首页 | 本学科首页   官方微博 | 高级检索  
相似文献
 共查询到18条相似文献,搜索用时 46 毫秒
1.
Many lizards and all snakes flick their tongues. It is known that this unique behavioral pattern serves to collect airborne and substrate chemicals which give the animal information via Jacobson's Organ about the location of food, conspecifics, and possibly other environmental factors. However, a comparative topographic analysis of tongue movements in squamate reptiles is lacking, and it might shed light on the evolution of this behavior. In this study, a survey was made of the lizards and snakes which tongue-flick. Observations and films were made of 25 lizard species representing 10 families and 30 snake species representing 5 families. The information from observations and film analyses of representative species was used to hypothesize the steps of the evolution of tongue-flicking from the simple downward extensions of primitive lizards to the complex multiple oscillations of snakes.  相似文献   

2.
The general histology and ultrastructure of the tongue and anterior process of the sublingual plica of four Taiwanese venomous snakes, the Chinese cobra (Naja naja atra), banded krait (Bungarus multicinctus), Taiwan habu (Trimeresurus mucrosquamatus), and bamboo snake (Trimeresurus stejnegeri stejnegeri) are described. The tongue fork exhibits a mid-dorsal invagination that broadens gradually toward its base. No mid-ventral invagination is observed. The epithelial cells on both dorsal and ventral aspects of the tongue fork have large and small microfacets, micropores and microvilli. The cell size, distribution pattern of the large microfacets, and the number of small microfacets present on both sides of the fork are essentially the same within a species, but vary among species. The function of these ultrastructures on the cell surface might be for the capture of chemical substances. The large microfacets are raised areas of the cell membrane, each with a pale granule contained within. The chemical nature of the pale granule is not yet known. The small pores surrounding the large microfacets are shallow hollows left after the release of the pale granules from the microfacets. The basic histological pattern of the tongue fork of these species is similar, being composed of a mucosal layer outside and dense musculature inside. No taste buds are discernible. The anterior processes are concave-like expansions of the anteriormost portions of the sublingual plicae. The oblique folds and micropapillae of this organ might be helpful for receiving the chemicals collected on the tongue, when the tongue makes contact with the elevated processes. The elevated processes may penetrate the ducts of Jacobson's organs to effect the final transfer.  相似文献   

3.
In almost all mammals a well developed, paired and blind ending vomeronasal Organ (VNO) situated within the basement of the nasal septum, communicates with the oral cavity. This contact is established by two nasopalatine ducts, which penetrate the rostral palate close to the incisors. These ducts open orally into the sulcus which moulds the palatine papilla. In several mammals taste buds were found in the epithelium of the patatine papilla located within the nasopalatine ducts or close to their oral openings. Presumably these taste buds interact with the vomeronasal olfaction. It is likely that they are leading to a chemosensory sensation comparable to the combination of normal taste and smell. As not all mammals with a functionable VNO possess taste buds in this position, an inspection of the rostral part of the tongue which touches the palatine papilla presented an interesting situation concerning the distribution of taste buds. This region of the tongue is almost completely free of taste buds in species like Tupaia glis and Didelphis marsupialis virginiana, which have taste buds in the epithelium of their palatine papilla. In Lemur catta however, where the palatine papilla is lacking taste buds, the respective tongue part is densely covered with them. In this case it appears likely that they in a way of substitution functionally are connected with vomeronasal olfaction.  相似文献   

4.
To understand the mechanisms for introducing urine or vaginal secretions into the vomeronasal organ, we used 16 mm cinematography and a freeze frame/slow motion technique to analyze the mouth and tongue movements of Brahman bulls while they examined the vulvas of restrained, estrogen-primed cows. Prior to flehmen, the mouth slowly opened, the curled tip of the tongue compressed the hard palate and the body of the tongue protruded from the mouth. The tongue maintained this form and moved forward. Once the tip of the tongue reached the incisive papilla, the body of the tongue retracted and the tip of the tongue relaxed. This tongue compression stroke (TCS) of the hard palate occurred 2 to 6 times, lasting 1 4 to 1 2 sec/stroke. Pressure changes in the vomeronasal organ are assumed to occur during and following TCSs, resulting in aspiration of any liquid in the incisive pit into the incisive and vomeronasal ducts. Such aspiration probably does not occur during flehmen because the tongue is relaxed and on the floor of the mouth.  相似文献   

5.
Most snakes ingest and transport their prey via a jaw ratchetingmechanism in which the left and right upper jaw arches are advancedover the prey in an alternating, unilateral fashion. This unilateraljaw ratcheting mechanism differs greatly from the hyolingualand inertial transport mechanisms used by lizards, both of whichare characterized by bilaterally synchronous jaw movements.Given the well-corroborated phylogenetic hypothesis that snakesare derived from lizards, this suggests that major changes occurredin both the morphology and motor control of the feeding apparatusduring the early evolution of snakes. However, most previousstudies of the evolution of unilateral feeding mechanisms insnakes have focused almost exclusively on the morphology ofthe jaw apparatus because there have been very few direct observationsof feeding behavior in basal snakes. In this paper I describethe prey transport mechanisms used by representatives of twofamilies of basal snakes, Leptotyphlopidae and Typhlopidae.In Leptotyphlopidae, a mandibular raking mechanism is used,in which bilaterally synchronous flexions of the lower jaw serveto ratchet prey into and through the mouth. In Typhlopidae,a maxillary raking mechanism is used, in which asynchronousratcheting movements of the highly mobile upper jaws are usedto drag prey through the oral cavity. These findings suggestthat the unilateral feeding mechanisms that characterize themajority of living snakes were not present primitively in Serpentes,but arose subsequently to the basal divergence between Scolecophidiaand Alethinophidia.  相似文献   

6.
BACKGROUND: Gamma-aminobutyric acid is an inhibitory neurotransmitter, synthesized by two isoforms of glutamate decarboxylase (GAD), GAD65 and -67. Unexpectedly, inactivation of GAD67 induces cleft palate in mice. Reduction of spontaneous tongue movement resulting from decreased motor nerve activity has been related to the development of cleft palate in GAD67(-/-) fetuses. In the present study, development of cleft palate was examined histologically and manipulated with culture of the maxilla and partial resection of fetal tongue. METHODS: GAD67(-/-) mice and their littermates were used. Histological examination and immunohistochemistry were performed conventionally. Organ culture of the maxilla was carried out as reported previously. Fetuses were maintained alive under anesthesia and tips of their tongues were resected. RESULTS: Elevation of palatal shelves, the second step of palate formation, was not observed in GAD67(-/-) mice. In wild-type mice, GAD67 and gamma-aminobutyric acid were not expressed in the palatal shelves, except in the medial edge epithelium. During 2 days of culture of maxillae dissected from E13.5-E14.0 GAD67(-/-) fetuses, elevation and fusion of the palatal shelves were induced. When E13.5-15.5 mutant fetuses underwent partial tongue resection, the palatal shelves became elevated within 30 min. CONCLUSIONS: These results suggest that the potential for palate formation is maintained in the palatal shelves of GAD67(-/-) fetuses, but it is obstructed by other, probably neural, factors, resulting in cleft palate.  相似文献   

7.
David  Cundall 《Journal of Zoology》1995,237(3):353-376
Cylindrophis ruffus ingests prey using two distinct mechanisms. During initial phases of prey transport, lateral movements of the rear of the braincase combine with small unilateral movements of the toothed bones of each side; prey is usually constricted during this phase to permit the snake to push its head over the prey. Once transport has carried the leading part of the prey into the anterior oesophagus, Cylindrophis begins to use bilaterally synchronized movements of the jaw apparatus combined with low-amplitude, short wave-length flexions of the anterior vertebral column. Transport of prey is many times faster during the bilateral phase than during the unilateral phase.
Radiographic and cinematographic evidence indicates that the mandibular tips of Cylindrophis do not separate more than 1.5–2.0 times the resting distance between the dentary tips. Although this limits potential gape size, the intramandibular joint is highly mobile, allowing the mandibles to conform to a variety of prey shapes. Manipulations of anaesthetized and fresh, dead specimens revealed that the palatomaxillary arches are tightly attached to the ventral bones of the snout, movements of each arch being reflected in equivalent movements of the ipsilateral elements of the snout.
Cylindrophis represents a functional stage intermediate between most lizards with limited palatomaxillary kinesis and advanced snakes with considerable palatomaxillary mobility. Contrary to previous hypotheses, however, upper jaw liberation in Cylindrophis is due to liberation of the ventral snout, not to reduction of attachments to the braincase and snout. This suggests that the nose played a crucial role in the evolution of the feeding apparatus in alethinophidian snakes.  相似文献   

8.
In the late 19th Century, the choanae (or internal nares) of the Plesiosauria were identified as a pair of palatal openings located rostral to the external nares, implying a rostrally directed respiratory duct and air path inside the rostrum. Despite obvious functional shortcomings, this idea was firmly established in the scientific literature by the first decade of the 20th Century. The functional consequences of this morphology were only re-examined by the end of the 20th Century, leading to the conclusion that the choanae were not involved in respiration but instead in underwater olfaction, the animals supposedly breathing with the mouth agape. Re-evaluation of the palatal and internal cranial anatomy of the Plesiosauria reveals that the traditional identification of the choanae as a pair of fenestrae situated rostral to the external nares appears erroneous. These openings more likely represent the bony apertures of ducts that lead to internal salt glands situated inside the maxillary rostrum. The 'real' functional choanae (or caudal interpterygoid vacuities), are situated at the caudal end of the bony palate between the sub-temporal fossae, as was suggested in the mid-19th Century. The existence of a functional secondary palate in the Plesiosauria is therefore strongly supported, and the anatomical, physiological, and evolutionary implications of such a structure are discussed.  相似文献   

9.
Pumping is a vital natural process, imitated by humans for thousands of years. We demonstrate that a hitherto undocumented mechanism of fluid transport pumps nectar onto the hummingbird tongue. Using high-speed cameras, we filmed the tongue–fluid interaction in 18 hummingbird species, from seven of the nine main hummingbird clades. During the offloading of the nectar inside the bill, hummingbirds compress their tongues upon extrusion; the compressed tongue remains flattened until it contacts the nectar. After contact with the nectar surface, the tongue reshapes filling entirely with nectar; we did not observe the formation of menisci required for the operation of capillarity during this process. We show that the tongue works as an elastic micropump; fluid at the tip is driven into the tongue''s grooves by forces resulting from re-expansion of a collapsed section. This work falsifies the long-standing idea that capillarity is an important force filling hummingbird tongue grooves during nectar feeding. The expansive filling mechanism we report in this paper recruits elastic recovery properties of the groove walls to load nectar into the tongue an order of magnitude faster than capillarity could. Such fast filling allows hummingbirds to extract nectar at higher rates than predicted by capillarity-based foraging models, in agreement with their fast licking rates.  相似文献   

10.
1. In Homo and the great apes (Pongidae) there occurs, besides the plica sublingualis a plica fimbriata at the ventral surface of the tongue. This duplicature of the mucosa does not occur in the Hylobytidae and in the other primates. 2. Some taste buds could be found in the epithelium of the plica sublingualis of the Pongidae. 3. There are many taste buds in the epithelium of the plica fimbriata of the Pongidae. On this sublingual structure there were counted 1776 taste buds in Pongo, 592 in Gorilla and 280 in Pan. A few taste buds could also be found on the plica fimbriata of a human newborn. 4. A glandula apicis linguae occurs in Homo, Pan, Gorilla and Pongo. 5. The fresh saliva of the glandula apicis linguae and the saliva on the floor of the mouth can be tested by the taste buds in the epithelium of the plica fimbriata, of papillae lenticulares and of areae gustatoriae at the ventral surface of the tongue. 6. It might be the function of the sublingual taste buds to taste the fresh saliva as a gradient for the central nervous comparison with the taste of the saliva on the dorsal surface of the tongue. 7. Because of the complete absence of a sublingua in the Platyrrhini and in the Cercopithecinae it is unlikely that the plica fimbriata of Homo and the great apes can be interpreted as a homalogon of the sublingua in the prosimians. 8. Because of the absence of a sublingua in other ordines of the Mammalia (Insectivora, Carnivora, Rodentia, Chiroptera, Ungulata) it is unlikely as well that the sublingua in the prosimians can be interpreted as a homologon of the tongues of the lower vertebrates. The sublingual structures occuring in the Marsupialia have to be investigated. 9. Because of these reasons the new development of the sublingua in the prosimians and the plica fimbriata in the Hominoidea, in complete independence from one another, seems to be a better explanation of the 2 structures and less contradictionary to anatomical and phylogenetic arguments. The different function of both structures in the recent primates gives a hint for the possible reason for their development during the process of evolution.  相似文献   

