.; Norman, J.P.; Barbieri, J.; Brown, E.B.; Gelbard, H.A. Mitochondrial membrane potential probes and the proton gradient: A practical usage guide. Biotechniques 2011, 50, 98–115.
  39. Gillan, L.; Evans, G.; Maxwell, W.M. Flow cytometric evaluation of sperm parameters in relation to fertility potential. Theriogenology 2005, 63, 445–457.
  40. Graham, J.K.; Kunze, E.; Hammerstedt, R.H. Analysis of sperm cell viability, acrosomal integrity, and mitochondrial function using flow cytometry. Biol. Reprod. 1990, 43, 55–64.
  41. Paoli, D.; Gallo, M.; Rizzo, F.; Baldi, E.; Francavilla, S.; Lenzi, A.; Lombardo, F.; Gandini, L. Mitochondrial membrane potential profile and its correlation with increasing sperm motility. Fertil. Steril. 2011, 95, 2315–2319.
  42. Garner, D.L.; Thomas, C.A. Organelle-specific probe JC-1 identifies membrane potential differences in the mitochondrial function of bovine sperm. Mol. Reprod. Dev. 1999, 53, 222–229.
  43. Gravance, C.G.; Garner, D.L.; Baumber, J.; Ball, B.A. Assessment of equine sperm mitochondrial function using JC-1. Theriogenology 2000, 53, 1691–1703.
  44. Martinez-Pastor, F.; Johannisson, A.; Gil, J.; Kaabi, M.; Anel, L.; Paz, P.; Rodriguez-Martinez, H. Use of chromatin stability assay, mitochondrial stain JC-1, and fluorometric assessment of plasma membrane to evaluate frozen-thawed ram semen. Anim. Reprod. Sci. 2004, 84, 121–133.
  45. Cheuquemán, C.; Bravo, P.; Treulén, F.; Giojalas, L.; Villegas, J.; Sánchez, R.; Risopatrón, J. Sperm membrane functionality in the dog assessed by flow cytometry. Reprod. Domest. Anim. 2012, 47, 39–43.
  46. Cheuquemán, C.; Merino, O.; Giojalas, L.; Von Baer, A.; Sánchez, R.; Risopatrón, J. Assessment of sperm function parameters and DNA fragmentation in ejaculated alpaca sperm (Lama pacos) by flow cytometry. Reprod. Domest. Anim. 2013, 48, 447–453.
  47. Lugli, E.; Troiano, L.; Cossarizza, A. Polychromatic analysis of mitochondrial membrane potential using JC-1. Curr. Protoc. Cytom. 2007.
  48. Amaral, A.; Ramalho-Santos, J.; St John, J.C. The expression of polymerase gamma and mitochondrial transcription factor A and the regulation of mitochondrial DNA content in mature human sperm. Hum. Reprod. 2007, 22, 1585–1596.
  49. Uribe, P.; Villegas, J.V.; Boguen, R.; Treulen, F.; Sánchez, R.; Mallmann, P.; Isachenko, V.; Rahimi, G.; Isachenko, E. Use of the fluorescent dye tetramethylrhodamine methyl ester perchlorate for mitochondrial membrane potential assessment in human spermatozoa. Andrologia 2017, 49.
  50. Darr, C.R.; Cortopassi, G.A.; Datta, S.; Varner, D.D.; Meyers, S.A. Mitochondrial oxygen consumption is a unique indicator of stallion spermatozoal health and varies with cryopreservation media. Theriogenology 2016, 86, 1382–1392.
  51. Evenson, D.P.; Darzynkiewicz, Z.; Melamed, M.R. Simultaneous measurement by flow cytometry of sperm cell viability and mitochondrial membrane potential related to cell motility. J. Histochem. Cytochem. 1982, 30, 279–280.
  52. Wang, X.; Sharma, R.K.; Gupta, A.; George, V.; Thomas, A.J.; Falcone, T.; Agarwal, A. Alterations in mitochondria membrane potential and oxidative stress in infertile men: A prospective observational study. Fertil. Steril. 2003, 80, 844–850.
  53. Troiano, L.; Granata, A.R.; Cossarizza, A.; Kalashnikova, G.; Bianchi, R.; Pini, G.; Tropea, F.; Carani, C.; Franceschi, C. Mitochondrial membrane potential and DNA stainability in human sperm cells: A flow cytometry analysis with implications for male infertility. Exp. Cell Res. 1998, 241, 384–393.
  54. Kasai, T.; Ogawa, K.; Mizuno, K.; Nagai, S.; Uchida, Y.; Ohta, S.; Fujie, M.; Suzuki, K.; Hirata, S.; Hoshi, K. Relationship between sperm mitochondrial membrane potential, sperm motility, and fertility potential. Asian J. Androl. 2002, 4, 97–103.
  55. Zhang, G.; Wang, Z.; Ling, X.; Zou, P.; Yang, H.; Chen, Q.; Zhou, N.; Sun, L.; Gao, J.; Zhou, Z.; et al. Mitochondrial biomarkers reflect semen quality: Results from the MARCHS study in Chongqing, China. PLoS ONE 2016, 11.
  56. Piasecka, M.; Kawiak, J. Sperm mitochondria of patients with normal sperm motility and with asthenozoospermia: Morphological and functional study. Folia. Histochem. Cytobiol. 2003, 41, 125–139.
  57. Pelliccione, F.; Micillo, A.; Cordeschi, G.; D’Angeli, A.; Necozione, S.; Gandini, L.; Lenzi, A.; Francavilla, F.; Francavilla, S. Altered ultrastructure of mitochondrial membranes is strongly associated with unexplained asthenozoospermia. Fertil. Steril. 2011, 95, 641–646.
  58. Amaral, A.; Paiva, C.; Attardo Parrinello, C.; Estanyol, J.M.; Ballescà, J.L.; Ramalho-Santos, J.; Oliva, R. Identification of proteins involved in human sperm motility using high throughput differential proteomics. J. Proteome. Res. 2014, 13, 5670–5684.
  59. Nowicka-Bauer, K.; Lepczynski, A.; Ozgo, M.; Kamieniczna, M.; Fraczek, M.; Stanski, L.; Olszewska, M.; Malcher, A.; Skrzypczak, W.; Kurpisz, M.K. Sperm mitochondrial dysfunction and oxidative stress as possible reasons for isolated asthenozoospermia. J. Physiol. Pharmacol. 2018, 69, 403–417.
More
Information
Subjects: Andrology
Contributors MDPI registered users' name will be linked to their SciProfiles pages. To register with us, please refer to https://encyclopedia.pub/register : , , , , ,
View Times: 338
Revisions: 2 times (View History)
Update Date: 30 Jun 2022
1000/1000
Video Production Service