11.
The fluid mechanics of bolus ejection from the oral cavity   总被引:1,自引:0,他引:1  
The squeezing action of the tongue against the palate provides driving forces to propel swallowed material out of the mouth and through the pharynx. Transport in respose to these driving forces, however, is dependent on the material properties of the swallowed bolus. Given the complex geometry of the oral cavity and the unsteady nature of this process, the mechanics governing the oral phase of swallowing are not well understood. In the current work, the squeezing flow between two approaching parallel plates is used as a simplified mathematical model to study the fluid mechanics of bolus ejection from the oral cavity. Driving forces generated by the contraction of intrinsic and extrinsic lingual muscles are modeled as a spatially uniform pressure applied to the tongue. Approximating the tongue as a rigid body, the motion of tongue and fluid are then computed simultaneously as a function of time. Bolus ejection is parameterized by the time taken to clear half the bolus from the oral cavity, t1/2. We find that t1/2 increases with increased viscosity and density and decreases with increased applied pressure. In addition, for low viscosity boluses (μapproximately 1000 cP), viscosity dominates. A transition region between these two regimes is found in which both properties affect the solution characteristics. The relationship of these results to the assessment and treatment of swallowing disorders is discussed.  相似文献   

12.
Water drinking in the mallard is accomplished by a fine-tuned set of movements of upper and lower jaw and of the tongue. During immersion of the tips of the bill, the oral cavity is formed into smaller volumes containing water and into connecting tubes. Two mechanisms serve the water transport: (1) lingual and jaw movements press water from the water-containing spaces into the tubes; (2) a quantitative simulation of the shape of the oral cavity during immersion shows that the two tubes are so narrow that capillary action also contributes to water transport. Thereafter, the tips of the bill are raised until they point upward. In this “tip-up” position, water flows into the esophagus because of gravity. We conclude that, in addition to normal tip-up drinking observed in almost all Passeriformes and Galliformes, a second type of tip-up drinking may be distinguished in Anseriformes. The integration of the drinking mechanism, keeping the water inside the mouth, and the straining mechanism, expelling the water along the beak rims, is effected by specific actions of the elaborate lingual apparatus.  相似文献   

13.
This study is concerned with reconciling theoretical modelling of the fluid flow in the airway surface liquid with experimental visualisation of tracer transport in human airway epithelial cultures. The airways are covered by a dense mat of cilia of length ∼ 6 μm beating in a watery periciliary liquid (PCL). Above this there is a layer of viscoelastic mucus which traps inhaled pathogens. Cilia propel mucus along the airway towards the trachea and mouth. Theoretical analyses of the beat cycle smithd, fulb predict small transport of PCL compared with mucus, based on the assumption that the epithelium is impermeable to fluid. However, an experimental study coord indicates nearly equal transport of PCL and mucus. Building on existing understanding of steady advection-diffusion in the ASL (Blake and Gaffney, 2001; Mitran,2004) numerical simulation of an advection-diffusion model of tracer transport is used to test several proposed flow profiles and to test the importance of oscillatory shearing caused by the beating cilia. A mechanically derived oscillatory flow with very low mean transport of PCL results in relatively little ‘smearing’ of the tracer pulses. Other effects such as mixing between the PCL and mucus, and significant transport in the upper part of the PCL above the cilia tips are tested and result in still closer transport, with separation between the tracer pulses in the two layers being less than 9%. Furthermore, experimental results may be replicated to a very high degree of accuracy if mean transport of PCL is only 50% of mucus transport, significantly less than the mean PCL transport first inferred on the basis of experimental results.  相似文献   

14.
Invasion of black-tailed deer (Odocoileus hemionus columbianus) by larvae of the nose bot flies (Cephenemyia apicata and C. jellisoni) was investigated by obssrving their expulsion by larviparous females and their subsequent activity on the host. The drying uterine fluid encasing larvae ( = larval packet) delays desiccation, ensures adhesion to hairs, and immediately dissolves upon contact with saliva. Contrary to the widely accepted nasal mode of invasion, larvae placed on muzzles of deer crawl ventrally toward the upper lip, enter the mouth, and then crawl caudally along the hard palate or tongue toward the throat. Hair located between the muzzle and nostril prevents larvae from entering the nostrils. A natural per os mode of invasion, heretofore unrecognized, is proposed. This is initiated by: (1) females depositing larval packets on the muzzle of deer, or (2) deer licking larval packets from contaminated areas around the muzzle. The positive thermotropism of larvae is compatible with such a per os mode of entry into the host.  相似文献   

15.
16.
Summary The scincid lizardTiliqua rugosa possesses a large external nasal gland which is located intraconchally. Highly ramified tubules, imbedded primarily in the periphery of the gland, unite to form collecting ducts which empty into a short excretory canal. The diameter of the tubules increases progressively from 30. at the distal extremity of the gland to over 200 at the level of the collecting ducts. The intraglandular portion of the excretory canal is often dilated to form an ampulla. The thickness of the epithelium increases from 12 at the level of the tubules to 25–30 in the excretory canal.The excretory canal is lined with an epidermal epithelium close to the point where it enters the vestibule. In all the rest of the gland the tubules are lined with two cell types: large, typical muco-serous cells and striated cells. At the distal end of the tubules the striated cells are narrow and poorly differentiated and alternate more-or-less regularly with the muco-serous cells. The relative proportion of these striated cells increases progressively, as does their size, as one moves proximally down the tubule. In the gland as a whole the striated cells are approximately twice as numerous as the muco-serous cells but, due to their smaller size, they occupy less than one third of the tubular volume.Electron microscopy of the striated cells ofTiliqua rugosa revealed the presence of extensive lateral interdigitations and expansions of the basal cytoplasmic membrane, anatomical specialisations which are normally indicative of active salt transport. These modifications are less marked however than in the external nasal glands of the lizardsLacerta muralis andVaranus griseus, which do not appear to function as salt glands. In addition there are few mitochondria present, although they are of large size. The combination of these ultrastructural features, plus the fact that the striated cells are intermixed with muco-serous cells in the tubules, makes it most unlikely that the external nasal gland ofTiliqua rugosa is capable of elaborating an hyperosmotic fluid. What is more, this has never been conclusively demonstrated in this species in physiological studies.The progressive specialisation of the striated cells from the distal to the proximal section of the tubules poses the problem of the origin and differentiation of this cell type.A review of results obtained from the study ofTiliqua rugosa and other species of lizards shows that the nature of the relationship between structure and function of the external nasal gland is far from clear. The existence of salt glands, capable of excreting hyperosmotic solutions, is invariably linked with the presence in the gland of well-developed striated segments composed almost entirely of cells possessing extensive interdigitations of the lateral membranes. Amongst terrestrial lizards, nasal salt glands are usually found in herbivorous species and they are primarily adapted to the extrarenal excretion of potassium ions. The problem for carnivorous species is more often that of an excess of sodium rather than potassium ions and with the possible exceptionAcanthodactylus species, functional nasal salt glands have not been demonstrated in terrestrial carnivores, despite the presence in some cases of well-developed striated segments in the gland having a similar structure to those found in herbivores. In humid regions, carnivorous lizards probably never require extrarenal excretory mechanism and in arid regions their survival is assured by their capacity to tolerate hypernatraemia when confronted with excessive salt loads. Salt glands capable of eliminating sodium ions to any extent have only been described in two littoral species, an herbivorous iguanid and a carnivorous varanid. Unfortunately the structure of their respective nasal glands has not yet been described and their further study would be desirable.  相似文献   

17.
Dyneins are highly complex, multicomponent, microtubule-based molecular motors. These enzymes are responsible for numerous motile behaviors in cytoplasm, mediate retrograde intraflagellar transport (IFT), and power ciliary and flagellar motility. Variants in multiple genes encoding dyneins, outer dynein arm (ODA) docking complex subunits, and cytoplasmic factors involved in axonemal dynein preassembly (DNAAFs) are associated with human ciliopathies and are of clinical interest. Therefore, clear communication within this field is particularly important. Standardizing gene nomenclature, and basing it on orthology where possible, facilitates discussion and genetic comparison across species. Here, we discuss how the human gene nomenclature for dyneins, ODA docking complex subunits, and DNAAFs has been updated to be more functionally informative and consistent with that of the unicellular green alga Chlamydomonas reinhardtii, a key model organism for studying dyneins and ciliary function. We also detail additional nomenclature updates for vertebrate-specific genes that encode dynein chains and other proteins involved in dynein complex assembly.

IntroductionDynein family motor proteins form multiple different dynein complexes in mammals, with important roles in a wide range of cellular functions (King, 2017; Osinka et al., 2019; Roberts, 2018). Dyneins can be broadly classified into two groups: cytoplasmic and axonemal. Dynein complexes “walk” toward the minus ends of microtubules; while doing so, they can transport a variety of cargoes within cells (Trokter et al., 2012). The motor activity of these complexes allows them to play key roles in enabling motility of whole cells, generating fluid flow across cell surfaces, and transporting organelles and other components within the cytoplasm.Dynein subunits are classified by mass into four categories: heavy (∼520 kD), intermediate (∼70 –140 kD), light intermediate (∼53–59 kD), and light (∼10–30 kD) chains (Pfister et al., 2006). The heavy and intermediate chains are specific to certain dynein complexes, while the light chains may be components of both cytoplasmic and axonemal dynein machinery, and in some cases, nondynein complexes. The light intermediate chains are present only in the cytoplasmic dynein class.Dynein-based movement is powered by the ATP-driven dynein heavy chain subunits (Schmidt and Carter, 2018). 15 genes in the human genome encode dynein heavy chains: 1 for each of the 2 cytoplasmic dynein complexes and 13 that encode heavy chain components of the various axonemal dynein complexes. A dynein-related gene, DNHD1 (dynein heavy chain domain 1) has been referred to as a “ghost gene”: it may be a remnant of an earlier duplication that has not decayed at a normal rate, as a truncated version might poison cytoplasmic dynein heavy chain dimerization and thus be lethal (Gibbons, 2018; Schmidt and Carter, 2018). DNHD1 is currently classified as an “orphan” dynein heavy chain-encoding gene (Kollmar, 2016) but may be in the process of becoming a pseudogene (Wickstead, 2018).Cytoplasmic dyneinsDynein 1 complexThe cytoplasmic dynein 1 complex (Table S1) is present throughout eukaryotes, with some notable exceptions such as green plants and red algae (Wickstead and Gull, 2007). It is involved in a wide variety of intracellular transport activities, transporting cargoes including chromosomes, mRNA, and protein complexes (Reck-Peterson et al., 2018). The dynein 1 complex also acts in cell division, helping to form and orient the mitotic spindle (Torisawa and Kimura, 2020), establish cell polarity (Lu and Gelfand, 2017), and position organelles (Allan, 2014; Oyarzún et al., 2019; Palmer et al., 2009).A dimer of DYNC1H1-encoded heavy chains forms the core of the cytoplasmic dynein 1 complex (Fig. 1 a) and acts as its ATPase motor (Palmer et al., 2009; Pfister et al., 2005). Each heavy chain contains six AAA+ domains, an antiparallel coiled-coil region with a microtubule-binding domain at its tip, and a C-terminal domain (Bhabha et al., 2016; Carter, 2013; Reck-Peterson et al., 2018; Roberts et al., 2013). Immediately N-terminal of the AAA1 domain is a linker that traverses the plane of the AAA ring and changes conformation during the ATPase cycle to drive motor activity. AAA1 exhibits ATP hydrolytic activity, acting as an ATPase and powering the dynein motor complex (Silvanovich et al., 2003), while nucleotide binding at several other AAA domains appears to modify how conformational change propagates through the AAA ring and affects microtubule-binding activity. The coordinated activity of both heavy chains within the dynein complex is required for processivity (Reck-Peterson et al., 2006).Open in a separate windowFigure 1.Cytoplasmic dynein complexes. (a) The cytoplasmic dynein 1 complex. The DYNC1H1 protein heavy chains have large globular heads at the C-termini that are composed of a ring of six AAA+ domains. The microtubule-binding domains are located at the tips of antiparallel coiled coils that derive from AAA4. The linker/N-terminal domains connect the AAA rings and the intermediate and light chains. (b) The cytoplasmic dynein 2 complex. The DYNC2H1 protein heavy chains power retrograde IFT and have the same general domain organization as DYNC1H1. However, the tails of the two heavy chains fold differently due to an asymmetry imposed by the two different intermediate chains: one is straight while the other forms a zigzag shape and interacts with the IFT-B train (Toropova et al., 2019). The linker/N-terminal domain connects the AAA ring and the intermediate and light chains. *It remains unknown whether the DYNLT2B protein forms a homodimer or a heterodimer with another Tctex-type light chain. (c) Schematic showing the interaction between the dynein 1 and dynactin complexes. The adapter molecule affects the type of cargo bound; in this figure, the hook microtubule tethering protein 3 (HOOK3)–encoded protein is acting as a cargo adapter.The intermediate chains of metazoan dynein 1 connect it to another multi-subunit complex known as dynactin (Loening et al., 2020). Dynactin is built around a filament of the protein encoded by ACTR1A (actin-related protein 1A). It activates dynein and regulates its binding to vesicles and organelles to be transported (Ketcham and Schroer, 2018). A coiled coil–containing cargo adaptor protein is required for dynein 1 activation (Fig. 1 c). A single adaptor protein sandwiches between dynactin and dynein, where it interacts with the dynein heavy chain tails and the light intermediate chain and along the length of the dynactin complex (Gonçalves et al., 2019; Reck-Peterson et al., 2018). There are currently ≥12 known cargo adaptor proteins, which are encoded by HOOK1, HOOK2, HOOK3, BICD2, BICDL1, BICDL2, RAB11FIP3, RASEF, CRACR2A, NIN, NINL, and SPDL1 (Barisic et al., 2010; Casenghi et al., 2005; Dona et al., 2015; Gonçalves et al., 2019; Horgan et al., 2010; Lee et al., 2018; Loening et al., 2020; Olenick et al., 2016; Vallee et al., 2021; Wang et al., 2019). Other protein cofactors may also be required for dynein recruitment to their cargoes. For example, the protein encoded by PAFAH1B1 (platelet activating factor acetylhydrolase 1b regulatory subunit 1; HUGO Gene Nomenclature Committee [HGNC] ID: 8574), also published using the alias LIS1 (lissencephaly 1) is required along with dynactin and BICD2 for dynein 1 to traffic many cargoes, such as nuclei, along microtubules (Faulkner et al., 2000; Splinter et al., 2012). LIS1 has most recently been suggested to stabilize the “open” conformation of cytoplasmic dynein 1 such that the heavy chains are able to undergo a mechanochemical cycle and cannot adopt the autoinhibited or “closed” state where movement of key mechanical elements is abrogated by interactions between heavy chains (Markus et al., 2020).The dynein light chains can be divided into three subfamilies: the t-complex associated (Tctex)–type family (encoded by DYNLT1, DYNLT2, DYNLT2B, DYNLT3, DYNLT4, and DYNLT5), the LC8-type family (encoded by DYNLL1, DYNLL2, and DNAL4), and the roadblock-type family (encoded by DYNLRB1 and DYNLRB2; Bowman et al., 1999; King et al., 1998; King et al., 1996; King and Patel-King, 1995). Most of these protein chains can be found in both cytoplasmic dynein complexes: the exceptions are DYNLT2, which is an axonemal dynein subunit; DYNLT2B, which is found in the dynein 2 complex and the I1/f inner dynein arm (IDA); DYNLT4 and DYNLT5, which are not well characterized; and DNAL4, which is present only in outer dynein arms (ODAs).Several proteins originally identified as dynein light chains are also found in numerous multimeric complexes unrelated to dyneins and appear to act as general dimerization engines or hubs (Williams et al., 2018). The LC8-type light chains (DYNLL1 and DYNLL2) are present in many enzymes including myosin V (Benashski et al., 1997; Espindola et al., 2000) and neuronal nitric oxide synthase (Jaffrey and Snyder, 1996). They also play a role in regulating apoptosis via an interaction with the BCL2 family protein encoded by BCL2L11 (Puthalakath et al., 1999). The DYNLT1 protein has reported roles in actin remodeling and neurite outgrowth (Chuang et al., 2005) and hypocretin signaling (Duguay et al., 2011). DYNLRB1 interacts with Rab6 family member proteins in the Golgi apparatus (Wanschers et al., 2008), and both roadblock-type dynein light chains are reportedly involved in a TGFβ signaling pathway (Jin et al., 2009).Dynein 2 complexCilia are highly complex microtubule-based organelles that extend from the cell surface and can be classified as either primary (or nonmotile) or motile (Satir and Christensen, 2008). Most eukaryotic cells, excluding blood cells and those actively dividing, have an associated primary or nonmotile cilium. These act as sensory organelles, detecting a broad range of signaling molecules (Kopinke et al., 2021; Mykytyn and Askwith, 2017; Saternos et al., 2020).The dynein 2 complex (also known as the intraflagellar transport [IFT] dynein or cytoplasmic dynein 1b in Chlamydomonas reinhardtii; Table S2) is found only in cells with associated cilia or flagella, where it locates within and around the base of these structures (Höök and Vallee, 2006). IFT trains are multiprotein complexes required for the assembly and function of cilia and flagella in eukaryotes (Dutcher, 2019; Wingfield et al., 2017). The anterograde IFT motor complex kinesin 2 moves IFT trains and associated cargoes plus the dynein 2 complex along microtubules, from the base to the tip of a cilium or flagellum (Toropova et al., 2019; Vuolo et al., 2020). The retrograde IFT motor complex dynein 2 transports IFT trains and associated factors from the tip back to the base (Hou and Witman, 2015; Pazour et al., 1998). The dynein 2 complex is required for the assembly of cilia and flagella (Pazour et al., 1999; Pfister et al., 2006) and also has key roles in ciliary signaling functions (Vuolo et al., 2020).The core of dynein 2 is composed of a dimer of two DYNC2H1-encoded heavy chains (Fig. 1 b). The tails of these identical heavy chains are directed into two different conformations by the other subunits in the complex (Toropova et al., 2019). Each heavy chain is stabilized by its interaction with a DYNC2LI1-encoded protein subunit. The C-terminal helix of one of these light intermediate subunits associates with a DYNC2I1 (previously WDR60)-encoded protein with a DYNLRB-encoded subunit, to enforce a distinct conformation on one heavy chain (Toropova et al., 2019; Vuolo et al., 2020).The DYNC2I1- and DYNC2I2-encoded intermediate chains bind the heavy chains via their C-terminal β-propeller domains. The N-terminal regions of these intermediate chains are dimerized by three DYNLL1/2 dimers and one of each of the other light chain dimers: DYNLT1/3, DYNLRB1/2, and DYNLT2B (Toropova et al., 2019). The DYNLT2B-encoded light chain is a unique accessory component of the dynein 2 complex. Whether it forms a homodimer or heterodimer with another light chain remains to be confirmed, although there is evidence to suggest that, unlike the other light chains, the DYNLT2B subunit may be monomeric (DiBella et al., 2001). Recent structural studies of Tetrahymena ODAs have revealed a Tctex-family heterodimer (Rao et al., 2021).Axonemal dyneinsMotile cilia (sometimes termed flagella when they occur singly or in small numbers on a cell) are more restricted to certain cell types. Their movement enables sperm to swim (Linck et al., 2016), respiratory cilia on epithelial cells to sweep away mucus containing trapped pathogens (Hansson, 2019), and oviduct epithelial cells to waft an ovum along a fallopian tube toward the uterus (Spassky and Meunier, 2017). Multiciliated cells in the brain help move the cerebrospinal fluid and also influence neuronal migration (Brooks and Wallingford, 2014). In the male reproductive tract, the epithelial cells of the efferent ducts are densely covered with multiple motile cilia necessary for the transport of sperm cells (Aprea et al., 2021a). Motility of nodal cilia in the embryonic left–right organizer is necessary for the determination of correct left–right body asymmetry (Nonaka et al., 1998).An axoneme is the microtubule superstructure core of the cilium and contains many tightly associated components. A motile cilium has a highly conserved “9 + 2” structure: 9 microtubule doublets that surround a central pair of 2 microtubule singlets (the “central apparatus”; Fig. 2). Axonemal dyneins are the motor complexes that drive a sliding motion between ciliary doublet microtubules, enabling movement. Motile cilia have IDAs and ODAs and radial spokes that are thought to be involved in signal transduction between the central pair and the outer microtubule ring (Ishikawa, 2017). Nonmotile cilia have only the outer doublet ring and have a 9 + 0 microtubule arrangement, although the number of outer doublets decreases and their arrangement changes beyond the proximal part of the cilium (Kiesel et al., 2020).Open in a separate windowFigure 2.Axonemal dynein complexes. (a) Axonemal ODA. The blue text denotes subunits found in ODA complexes in respiratory cilia, and red text denotes subunits found in ODA complexes in sperm flagella. (b) Axonemal inner arm I1/f complex subunits (IDA). (c) Monomeric IDAs. Each inner arm species is constructed around a distinct monomeric heavy chain associated with an actin monomer and either DNALI1 or centrin; species d contains two additional components. In most cases, the precise equivalence between the human and C. reinhardtii monomeric heavy chain species is uncertain.The ODAs (Table S3 and Fig. 2 a) and IDAs (Table S4 and Table S5, and Fig. 2, b and c) in motile cilia are arranged in two rows with a complex 96-nm repeat organization. They are permanently attached to the A-tubule of one outer doublet microtubule (see Fig. 3, a and b) and transiently interact in an ATP-dependent manner with the B-tubule of the adjacent doublet to generate a sliding force (King, 2017). IDAs with a single heavy chain are termed monomeric (Table S4), while the I1/f IDA (Table S5) is dimeric, with two nonidentical heavy chains. These different types of dyneins vary in terms of their enzymatic and motor properties, likely reflecting their precise roles in the generation of ciliary motility (King, 2017).Open in a separate windowFigure 3.Organization of a mammalian motile cilium. (a) The diagram illustrates the general 9 + 2 microtubule arrangement within the ciliary axoneme. The inner and outer rows of dynein arms generate the force required for ciliary beating. The N-DRC complex is a key regulatory structure that interconnects the doublet microtubules. The radial spokes regulate the beat of cilia by transducing signals between the doublets and the central microtubule pair. (b) Tomographic image of an averaged 96-nm repeat for a single human ciliary doublet microtubule, revealing the microtubule-associated dynein arms, N-DRC, and radial spoke. The scale bar represents 25 nm. This image was generated by Jason Schrad (Nicastro laboratory) using data from Lin et al. (2014). (c and d) Cross-section (c) and longitudinal (d) views of the 48-nm repeat organization of a bovine doublet microtubule. The components of the ODA-DC are individually colored and indicated. This ribbon diagram was generated with the PyMol molecular graphics system (Schrödinger) using Protein Data Bank accession no. 7RRO (Gui et al., 2021).ODA docking complex (ODA-DC)The correct functioning of cilia and flagella in most eukaryotes is dependent on the ODA chains attaching to the outer doublet microtubules at 24-nm intervals (Dean and Mitchell, 2015; King, 2017). The ODA-DC facilitates binding and may also play a role in regulating the activity of the ODAs (Takada et al., 2002). The ODA-DC in C. reinhardtii consists of three protein subunits, encoded by DCC1 (DC1), DCC2 (DC2), and DLE3 (DC3). In mammals, it consists of five protein subunits (Gui et al., 2021) encoded by five genes, now named ODAD1, ODAD2, ODAD3, ODAD4, and CLXN (calaxin; Fig. 2 a and Fig. 3, b and c). CLXN (previously EFCAB1) has been assigned the alias symbol ODAD5, and authors may refer to it as such in publications if they wish, referencing the approved gene symbol at least once to aid data retrieval. Only ODAD1 and ODAD3 have orthologues in C. reinhardtii (DCC2 and DCC1, respectively).Dynein axonemal assembly factors (DNAAFs)Genes encoding proteins that act as axonemal dynein assembly factors are named using the root symbol DNAAF. These proteins play an important role in the preassembly of IDAs and ODAs in the cytoplasm before their transport to cilia (Fabczak and Osinka, 2019; King, 2021).Historically, the DNAAF root has been used only for proteins directly involved in the preassembly of axonemal dynein arms in the cytoplasm. We wrote to authors who have published on the genes that we are reporting in this publication as newly updated DNAAFs (see their symbols in bold in Table S6) and discussed this issue with our specialist advisors for this gene group (https://www.genenames.org/data/genegroup/#!/group/1627). This effort resulted in an agreement to use the term “DNAAF” more broadly. Therefore, a DNAAF symbol can now also be assigned to genes encoding proteins that play a role in trafficking dynein arms from the cytoplasm to cilia.Association with human phenotypesHumans have four described cilia types, and defects in all types are associated with various diseases: motile 9 + 2 cilia (e.g., respiratory cilia, ependymal cilia, sperm flagella); motile 9 + 0 cilia (e.g., nodal cilia); nonmotile 9 + 2 cilia (e.g., the kinocilium of hair cells and the proximal region of olfactory cilia); and nonmotile 9 + 0 cilia (e.g., renal monocilia and the connecting cilia of photoreceptor cells). Cilia are located on almost all polarized cell types of the human body; therefore, cilia-related disorders (ciliopathies) affect many organ systems (Fliegauf et al., 2007). Genetic mutations that impair cilia and/or flagella beating cause a heterogeneous group of rare disorders referred to as motile ciliopathies (Wallmeier et al., 2020). The pathogenic mechanisms, clinical symptoms, and severity of the diseases depend on the specific affected genes and the tissues in which they are expressed. Defects in ependymal cilia can result in hydrocephalus. Reduced fertility can be due to defective cilia in the fallopian tubes or the efferent ducts as well as sperm flagella. The malfunction of motile monocilia on the left–right organizer during early embryonic development can lead to laterality defects such as situs inversus and heterotaxy. Severe impairment of mucociliary clearance in the respiratory tract leads to chronic bronchial problems. Primary ciliary dyskinesia (PCD), which can present with a variety of these features, is the most common motile ciliopathy.The genetic disorder PCD is heterogeneous and has been linked to variants in genes encoding dyneins, axonemal dynein assembly factors, and ODA-DC subunits (Wallmeier et al., 2020). PCD-associated phenotypes include chronic respiratory problems, recurrent middle ear infections, male infertility, and subfertility in females (Leigh et al., 2019). Roughly 50% of PCD patients are diagnosed with Kartagener syndrome, a subtype defined by a triad of symptoms: chronic sinusitis, bronchiectasis, and situs inversus, where the positions of major body organs are reversed (Zariwala et al., 2011). Situs inversus totalis is observed when all thoracic and abdominal viscera are reversed; individuals with situs inversus or situs ambiguus show more variable organ positioning (seen in ≥6% of PCD cases; Kennedy et al., 2007; Sempou and Khokha, 2019).Table 1.Human phenotypes associated with variants of genes encoding dyneins and dynein-associated proteins
PhenotypeAssociated dynein or dynein-related gene variantsaSelected associated publications (PubMed ID)OMIM MIM number (phenotype subtype)
Primary ciliary dyskinesia (PCD): abnormal ciliary motility, respiratory distress, sinusitis, otitis media, bronchiectasis, laterality defects, infertility DNAH1 11371505 20301301 24360805617577 (CILD37)
DNAH5 11062149 11788826608644 (CILD3)
DNAH9 30471717 30471718618300 (CILD40)
DNAH11 12142464611884 (CILD7)
DNAI1 10577904604366 (CILD1)
DNAI2 18950741612444 (CILD9)
DNAL1 21496787614017 (CILD16)
NME8 (alias DNAI8 and TXNDC3)17360648610852 (CILD6)
ODAD1 23261302 23261303 23506398 30291279 32855706615067 (CILD20)
ODAD2 23849778 24203976 25186273615451 (CILD23)
ODAD3 24067530 25192045 25224326 30504913 31383820616037 (CILD30)
ODAD4 27486780617092 (CILD35)
DNAAF1 19944400 19944405 27261005613193 (CILD13)
DNAAF2 31107948 32638265 34785929612518 (CILD10)
DNAAF3 22387996 31186518606763 (CILD2)
DNAAF4 23872636615482 (CILD25)
DNAAF5 29358401 25232951 23040496614874 (CILD18)
DNAAF6 32170493300991 (CILD36)
ZMYND10 23604077 23891469 23891471615444 (CILD22)
LRRC6 23122589614935 (CILD19)
LRRC56 30388400618254 (CILD39)
SPAG1 24055112 26228299615505 (CILD28)
CFAP298 24094744615500 (CILD26)
CFAP300 29727692 29727693618063 (CILD38)
Spinal muscular atrophy (SMALED type 1): lower limb atrophy and weakness, mild to moderate cognitive impairment DYNC1H1 24307404 25609763 32788638158600 (SMALED)
BICD2 26998597 29353221 32709491615290 (SMALED2A) 618291 (SMALED2B)
Charcot-Marie-Tooth type 2: distal lower limb weakness, abnormal gait DYNC1H1 24307404 20697106 22459677 22847149 33242470614228 (CMT2O)
DNAH10 26517670Not listed in OMIM
Asphyxiating thoracic dystrophies (including Jeune syndrome): skeletal abnormalities that may include short ribs and a chest wall deformity, shortened arm and leg bones, an unusually shaped pelvis, polydactyly, renal and hepatic disease (more rarely, retinal disease) DYNC2H1 19442771 26874042 27925158 31935347613091 (SRTD3)
DYNC2I1 23910462 26874042 29271569615503 (SRTD8)
DYNC2I2 24183449 24183451615633 (SRTD11)
DYNC2LI1 26130459617088 (SRTD15)
DYNLT2B 25830415 26044572 28475963617405 (SRTD17)
Retinal degeneration DYNC2H1 32753734Not listed in OMIM
Nonsyndromic rod-cone dystrophy DYNC2I2 33124039Not listed in OMIM
Neurodevelopmental disorder with microcephaly and structural brain anomalies DYNC1I2 31079899618492 (NEDMIBA)
Mirror movements type 3: movements on one side of the body are involuntarily mirrored on the other side of the body DNAL4 25098561616059 (MRMV3)
Mental retardation autosomal dominant 13 DYNC1H1 23603762 22368300614563 (MRD13)
Spermatogenic failure DNAH1 24360805 33989052617576 (SPGF18)
DNAH2 30811583619094 (SPGF45)
DNAH8 32619401619095 (SPGF46)
DNAH17 31178125 31658987 31841227618643 (SPGF39)
Lissencephaly: developmental delay, myoclonic jerks and spasms, seizures, hypotonia, microcephaly, dysmorphic facies PAFAH1B1 32692650 20301752 32341547 28886386601545 (LIS)
Seckel syndrome: growth retardation, microcephaly, developmental delay NIN 27053665 22933543614851 (SCKL7)
Open in a separate windowaNote that for some of these phenotypes, there are several variants with varying degrees of severity, and different genes may be associated with different types of these genetic conditions.Mutations in DNAH5 encoding an axonemal ODA heavy chain are the most common genetic defect observed in PCD (Hornef et al., 2006). DNAH5 mutations result in dysmotility of respiratory as well as nodal cilia (Olbrich et al., 2002). Defective nodal cilia motility during early embryogenesis caused by mutations in genes encoding components essential for ciliary motility (e.g., due to DNAH5 mutations) result in situs inversus or situs ambiguus in approximately half of affected individuals due to the randomization of their left–right body asymmetry. Consistently, mice deficient for DNAH5 show immotility of respiratory cilia and embryonic nodal monocilia and exhibit ODA defects in both cilia types (Nöthe-Menchen et al., 2019). DNAH5 mutations also result in ODA defects and dysmotility of ependymal cilia (Ibañez-Tallon et al., 2004). DNAH5-deficient mice develop hydrocephalus during early postnatal life because the flow of cerebrospinal fluid around the brain is obstructed by the abnormal closure of the aqueduct of Sylvii connecting the third and fourth brain ventricles. Possibly due to the larger human brain size, the active propulsion of cerebrospinal fluid along the narrow passages of the ventricular system is not essential in most individuals with PCD; however, they still carry a slightly increased risk of developing hydrocephalus. This suggests that the non–motility-related functions of ependymal cilia might also be important (Wallmeier et al., 2020).All known motile cilia types with DNAH5 loss-of-function mutations display aberrant motility, with the exception of sperm flagella. This is because the paralogous protein DNAH8 is present in sperm and exhibits functional overlap. The male reproductive tracts of mice deficient for DNAH5 have immotile efferent duct cilia, which results in severe stasis of sperm cell transport; this is due to disruption of the ODA composition. In human individuals with loss-of-function DNAH5 mutations, reduced sperm count in the ejaculate (oligozoospermia) and dilatations of the epididymal head were observed, consistent with DNAH5 in efferent duct cilia having an important role in sperm cell transport (Aprea et al., 2021a).In females, the ODA composition of cilia in the Fallopian tube resembles that of respiratory cilia, with the ODA DNAH5 (dynein axonemal heavy chain 5) and DNAI1 (dynein axonemal intermediate chain 1) both being present (Raidt et al., 2015). The coordinated beating of the Fallopian tube ciliated cells produces a fluid flow from the distal site of the Fallopian tubes (ovaries), which transports the egg to the proximal end of the reproductive tract (uterine cavity; Lyons et al., 2006). Interestingly, some females with defective DNAH5 and DNAI1 are still able to conceive children. Thus, the motility of Fallopian tube cilia may not be essential for gamete transport, as Fallopian tube muscle contractions might aid in transporting the egg to the uterine cavity.Mutations in genes encoding DNAAFs cause variable degrees of absence of ODAs and IDAs in respiratory cilia and sperm flagella (Aprea et al., 2021b), indicating that the process of cytoplasmic assembly of dynein arms is critical in both cell types. DNAAF mutant individuals consistently exhibit severely hampered motility of both sperm flagella and respiratory cilia. The sperm flagella of some DNNAF mutant males have shortened flagella axonemes, indicating that their length is also influenced by DNAAF function during dynein arm assembly.Most defects of DNAAFs and axonemal dynein components affect motility of cilia and sperm flagella, contributing to motile ciliopathies (Leigh et al., 2019; Reiter and Leroux, 2017; Wallmeier et al., 2020). However, mutations in genes encoding cytoplasmic dynein subunits can affect the function of both motile and nonmotile cilia, as well other cellular processes. Thus, the clinical phenotype can vary enormously depending on the cell types that are affected. A variant of DYNC1H1 has been associated with a particular form of the ciliopathy SMALED (spinal muscular atrophy lower extremity dominant). This form of the condition mainly affects the lower limbs, causing progressive muscle weakness (Das et al., 2018). A different point mutation in DYNC1H1, also within the tail domain of the heavy chain protein, has been associated with the related neuropathy Charcot Marie Tooth disease. Dysfunction of the dynein heavy chains encoded by DYNC1H1 may also adversely affect maintenance of the morphology of mitochondria and may contribute to disease pathology (Eschbach et al., 2013).Variants of several genes encoding dynein 2 subunits (Dagoneau et al., 2009). If retrograde IFT trafficking of cargoes from the tip to the base of the cilium is compromised, then so is hedgehog (Hh) signaling in the developing embryo, and the resulting incorrect embryonic patterning can produce a range of phenotypes (Goetz and Anderson, 2010). Patients with these conditions have skeletal abnormalities including a narrow thorax, short ribs, and bony spurs in a three-pronged formation observed at the hip joint; they may also display polydactyly.Variants of some of the genes encoding dynein 2 subunits have also been linked to phenotypes affecting vision. The outer segment of photoreceptors is a modified cilium, and a constant turnover of outer segment constituents is required; IFT is key to this process. Four variants in DYNC2H1 in human are associated with nonsyndromic retinal degeneration (Vig et al., 2020). Some of these variants are suggested to affect the ciliary transport of the protein encoded by IFT88, an IFT component that is essential for the assembly and maintenance of vertebrate photoreceptors (Pazour et al., 2002).Standardizing gene nomenclatureThe HGNC (https://www.genenames.org) is the international authority assigning standardized nomenclature to human genes, and hence facilitating communication between researchers. We aim to assign unique, informative symbols and names to human genes that can be used in all domains, and across major biological and clinical databases and publications. Our sister project, the Vertebrate Gene Nomenclature Committee (VGNC; https://vertebrate.genenames.org), names genes across selected vertebrates in line with their human orthologues. VGNC species currently include chimp, macaque, cow, dog, horse, pig, and cat. We also work with other nomenclature committees responsible for naming genes in model vertebrates, such as mouse, rat, and Xenopus, to ensure consistency across species when possible (Tweedie et al., 2021).Every named human gene has a symbol report on the HGNC website listing key data, including the approved nomenclature, published aliases, and locus type. An HGNC symbol report also contains links to multiple relevant sequence databases and clinical resources. It may additionally contain a link to a gene group page (see below), links to VGNC pages for orthologues in selected vertebrate species, and links to key publications in Europe PMC and PubMed. All data including our nomenclature guidelines (Bruford et al., 2020) can be accessed via our website.The green alga C. reinhardtii is a key model organism for studying eukaryotic cilia and flagella and the dynein motor complexes that aid in their assembly and drive their movement. The alveolate Tetrahymena thermophila and sea urchins such as Strongylocentrotus purpuratus are also key model organisms for studying ciliary function. The nomenclature of human dyneins has been largely based on orthology with C. reinhardtii, but also partly based on sea urchin nomenclature. Unfortunately, there are inconsistencies in the naming of orthologues among these species due to historic numbering assignments based on protein migration in SDS/urea-polyacrylamide gels. We have brought mammalian dynein nomenclature more into line with that of C. reinhardtii where possible and have established a naming system for genes encoding dynein chains that are unique to vertebrate species.While the stability of gene symbols, particularly those associated with phenotypes, is now a priority for the HGNC, we are still willing to consider updates for genes approved with placeholder symbols or for genes with domain-based nomenclature that may not give a clue to the function of the encoded protein, for example, genes named based on whether their encoded proteins contain transmembrane domains or coiled-coil regions (CCDC). Symbol changes are made only if an approved symbol has not become entrenched in the literature and if the community working on the gene in question is supportive of change to something more functionally informative.In 2005, the nomenclature for the mammalian cytoplasmic dynein genes was revised (Pfister et al., 2005). The introduction of new DYNC1 and DYNC2 root symbols helped clarify whether genes encoded subunits that were components of the dynein 1 or dynein 2 complex. New root symbols were also introduced to subdivide the known human dynein light chains into three families: roadblock (DYNLRB), Tctex (DYNLT), and LC8 (DYNLL). A 2011 paper (Hom et al., 2011) reported updates made to C. reinhardtii dynein gene nomenclature based on the structural properties of their encoded protein products. This more systematic naming system helped to make the cross-species comparison of orthologues more straightforward and provided a framework for naming newly characterized dynein-encoding genes. Note that there are several human genes encoding dynein chains without orthologues in C. reinhardtii, as it lacks an equivalent of the cytoplasmic dynein 1/dynactin system, so some of the nomenclature is mammal specific.Here we discuss our recent nomenclature updates for genes encoding dynein complex subunits, ODA-DC subunits, and axonemal dynein assembly factors in the human genome (Tweedie et al., 2021), as well as in the model organisms that follow HGNC nomenclature such as mouse, rat, and Xenopus.Table 2.Summary table of nomenclature updates reported here
Approved HGNC SymbolNameAliases (previously approved symbols in bold)Chlamydomonas orthologuea (genes and proteins)Protein present in
DYNLT2 Dynein light chain Tctex-type 2TCTE3, TCTEX1D3, TCTEX2, Tctex4DLT2 (LC2)Axonemal ODA complex
ODAD1 Outer dynein arm docking complex subunit 1CCDC114, FLJ32926, CILD20DCC2 (ODA1) and DCC3 (ODA5)bAxonemal ODA complex
ODAD2 Outer dynein arm docking complex subunit 2ARMC4, FLJ10817, FLJ10376, DKFZP434P1735, CILD23, guduNo orthologueAxonemal ODA complex
ODAD3 Outer dynein arm docking complex subunit 3CCDC151, MGC20983, ODA10DCC1 (ODA3) and ODA10 (ODA10)bAxonemal ODA complex
ODAD4 Outer dynein arm docking complex subunit 4TTC25, DKFZP434H0115No orthologueAxonemal ODA complex
DNAI3 Dynein axonemal intermediate chain 3WDR63, DIC3, FLJ30067, NYD-SP29DIC3 (IC140)Axonemal IDA I1/f complex
DNAI4 Dynein axonemal intermediate chain 4WDR78, DIC4, FLJ23129DIC4 (IC138)Axonemal IDA I1/f complex
DNAI7 Dynein axonemal intermediate chain 7CFAP94, CASC1, LAS1, FLJ10921, PPP1R54, IC97DII6 (FAP94)Axonemal IDA I1/f complex
DYNLT2B Dynein light chain Tctex-type 2BTCTEX1D2, MGC33212DLT4 (Tctex2b)Axonemal IDA I1/f complex
Cytoplasmic dynein 2 complex
DYNC2I1 Dynein 2 intermediate chain 1WDR60, FLJ10300, FAP163, CFAP163, DIC6DIC6 (FAP163)Cytoplasmic dynein 2 complex
DYNC2I2 Dynein 2 intermediate chain 2WDR34, DIC5, MGC20486, bA216B9.3, FAP133, CFAP133DIC5 (FAP133)Cytoplasmic dynein 2 complex
DYNLT3 Dynein light chain Tctex-type 3TCTE1L, TCTEX1LDLT1 (LC9)Cytoplasmic dynein 2 complex
DNAAF8 Dynein axonemal assembly factor 8C16orf71, FLJ43261, DKFZp686H2240Axonemal dynein assembly factor
DNAAF9 Dynein axonemal assembly factor 9C20orf194, DKFZp434N061 DNAAF9 Axonemal dynein assembly factor
DNAAF10 Dynein axonemal assembly factor 10WDR92, FLJ31741, Monad DNAAF10 Axonemal dynein assembly factor
DNAAF11 Dynein axonemal assembly factor 11LRRC6, TSLRP, LRTP, CILD19, tilBDNAAF11, MOT47, LRRC6, SeahorseAxonemal dynein assembly factor
LRRC56 Leucine rich repeat containing 56DNAAF12, FLJ00101, DKFZp761L1518DLU2 (ODA8)Axonemal dynein assembly factor
SPAG1 Sperm associated antigen 1DNAAF13, SP75, FLJ32920, HSD-3.8, TPIS, CT140, CILD28,SPAG1 (SPAG1)Axonemal dynein assembly factor
PIH1D1 PIH1 domain containing 1DNAAF14, FLJ20643, Pih1, MOT48,DAP2 (MOT48)Axonemal dynein assembly factor
PIH1D2 PIH1 domain containing 2DNAAF15Axonemal dynein assembly factor
CFAP298 Cilia and flagella associated protein 298FLJ20467, DAB2, FBB18, CILD26, Kur, C21orf48, C21orf59, DNAAF16 DAB2 Axonemal dynein assembly factor
CCDC103 Coiled-coil domain containing 103FLJ13094, FLJ34211, PR46b, CILD17, DNAAF17c CCDC103 Axonemal dynein assembly factor
DAW1 Dynein assembly factor with WD repeats 1FLJ25955, ODA16, WDR69, DNAAF18 DAW1 Axonemal dynein assembly factor
Open in a separate windowaInformation about C. reinhardtii ciliary proteins, including dynein components, is curated and available at http://chlamyfp.org/.bChlamydomonas encodes two paralogous proteins that both have the same human orthologue.cReserved symbol/alias symbol. This gene will either be updated as a DNAAF or a DNAAF symbol will be added as an alias if further future publications support this.Gene groupsHGNC gene groups are manually curated using data from publications and advice from our specialist advisors. The groups for genes encoding the subunits of human dynein complexes can be viewed here: https://www.genenames.org/data/genegroup/#!/group/537 and reflect the data shown in Table S1, Table S2, Table S3, Table S4, Table S5, and Table S6.Discussion of HGNC nomenclature updates for dyneins and their cytoplasmic assembly factors

Dynein light chain nomenclature updates (dynein light chain Tctex-type [DYNLT])

Based on advice from experts in the field, we have updated the nomenclature of all the Tctex family genes to better reflect the function of their encoded proteins as dynein subunits. The six paralogs in this set now use the root symbol DYNLT in human.

DYNLT1 and DYNLT3

The gene currently approved as DYNLT1 (HGNC ID: 11697) was first approved using the symbol TCTEL1 based on homology with the mouse gene Tcte1 (t-complex associated testis expressed 1; Watanabe et al., 1996), which was reported to be specifically expressed in murine testes (Lader et al., 1989; Sarvetnick et al., 1989). The t-complex is a region of the mouse genome that shows non-Mendelian segregation, and some of the genes in it are associated with spermatogenesis (Castaneda et al., 2020). The alias symbol Tctex1 was also used to publish on this gene; it was characterized as encoding a cytoplasmic dynein light chain (Dedesma et al., 2006; King et al., 1998) and later also identified in axonemal inner arm I1/f (Harrison et al., 1998); in C. reinhardtii, a closely related protein is present in the ODA (DiBella et al., 2005).The most closely related paralogous gene to DYNLT1, now approved as DYNLT3 (HGNC ID: 11694), was originally assigned the symbol TCTE1L (Tcte1-like) in human, again to reflect its homology to mouse Tcte1. It was also published as a candidate for the retinitis pigmentosa RP3 locus (Roux et al., 1994), although this link was later disproven (Meindl et al., 1996) when RPGR (retinitis pigmentosa GTPase regulator) was identified as the causative gene for this phenotype (Ferrari et al., 2011). DYNLT3 was reported to encode a cytoplasmic dynein light chain in 1998 (King et al., 1998) and was later published as also playing a role in regulating primary cilium length (Palmer et al., 2011). We have constructed a phylogenetic tree (Fig. 4) that shows there is no clear 1:1 orthology relationship for either human DYNLT1 or DYNLT3 with respect to invertebrate species.Open in a separate windowFigure 4.Maximum-likelihood phylogenetic tree to show the relationship of Tctex-type dynein light chains in selected species. This tree is shown with a midpoint rooting. The figures on the nodes show the Shimodaira–Hasegawa likelihood ratio test and the Ultrafast bootstrap support values for the branches (SH-aLRT %/UFBoot %). Bootstrap values of ≥70% only are shown. The scale bar represents the expected number of amino acid substitutions per site. M. musculus has multiple Dynlt1 and Dynlt2 paralogs, but as these are identical at the amino acid level, only one sequence has been included in each case. The colors highlight supported clades: green for DYNLT1 and DYNLT3 and their orthologues, blue for DYNLT2 and its orthologues, red for DYNLT2B and its orthologues, yellow for DYNLT4 and its orthologue, and purple for DYNLT5 and its orthologues.

DYNLT2 and DYNLT2B

We have updated the nomenclature of the gene previously approved as TCTE3 (HGNC ID: 11695) to DYNLT2, and that of its closely related paralog previously approved as TCTEX1D2 (Tctex1 domain containing 2; HGNC ID: 28482) to DYNLT2B. These new symbols are more functionally informative, and this update brings the human nomenclature into line with that of C. reinhardtii, S. purpuratus, and T. thermophila (see Fig. 4). The phylogeny (Fig. 4) shows the paralogous relationship between DYNLT2 and DYNLT2B and that their 1:1 orthologues in the other species fall into two separate subclades.Although DYNLT2 and DYNLT2B are paralogs, their protein products are components of distinct dynein complexes. DYNLT2 encodes an axonemal dynein subunit, required for outer arm assembly (Patel-King et al., 1997), and has not been reported as being part of any cytoplasmic dynein complex. The DYNLT2B-encoded protein is part of the cytoplasmic dynein 2 complex (Hamada et al., 2018; Schmidts et al., 2015) and is also an axonemal inner arm I1/f complex subunit (DiBella et al., 2004).

DYNLT4 and DYNLT5

We have updated the nomenclature of the gene previously approved as TCTEX1D4 (HGNC ID: 32315) to DYNLT4. This gene encodes a dynein light chain protein that belongs to the TCTEX1 family. Freitas et al. (2014) discussed its role in sperm motility and IFT.While discussing the update for TCTEX1D4 with experts, we also proposed a nomenclature update for TCTEX1D1 (HGNC ID: 26882). This gene could not be updated to DYNLT1 in line with the TCTEX1D1 numbering, as this symbol was already in use, so we proposed an update to DYNLT5. There is currently a single paper published on this human gene (Spitali et al., 2020), linking a variant of it with the phenotype Duchenne muscular dystrophy. Although it seems likely that, as a paralog of the other DYNLT genes, DYNLT5 will be found to encode a dynein light chain, we have included the term family member in its current gene name to indicate that although it is related to the other DYNLT genes, a shared function has not yet been established. The phylogeny (Fig. 4) reveals that S. purpuratus has a 1:1 orthologue of DYNLT5, while C. reinhardtii and T. thermophila do not.DNAI nomenclature updates

DNAI3 and DNAI4

We have updated the nomenclature of the human orthologues of C. reinhardtii DIC3, encoding IC140 (alias IDA7); and DIC4, encoding IC138 (alias BOP5), to DNAI3 (HGNC ID: 30711) and DNAI4 (HGNC ID: 26252), respectively. These genes were previously approved as WDR63 (WD repeat domain 63) and WDR78. In C. reinhardtii, IC140 and IC138 have been well characterized as intermediate chain subunits of an IDA complex (I1 dynein complex, also known as dynein-f; Hendrickson et al., 2004; Yang and Sale, 1998). Updating WDR63 and WDR78 using the DNAI root brings their nomenclature in line with the other human genes encoding axonemal dynein intermediate chains, DNAI1 and DNAI2. It also keeps the numbering system used equivalent to that of the C. reinhardtii orthologues.The DNAI3-encoded protein is not essential for fertility in male mice, as other intermediate chains of the IDA I1/f complex may compensate for this role in mouse sperm motility (Young et al., 2015). The mouse orthologue of DNAI4 encodes a dynein intermediate chain in vertebrates. The DNAI4 protein interacts with multiple subunits of the axonemal inner arm I1/f dynein complex and is essential for the ciliary assembly of this complex in vertebrates (Zhang et al., 2019).

DNAI7 and NME8 (alias DNAI8)

We originally considered updating the nomenclature of the gene previously approved as CASC1 (cancer susceptibility 1; HGNC ID: 48939) to DNAI5. However, after discussion with experts, we realized this could be confusing, as it is not the orthologue of C. reinhardtii DIC5, and all the other human DNAI genes are numbered in line with their C. reinhardtii orthologues. There is also a DIC6 gene in C. reinhardtii, and its orthologue is the human gene now approved as DYNC2I1 (dynein 2 intermediate chain 1).We were also reluctant to reassign CASC1 as DII6, the symbol used for the C. reinhardtii orthologue of this gene (Hom et al., 2011). We do not have an established DII# (dynein inner arm interacting) root approved in human, and most of the orthologues of the DII# C. reinhardtii genes are already approved and published using alternative symbols. These genes include DNALI1 (dynein axonemal light intermediate chain 1), the orthologue of DII1; ACTG1 (actin γ1), the orthologue of DII4; and ANK2 (ankyrin 2), the orthologue of DII7. In addition, with the exception of DNALI1, it is possible that one or more of these genes may not necessarily encode proteins that are dynein-arm interacting in vertebrates. Therefore, we updated CASC1 as DNAI7, reflecting that its protein product is a dynein intermediate chain in human. The mouse orthologue of DNAI7 encodes an intermediate chain in vertebrates that forms part of the inner arm I1/f dynein complex required for ciliary beating (Zhang et al., 2019).This leaves NME8 (NME/NM23 family member 8) as the only remaining human gene known to encode a dynein intermediate chain but not named as such. This gene was previously approved as TXNDC3 (thioredoxin domain containing 3; Duriez et al., 2007) and has also been published using the alias symbol SPTRX2 (sperm-specific thioredoxin 2; Sadek et al., 2001).There are 10 genes in the human NME/NM23 family, at least five of which encode active nucleoside diphosphate kinases (Ćetković et al., 2015). NME8 (HGNC ID: 16473) is the human orthologue of the sea urchin IC1 gene (Duriez et al., 2007), which encodes a sea urchin ODA intermediate chain and, like its human orthologue, contains an N-terminal thioredoxin-like domain (Ogawa et al., 1996). In C. reinhardtii, the ODA contains two paralogous thioredoxin-like light chains (LC3 and LC5) but lacks a nucleoside diphosphate kinase (Patel-King et al., 1996).NME8 encodes a protein with a ciliary role, and its gene product is suggested to be bifunctional, with isoforms expressed at varying levels in different tissues (Duriez et al., 2007). The TXNDC3d7 protein isoform can bind microtubules, plays a role in ciliary function, and may be a component of ODAs (Duriez et al., 2007). As NME8 is already named as part of a gene group, is a functionally informative symbol, and has been used in the literature, we have decided to retain this nomenclature. However, this gene has been assigned the alias symbol DNAI8 and added to our dynein axonemal outer arm complex subunits gene group page (https://www.genenames.org/data/genegroup/#!/group/2031).

DYNC2I1 and DYNC2I2

We have updated the nomenclature of the human orthologue of C. reinhardtii DIC6 encoding D1bIC1 (alias FAP163) from WDR60 to DYNC2I1 (HGNC ID: 28296). We have also updated the nomenclature of the human orthologue of C. reinhardtii DIC5, encoding D1bIC2 (alias FAP133) from WDR34 to DYNC2I2 (HGNC ID: 21862). The numbering was assigned in this way so that the human gene nomenclature corresponds to that of the C. reinhardtii proteins.DIC5/FAP133 in C. reinhardtii is associated with the IFT dynein motor (dynein 2, usually known as dynein 1b in C. reinhardtii) complex (Rompolas et al., 2007). DIC6/FAP163 encodes a C. reinhardtii intermediate chain that is closely related to DIC5/FAP133 and is also a component of the dynein 2 complex (Patel-King et al., 2013). Previous studies linked these two genes to ciliopathies including short rib polydactyly and Jeune syndrome (McInerney-Leo et al., 2013; Schmidts et al., 2013) and suggested that these orthologues of C. reinhardtii dynein intermediate chains may also encode components of the dynein 2 complex. Indeed, it was confirmed that both human genes encode dynein 2 intermediate chains (Asante et al., 2014).ODA-DC (ODAD) nomenclature updates

ODAD1, ODAD2, ODAD3, ODAD4, and CLXN (ODAD5)

The ODA-DC has only recently been characterized in human (Hjeij et al., 2014; Onoufriadis et al., 2013; Wallmeier et al., 2016), and it became apparent that the nomenclature of the genes encoding the constituent proteins was not as functionally informative as it could be. The nomenclature of four of the ODA-DC subunits was initially based on the presence of structural domains in the encoded proteins: ARMC4 (armadillo repeat containing 4), CCDC114 and CCDC151 (coiled-coil domain containing 114 and 151, respectively), and TTC25 (tetratricopeptide repeat domain 25), as there was no functional information published when they were initially named.These four genes have now been reassigned using the root symbol ODAD (ODA-DC subunits). The ODAD genes have been assigned numbers in the order in which they were characterized as encoding ODA-DC subunits in human and in line with the ODA numbering in C. reinhardtii where possible. We could not use the DCC root in human for these genes, as it clashed with the approved symbol for an unrelated gene, DCC (DCC netrin 1 receptor; HGNC ID: 2701).ODAD1 is the orthologue of C. reinhardtii DCC2 (encoding DC2, alias ODA1), which encodes a docking complex subunit, and of its paralog DCC3, which encodes the ODA5 assembly factor (Takada et al., 2002). ODAD3 is the orthologue of DCC1, which encodes the protein DC1 (alias ODA3; Koutoulis et al., 1997), a docking complex subunit in C. reinhardtii, and of its paralog ODA10, which encodes a dynein assembly factor in C. reinhardtii (Dean and Mitchell, 2013). ODAD2 and ODAD4 have no known C. reinhardtii orthologues.A fifth gene has recently been published in a study examining mammalian tracheal cilia as encoding an ODA-DC subunit (Gui et al., 2021; Fig. 3, c and d). Its encoded protein, calaxin, is a member of a neuronal calcium sensor family and was originally identified in ODAs from the sea squirt Ciona intestinalis; subsequent studies revealed it is required for normal ciliary motility in mice (Mizuno et al., 2009; Mizuno et al., 2012; Sasaki et al., 2019). We have updated its approved nomenclature from the previously approved but less frequently used EFCAB1 (EF-hand calcium binding domain 1) to CLXN (calaxin), aliasing it as ODAD5 after discussion with authors.The symbol ODAD6 is reserved for the gene currently approved as CCDC63, a closely related paralog of ODAD1. We will continue to monitor the literature and may update the nomenclature of this gene, either approving ODAD6 or adding it as an alias if CCDC63 is shown to encode an ODA-DC subunit. The ODA-DC gene group page can be seen on our website (https://www.genenames.org/data/genegroup/#!/group/2019).DNAAFsWe have updated the nomenclature of four genes as DNAAFs, including two previously assigned using placeholder C#orf# symbols (see Table S6). There are now 18 genes included in our axonemal dynein assembly factor gene group set (https://www.genenames.org/data/genegroup/#!/group/1627).We have updated the nomenclature of the gene previously approved using the placeholder symbol C16orf71 (chromosome 16 open reading frame 71; HGNC ID: 25081) to DNAAF8. The Xenopus orthologue was recently published using the alias symbol Daap1 (dynein axonemal-associated protein 1; Lee et al., 2020), but following discussion, this gene has been approved as dnaaf8 in line with its human orthologue.We have also updated the nomenclature of the gene previously approved as C20orf194 (chromosome 20 open reading frame 194; HGNC ID: 17721) to DNAAF9. The Tetrahymena orthologue of this gene was published using the alias name “shulin” (Mali et al., 2021). Those authors’ work showed that the encoded protein has a role in keeping the axonemal ODAs in a nonfunctional state before delivery to cilia. With these authors, our experts, and all researchers who had previously published on this gene, we discussed assigning this gene as DNAAF9, and they were supportive of this update. A gene (Cre11.g467556) exhibiting some similarity to DNAAF9 is present in C. reinhardtii; this is in a potentially poorly assembled genomic region, and further characterization will be required to determine whether it is the true orthologue of this human gene.Two other genes, previously approved as WDR92 and LRRC6 (leucine rich repeat containing 6), have also been updated to DNAAF10 and DNAAF11, respectively. Both have been shown to encode proteins that are involved in axonemal dynein assembly (Patel-King and King, 2016; Fabczak and Osinka, 2019; Liu et al., 2019; Patel-King et al., 2019; Li et al., 2021; Zur Lage et al., 2018). The DNAAF10 protein product interacts with the protein encoded by SPAG1 (sperm associated antigen 1; see below) during dynein preassembly (Zur Lage et al., 2018). The DNAAF11 protein product interacts with the protein encoded by ZMYND10 (zinc finger MYND-type containing 10), which is aliased as DNAAF7 (Zariwala et al., 2013). ZMYND10 has been retained as the approved symbol because it has been well used in publications, and the current nomenclature reflects the fact that the encoded protein contains a MYND-type zinc finger domain.We also assigned four other genes (LRRC56, SPAG1, PIH1D1, and PIH1D2) with DNAAF aliases to reflect the roles of their encoded proteins in dynein assembly (Bonnefoy et al., 2018; Knowles et al., 2013; Yamaguchi et al., 2018). These were assigned the alias symbols DNAAF12, DNAAF13, DNAAF14, and DNAAF15, respectively. Although it seems very likely based on two publications (Bonnefoy et al. [2018] and Desai et al. [2015]) that LRRC56 encodes a DNAAF, we are continuing to monitor the literature and could consider updating the nomenclature of this gene to DNAAF12 if there is sufficient evidence published to support this.The SPAG1 and PIH1D1 symbols are already well established in the literature, and SPAG1, PIH1D1, and PIH1D2 all encode proteins that are subunits of complexes with many other functions as well as being involved in dynein assembly (von Morgen et al., 2015). The PIH1D2- and SPAG1-encoded proteins are part of the R2SP complex (Chagot et al., 2019), and the PIH1D1-encoded protein is part of the R2TP complex (Rodríguez and Llorca, 2020). Therefore, we have chosen to retain their currently approved symbols but have added them to our DNAAF gene group page. While we always ask that authors reference the approved gene symbols at least once in all publications, they can of course also use the DNAAF aliases.We also discussed a DNAAF symbol update for the orthologue of C. reinhardtii DAB2 with authors and our expert advisors. DAB2 accumulates in cilia, and their motility is impaired (Austin-Tse et al., 2013). Variants of the Danio rerio orthologue of this gene, Kurly, are found in zebrafish mutants that display abnormalities in their development and have dynein arm defects, suggesting that the Kurly protein plays a role in ciliary motility but is also involved in regulating planar cell polarity (Jaffe et al., 2016). The human orthologue, previously approved as C21orf59, encodes a protein that has been shown to interact with known DNAAFs, including proteins encoded by ZYMND10 and DNAAF11 (previously LRRC6; Cho et al., 2018), and has been associated with the human phenotype PCD (Bolkier et al., 2021). Discussion with authors and our specialist advisors for the DNAAFs and cilia- and flagella-associated proteins (CFAPs) revealed community support for assigning a more general CFAP symbol for this gene. Its association with cilia and flagella is clear, and it also has a wider function beyond its role as an axonemal dynein arm assembly factor. However, while we have updated this gene as CFAP298 (HGNC ID: 1301), we have also assigned it the alias symbol DNAAF16. We also updated another cilia-associated gene, the orthologue of C. reinhardtii FBB5, as CFAP300 (previously approved as C11orf70) and have assigned it the alias symbol DNAAF17. Phylogenetic analysis strongly suggests that this gene is specific to organisms with motile cilia (being part of the MotileCut grouping; Merchant et al., 2007), and our CFAP nomenclature specialist advisor supported this change. As more becomes known about the function of the CFAP300 protein, we can consider whether a further symbol change would be helpful for this gene.We are retaining the symbol DAW1 (dynein assembly factor with WD repeats 1), as it is the orthologue of C. reinhardtii DAW1 and its current nomenclature is functionally informative. However, we have aliased it as DNAAF18 and added it to the DNAAF gene group. We have also reserved the gene symbol DNAAF19 for the gene currently approved as CCDC103. The CCDC103 protein affects dynein assembly (King and Patel-King, 2020; Panizzi et al., 2012), but its exact role has still to be defined.ConclusionIn total, we have updated the nomenclature of nine genes encoding human dynein chains, four genes encoding proteins that form the ODA-DC, and four genes encoding axonemal dynein assembly factors. Several other genes have retained their current symbols but have been aliased as ODADs or DNAAFs and added to the appropriate HGNC gene group pages. All updates were made following consultation with experts from the community, and these changes were widely supported among the authors publishing in this field. While we aim to limit changes in gene nomenclature, especially when the genes are linked to a phenotype, these updates have largely replaced uninformative placeholder or domain-based symbols, and we view the new informative symbols as stable. As such, users should regard these new symbols as the permanent gene symbols for these human genes.We hope that all researchers will use the new nomenclature in their future publications to aid communication and data retrieval within the field. Approved symbols should be mentioned at least once in publications, along with the associated HGNC ID if possible.Materials and methodsDynein light chain phylogenetic treeAmino acid protein sequences for dynein light chains were obtained for each of the six selected species from NCBI. A multiple alignment was built using the MUSCLE online tool (https://www.ebi.ac.uk/Tools/msa/muscle/; Madeira et al., 2019) and edited using AliView 1.20 (Larsson, 2014). The ends of the alignment were trimmed, and all indels were removed. The IQ-TREE web server (http://iqtree.cibiv.univie.ac.at/) was used to construct a maximum-likelihood tree. The substitution model was autoselected with ultrafast bootstrapping and SH-aLRT branch test methods applied.Online supplemental materialThe supplementary tables show HGNC approved nomenclature for genes encoding subunits of dynein complexes alongside their known published alias symbols and their orthologs in C. reinhardtii. Table S1 shows cytoplasmic dynein 1 subunits. Table S2 shows cytoplasmic dynein 2 subunits. Table S3 shows axonemal ODA subunits. Table S4 shows monomeric dynein heavy chains and their accessory subunits. Table S5 shows axonemal inner arm dynein I1/f subunits. Table S6 shows axonemal dynein assembly factors (DNAAFs).  相似文献   

18.
Fertilization is defined as the union of two gametes. During fertilization, sperm and egg fuse to form a diploid zygote to initiate prenatal development. In mammals, fertilization involves multiple ordered steps, including the acrosome reaction, zona pellucida penetration, sperm–egg attachment, and membrane fusion. Given the success of in vitro fertilization, one would think that the mechanisms of fertilization are understood; however, the precise details for many of the steps in fertilization remain a mystery. Recent studies using genetic knockout mouse models and structural biology are providing valuable insight into the molecular basis of sperm–egg attachment and fusion. Here, we review the cell biology of fertilization, specifically summarizing data from recent structural and functional studies that provide insights into the interactions involved in human gamete attachment and fusion.

IntroductionDuring sexual reproduction, the oocyte and sperm fuse to generate a new and unique embryo. The journey of a sperm to an egg ends in the ampulla of the female oviduct. From there, the sperm must overcome a number of physical and biochemical barriers. After undergoing the acrosome reaction and binding the ova, the sperm penetrates through the cumulus oophorus cells and the zona pellucida (ZP) to reach the perivitelline space (PVS) and oocyte membrane. Upon fusion of the sperm and egg membranes, the sperm nucleus and organelles are incorporated into the egg cytoplasm.An understanding of the mechanisms of mammalian fertilization is crucial to treat infertility and develop new methods of birth control. Infertility affects 15% of couples globally, and in one third of these cases, the underlying cause is unknown (Gelbaya et al., 2014). Developments in assisted reproductive technologies have provided couples with new options to conceive but may have epigenetic side effects (Mani et al., 2020). Furthermore, only 40% of couples manage to have a child despite 2 yr of treatment. Safety, efficacy, and acceptability of contraceptives are also critically important, but many current female contraceptive methods have side effects that limit long-term use (Aitken et al., 2008), while male contraceptives are limited to condoms or vasectomy (Kanakis and Goulis, 2015). A better understanding of the molecular players involved in fertilization is necessary to drive innovation in both assisted reproductive technologies and contraception.In this review, we will first briefly review the events that prepare the gametes for fertilization. We will then discuss how recent studies of genetically altered mice and structural biology efforts have shed light on the molecular mechanisms of sperm–egg attachment and fusion. We will also discuss the gaps in current knowledge and suggest new perspectives and future directions in the search for other protein factors involved at the gamete fusion synapse.Cell biology of gametesFertilization requires proper gametogenesis (oogenesis in the female and spermatogenesis in the male), which produces haploid cells and introduces diversity. Primordial germ cells (PGC) are the embryonic precursors to spermatocytes and ova. The cells produced by the first few divisions of the fertilized egg are totipotent and capable of differentiating into any cell type, including germ cells. PGCs originate within the primary ectoderm of the embryo and then migrate into the yolk sac. Between weeks 4 and 6, the PGCs migrate back into the posterior body wall of the embryo, where they stimulate cells of the adjacent coelomic epithelium and mesonephros to form primitive sex cords and induce the formation of the genital ridges and gonads. The sex (gonadal) cords surround the PGCs and give rise to the tissue that will nourish and regulate the development of the maturing sex cells (ovarian follicles in the female and Sertoli cells in the male).EggOogenesis is a complex differentiation process by which mature functional ova develop from germ cells (Fig. 1 A; Edson et al., 2009). In humans, oogenesis begins in the ovary at 6–8 wk of fetus development, when PGCs differentiate into oogonia. By the 12th week, several million oogonia enter prophase, the first meiotic division and become dormant until shortly before ovulation (Hayashi et al., 2020). Due to their large and watery nuclei, these cells are referred to as germinal vesicles (Pan and Li, 2019). These primary oocytes become enclosed by follicle cells to form primordial follicles. The number of primordial follicles peaks at ∼7 million by the fifth month of fetal life, with ∼700,000 left at birth and 400,000 by puberty (Marcozzi et al., 2018). All of the egg cells that the ovaries will release are already present at birth.Open in a separate windowFigure 1.Gametogenesis and fertilization. (A–C) Illustration of oogenesis and follicle development (A), spermatogenesis (B), and the major steps in fertilization (C): (1) initial contact, (2) acrosome reaction, (3) ZP penetration, (4) sperm–egg fusion, (5) entry of sperm nucleus, (6) cortical reaction, and (7) fusion of the sperm and egg nuclei. The oocyte with its ZP measures 130 μm in diameter. Created with BioRender.During each menstrual cycle, hormones from the hypothalamic–pituitary–gonadal axis restart the division of the primary oocytes in meiosis I and follicular development (Atwood and Vadakkadath Meethal, 2016). Primary follicles develop into secondary follicles, containing each growing oocyte surrounded by two or more layers of proliferating follicle cells. ZP glycoproteins are secreted by the oocyte of the primary follicle and possibly the follicular cells (Törmälä et al., 2008). Although these glycoproteins form a physical barrier between the follicle cells and the oocyte, follicle cells and the oocyte remain connected through transzonal cytoplasmic projections from the follicle cells until fertilization (Makabe et al., 2006). A reciprocal dialog between the oocyte and its surrounding follicular cells coordinates the different phases of follicular development and the maintenance of meiotic arrest (Dalbies-Tran et al., 2020). Oocyte-derived microvilli control female fertility by optimizing ovarian follicle selection in mice (Zhang et al., 2021). The epithelium of 5–12 primary follicles proliferates to form a multilayered capsule around the oocyte. A few of these growing follicles continue to enlarge in response to follicle-stimulating hormone (FSH; Visser and Themmen, 2014). A single follicle becomes dominant, and the others degenerate by atresia (Atwood and Vadakkadath Meethal, 2016). Meiosis of the oocyte in the mature preovulatory follicle is blocked until a surge in levels of FSH and luteinizing hormone that occurs midway through the menstrual cycle. The membrane of the germinal vesicle nucleus breaks down, the chromosomes align in metaphase, and the oocyte expels its first polar body. The secondary oocyte then begins the second meiotic division, which is arrested at the meiotic metaphase II stage until ovulation (Gougeon, 1996). Ovulation depends on the breakdown of the follicle wall and occurs ∼38 h after the increase in levels of FSH and luteinizing hormone (Holesh et al., 2021). The disruption of the follicle wall expulses the oocyte, which is captured by the fimbriated mouth of the oviduct and moved into the ampulla. The oocyte retains its ability to be fertilized for ∼24 h and completes meiosis only if it is fertilized.SpermIn contrast to oogenesis, which is complete before birth, spermatogenesis is a continuous process that begins at puberty (Fig. 1 B). In humans, spermatogenesis takes 74 d to complete; thus, multiple spermatogenesis events occur simultaneously to allow for continual sperm production. Spermatogenesis occurs in the testis in a stepwise manner, beginning with diploid spermatogonia at the basal surface of seminiferous tubules and ending with mature elongated spermatozoa that are released in tubule lumens in a process called spermiation (Clermont, 1972; Yang and Oatley, 2014). During spermatogenesis, mitosis results in gene amplification, meiosis results in genome reduction, and finally maturation occurs (Hess and Renato de Franca, 2008). At this stage, sperm are not motile and are fertilization incompetent. Two additional sperm maturational processes are required outside the testis. First, sperm undergo a maturation process during epididymal transit (Bedford et al., 1973) involving posttranslational modifications of previously synthesized proteins and acquisition of proteins from the epididymal epithelium (James et al., 2020; see text box). After ejaculation into the female reproductive tract, dilution triggers additional changes in sperm, collectively termed capacitation (see text box), that prepare the sperm for the acrosome reaction.Epididymal maturationSperm exchange with the epididymal epithelium occurs by direct interaction with epithelial cells, by interaction with soluble proteins in the epididymal fluid or via extracellular exosome-like vesicles released by epithelial cells called epididymosomes (James et al., 2020). The purposes of this exchange are to redistribute sperm proteins and change the composition and lipid balance of the sperm membrane. These changes take place during the transit from the epididymis initial segment, through the caput and the corpus, to the cauda where sperm are stored (Cornwall, 2009). Epididymal transit lasts 10–12 d in mammals, but storage is dependent on sexual activity. Since fertilization is not immediate, fertilizing capacities of the spermatozoa are preserved by decapacitation factors that are active in the epididymis. An example of a decapacitation factor is SPINK3, which is secreted by seminal vesicles; it impairs sperm membrane hyperpolarization and calcium influx through CatSper (Zalazar et al., 2020). Epididymal plasma and sperm represent only a small fraction (5%) of semen in men (Batruch et al., 2011). Two thirds of the volume of semen comes from the seminal vesicles and the other third from the prostate. These secretions protect the sperm and prevent early maturation.Sperm capacitationMore than 70 yr ago, Austin and Chang described capacitation as the changes required for sperm to fertilize oocytes in vivo (Austin, 1952; Chang, 1951). Once sperm enter the female reproductive tract, they undergo capacitation. Capacitation results in hyperactivation of sperm movement and initiation of the acrosome reaction (Saling et al., 1979; Florman and First, 1988). During capacitation, stabilizing or decapacitation factors that are adsorbed on the sperm plasma membrane are removed (Bedford and Chang, 1962). These agents that initiate removal of decapacitation factors are electrolytes, energy substrates, and proteins such as seminal plasma protein or albumin. Removal of decapacitation factors increases sperm plasma membrane fluidity, allowing an increase in the permeability to calcium, chloride, and bicarbonate ions (Gangwar and Atreja, 2015). Sperm motility depends on the membrane potential, intracellular pH, and balance of intracellular ions (reviewed in Nowicka-Bauer and Szymczak-Cendlak, 2021). The most important ion for this function is Ca2+ (Hwang et al., 2019). This secondary messenger is an important signaling pathways activator that regulates sperm motility (Finkelstein et al., 2020). The activation of soluble adenyl cyclases generates cyclic adenosine monophosphate that in turn activates serine/threonine protein kinase A, which induces a cascade of protein phosphorylation initiating the induction of sperm motility (Chen et al., 2000). Protein phosphorylation, sperm hyperactivation, and the acrosome reaction are used in vitro to evaluate capacitation. Capacitation can be induced in vitro by incubation in medium containing calcium, bicarbonate ions, and serum albumin (Touré, 2019).Mammalian sperm capacitation occurs during sperm migration in the female tract. Mammalian males ejaculate millions of sperm cells into the female reproductive tract, but only a few hundred sperm at most reach the oocytes. This massive elimination process likely prevents polyspermy (reviewed in Kölle, 2015). Selection of human sperm during the journey begins in the acidic environment of the vagina. In the cervix, only morphologically normal sperm can migrate. Some sperm immediately pass into the cervical mucus, whereas the remaining sperm becomes a part of the coagulum. The next selection occurs at the uterus–tubal junction, the connection between the uterus and the oviduct that represents a major obstacle for sperm migration (Kölle, 2015). Experiments in mice indicate that sperm motility alone is insufficient for sperm migration through the uterus–tubal junction (Fujihara et al., 2018). Uterine contractions facilitate sperm transport as do molecular interactions. Several proteins, such as ADAM3 and other ADAM family members, are known to be involved in this step in mice (Yamaguchi et al., 2009; Xiong et al., 2019); most ADAM proteins have human orthologues.Spermatogenesis takes place in a species-specific cycle called the seminiferous epithelial cycle and is regulated in particular through the hypothalamic–pituitary–testicular axis. Indeed, at puberty, the testes (interstitial steroidogenic Leydig cells) secrete an increased amount of testosterone, which triggers growth of the testes, maturation of the seminiferous tubules, and the commencement of spermatogenesis. The Sertoli cells are the major somatic cells present in the seminiferous tubules and are considered to be the main regulators of spermatogenesis. They orchestrate spermatogenesis by supporting spermatogonial stem cells, determining the testis size, organizing meiotic and postmeiotic development and sperm output, supporting androgen production by maintaining the development and function of Leydig cells, and regulating other aspects of testis function like peritubular myoid cells, immune cells, and the vasculature, which participate in the maintenance of the spermatogonial stem cell niche.Acrosome reactionThe acrosome is a secretory vesicle located on the anterior region of sperm that originates from the spermatid Golgi apparatus. An acrosomal granule is formed by the fusion of proacrosomal vesicles in the vicinity of the nucleus. The region increases in size and spreads over the anterior part of the nucleus. The acrosome reaction is driven by SNARE complexes and results in the exocytosis of the contents of the acrosome upon fusion of the plasma membrane with the outer acrosomal membrane (Fig. 1 C; reviewed in Okabe, 2016; De Blas et al., 2005). The timing of the acrosome reaction is critical. Only sperm that have undergone this reaction are fertilization competent, but when a high proportion of sperm undergo the acrosome reaction prematurely, success of in vitro fertilization is low (Wiser et al., 2014). Several studies indicate that only a fraction of sperm is capable of undergoing spontaneous acrosome reaction. In human and mice sperm samples, 15–20% of cells undergo spontaneous acrosome reaction (Nakanishi et al., 2001), whereas only 20–30% undergo progesterone-induced acrosome reaction (Stival et al., 2016), suggesting physiological heterogeneity of sperm population. In addition, Inoue et al. demonstrated that acrosome-reacted mouse spermatozoa recovered from the PVS can fertilize other eggs (Inoue et al., 2011).Based on in vitro data, it was thought that the acrosome reaction occurs when the sperm contacts the ZP, particularly the ZP3 protein (Litscher and Wassarman, 1996). Using transgenic mice that express fluorescent markers in the acrosome (Nakanishi et al., 1999) and the midpiece mitochondria (Hasuwa et al., 2010), real-time observation of acrosomal exocytosis was possible. These experiments showed that most mouse spermatozoa capable of fertilization had undergone the acrosome reaction before contact with the oocyte ZP (Jin et al., 2011). Most spermatozoa begin to react in the isthmus of the oviduct before reaching the ampulla (Hino et al., 2016; La Spina et al., 2016). Contact with the ZP in vitro probably makes it possible to complete a partial acrosome reaction. The most important function of the acrosome reaction is to induce changes in the sperm membrane (Okabe, 2016). The relocations of IZUMO1 and SPACA6, proteins essential for sperm–egg fusion, that occur after the acrosome reaction are illustrative examples of these changes (Sosnik et al., 2009; Barbaux et al., 2020; Satouh et al., 2012). The presence of these proteins on the sperm membrane, in addition to the classic markers Pisum sativum agglutinin, Peanut agglutinin lectins, or CD46, can be used as markers for the acrosome reaction (Ito and Toshimori, 2016). The acrosome and its disruption are both crucial for effective fertilization, as low fertilization rates are observed upon intracytoplasmic sperm injection of acrosome-intact sperm (Morozumi and Yanagimachi, 2005) or round spermatozoa lacking acrosomes (Dávila Garza and Patrizio, 2013).ZP penetrationThe ZP is a physical barrier between the oocyte and the follicular cells that forms from glycoproteins secreted from the primary follicles (Fig. 1 C). The human ZP consists of four glycoproteins (hZP1–hZP4; Harris et al., 1994). Mice, which have been used for most of the ZP studies in mammals, express only three ZP glycoproteins (mZP1–mZP3; Litscher and Wassarman, 2007). Analysis of mouse lines expressing human ZP proteins demonstrated that only hZP2 is important in human sperm–egg binding (Gupta, 2021). Experiments using purified native or recombinant human ZP proteins have shown that hZP1, hZP3, and hZP4 bind to the capacitated human spermatozoa and induce the acrosome reaction (Gupta, 2021). ZP1 is required for the structural integrity of the ZP (Chakravarty et al., 2008). To better understand the roles of ZP glycoproteins, further studies, particularly on ZP protein glycosylation, are needed. The species-specific binding of the ZP to sperm is presumably related to these carbohydrate moieties (Clark, 2014). The sialyl-Lewis(x) sequence is the major carbohydrate ligand for human sperm–egg binding (Pang et al., 2011). The current hypothesis that hZP1, hZP3, and hZP4 bind to capacitated sperm and hZP2 binds to sperm with intact acrosomes will need to be revisited due to the recent demonstration that the acrosome reaction takes place before ZP contact. Regardless, the role of the ZP in preventing polyspermy is clear. Indeed, ZP hardening is due to ZP2 cleavage by ovastacin, a protease released into the PVS by cortical granules after the first sperm–egg fusion (Burkart et al., 2012).Sperm–egg attachment and membrane fusionAfter penetration of the ZP, the sperm enters the PVS and can attach and fuse with the egg plasma membrane. The development of genetic knockout animal models has proven critical in determining the importance of various sperm and egg proteins in sperm–egg attachment and fusion. Surprisingly, genetic knockout studies revealed that many factors originally thought to be important for fertilization were in fact not necessary (reviewed in Okabe, 2018, 2015). The proteins from sperm and egg that are essential for sperm–egg membrane interaction and fusion are listed in
ProteinYear identifiedRole in fertilizationStructural featuresReferences
CD91999CD9 is expressed on the surface of the oocyte and accumulates during the attachment event; it may modulate the integrity of the oocyte membrane; its precise role in sperm–egg fusion remains unclearCD9 is a tetraspanin with four transmembrane domains and two extracellular loops (short and long)Miyado et al., 2000; Le Naour et al., 2000; Kaji et al., 2000; Chen et al., 1999; Umeda et al., 2020; Zimmerman et al., 2016; Zhang and Huang, 2012; Dahmane et al., 2019; Runge et al., 2007; Zhu et al., 2002; Chalbi et al., 2014; Rubinstein et al., 2006; Ziyyat et al., 2006
IZUMO12005IZUMO1 relocates to the equatorial region of the sperm head after the acrosome reaction; high-affinity binding of IZUMO1 to JUNO results in initial attachment of sperm and egg in the PVSThe protein has an N-terminal 4HB, followed by a β-hinge and an IgSF domain; the structure is stabilized by five disulfide bondsInoue et al., 2005; Ellerman et al., 2009; Young et al., 2015; Satouh et al., 2012; Aydin et al., 2016; Ohto et al., 2016; Nishimura et al., 2016; Kato et al., 2016
JUNO2014JUNO is expressed on the surface of the oocyte membrane and serves as the receptor of IZUMO1JUNO has structural similarity to folate receptors; it is a globular α/β protein composed of five α helices, three 310 helices, and four short β strands stabilized by eight disulfide bondsBianchi et al., 2014; Kato et al., 2016; Han et al., 2016; Jean et al., 2019; Yamaguchi et al., 2007; Aydin et al., 2016; Ohto et al., 2016
SPACA62014SPACA6 is expressed in sperm and localized to the equatorial segment after the acrosome reaction, but its specific role in sperm–egg fusion remains unknownThe three-dimensional structure of SPACA6 is currently unknown; SPACA6 is similar in organization to IZUMO1 with a signal peptide, followed by an α-helical domain, an IgSF domain, a transmembrane helix, and a cytoplasmic tailLorenzetti et al., 2014; Noda et al., 2020; Barbaux et al., 2020
TMEM952014TMEM95 is localized to the equatorial segment of sperm and is essential for sperm–egg fusion and male fertility in mice, but its specific role in sperm–egg fusion is currently unknownThe structure of TMEM95 is currently unknown; TMEM95 consists of a signal peptide, an N-terminal helix-rich region, a transmembrane helix, and a leucine-rich cytoplasmic domainPausch et al., 2014; Zhang et al., 2016; Noda et al., 2020; Fernandez-Fuertes et al., 2017; Lamas-Toranzo et al., 2020
SOF12020SOF1 is predicted to be a secreted factor essential for fusion; its role is still not fully understoodNo structural information to date; primary sequence analysis revealed the presence of conserved LLLL and CFNLAS motifs Noda et al., 2020
FIMP2020FIMP is involved in sperm–egg fusion; only the transmembrane form is important in fertilization, but its role is still not fully determinedNo structural information to date Fujihara et al., 2020
DCST1/DCST22021DCST1 and DCST2 are involved in sperm–egg fusion; stability of SPACA6 is regulated by DCST1/2; DCST1/DCST2 are evolutionary conserved in vertebrates and invertebratesNo structural information to date; contains six putative transmembrane helices Inoue et al., 2021
Open in a separate windowSperm IZUMO1In 2005, Inoue et al. discovered that homozygous Izumo1−/− mice are healthy and show normal mating behavior, but males are infertile. IZUMO1 is named after a shrine in Japan that honors the deity for marriage (Inoue et al., 2005). The spermatozoa of Izumo1−/− mice can undergo acrosomal reaction and penetrate the ZP but fail to fuse with oocytes. When the fusion step is bypassed using intracytoplasmic sperm injection, Izumo1−/− spermatozoa can fertilize oocytes, resulting in offspring; thus, IZUMO1 is only necessary at the adhesion/fusion stage of fertilization. An anti-IZUMO1 antibody, OBF13, completely abolishes gamete fusion by blocking IZUMO1 from binding to its receptor. There are four IZUMO family members (Ellerman et al., 2009), but in mice, IZUMO1 is the only paralog that is essential to fertilization (Inoue et al., 2005). IZUMO1 is a type I transmembrane protein consisting of 350 residues that is expressed exclusively in sperm (Inoue et al., 2005; Ellerman et al., 2009; Young et al., 2015). As sperm transit through the epididymis, IZUMO1 undergoes posttranslational modifications. In immature spermatozoa isolated from the proximal caput region of the epididymis, IZUMO1 is localized to both the acrosome and flagella of spermatozoa and is phosphorylated at two sites (S339 and S346; Young et al., 2015). In the cauda epididymis, IZUMO1 is found predominantly in the acrosome of spermatozoa and is phosphorylated at seven residues (S346, S352, S356, S366, T372, S374, and S375; Young et al., 2015). Cell-based fluorescence studies show that after the acrosomal reaction, IZUMO1 is relocated to the membrane surface in the equatorial segment of the acrosome (Satouh et al., 2012).Three crystal structures of the human and mouse IZUMO1 ectodomain were recently published (Aydin et al., 2016; Ohto et al., 2016; Nishimura et al., 2016). In one structure, IZUMO1 is in an upright conformation; however, other crystallographic structures are angled at the hinge region in a “boomerang” shape, which is also observed in solution small-angle x-ray scattering studies (Aydin et al., 2016). The structural discrepancy is not unusual, because the crystal lattice can induce distortions. The crystal structures of the human and mouse IZUMO1 ectodomain show that the extracellular region is organized into two domains, an N-terminal four-helix bundle (4HB) and an Ig-superfamily (IgSF) domain (Fig. 2 A; Aydin et al., 2016; Ohto et al., 2016; Kato et al., 2016). The two domains are connected by a β-hairpin that serves as a flexible hinge. There are five disulfide bonds, one buried at the protein core and four others that are solvent exposed on the surface. Three disulfide bonds connect the N-terminal 4HB domain to the hinge region, and the fourth links the hinge region to the IgSF domain. Interestingly, IZUMO1 shows marked similarities to two protozoan Plasmodium sp. parasite proteins: TRAP, which plays a critical role in gliding motility and host invasion (Song et al., 2012), and SPECT1, which plays a role in host cell fusion and hepatocyte invasion (Ishino et al., 2004; see text box).Open in a separate windowFigure 2.Structures of cellular factors involved in sperm–egg attachment and fusion. (A) Human IZUMO1 is shaped like a boomerang in an unbound state (Protein Data Bank [PDB] accession no. 5F4T). The 4HB, hinge, and IgSF domains are shown in orange, green, and cyan, respectively. (B) Human JUNO (PDB accession no. 5F4Q) belongs to the folate receptor family. (C) The structure of human IZUMO1–JUNO complex (PDB accession no. 5F4E) reveals that JUNO binds to IZUMO1 via the β-hairpin hinge, four residues from the 4HB domain and two from the IgSF domain. (D) Human CD9 (PDB accession no. 6K4J) adopts a conical shape formed by four transmembrane helices (TM1–TM4; blue) and two extracellular loops (SEL, pink; and LEL, red).Perspectives: Similarity to Plasmodium host invasion proteinsThe β-hinge region of IZUMO1 is highly similar to an extensible β-ribbon region in TRAP (root-mean-square deviation [RMSD] 1.4 Å; Nishimura et al., 2016). The TRAP β-ribbon has been proposed to undergo conformational changes upon binding to a host cell to mediate sporozoite gliding and host cell invasion. The IZUMO1 4HB domain shows structural similarities to Plasmodium berghei SPECT1 (RMSD 3.3 Å; Nishimura et al., 2016; Aydin et al., 2016). SPECT1 is required for cell traversal of sporozoites. Both SPECT1 and IZUMO1 adopt 4HBs with the same connectivity. In SPECT1, the 4HB is proposed to be a metastable structure that transitions from a solvent-accessible to a membrane-associated state. It has also been proposed that SPECT1 interacts with SPECT2, which has a membrane-attack complex/perforin domain, to form a pore. How the two proteins cooperatively mediate pore formation remains to be determined, but the similarity of IZUMO1 to proteins involved in parasite entry is intriguing.Oocyte JUNOIn 2014, Bianchi et al. made the groundbreaking discovery of the oocyte receptor for IZUMO1 (Bianchi et al., 2014). The group iteratively cloned and expressed the entire mouse oocyte cDNA library in mammalian cells and tested each clone for IZUMO1 binding using avidity-based extracellular interaction screening (AVEXIS; Kerr and Wright, 2012). Folate receptor δ (or folate receptor 4), which was aptly renamed JUNO, after the Roman goddess of marriage and fertility, was the only protein that bound to IZUMO1. Mouse JUNO shares 58% sequence identity with human folate receptors FOLR-α and FOLR-β but does not bind to folate (Kato et al., 2016; Han et al., 2016). Juno−/− mice show normal development and mating behaviors, but females are infertile, and eggs from Juno−/− mice are unable to fuse with wild-type sperm (Bianchi et al., 2014). Moreover, an anti-JUNO antibody incubated with human zona-free oocytes effectively blocks fertilization (Jean et al., 2019).While JUNO is primarily expressed on the surface of oocytes, it is also expressed on CD4+ CD25+ regulatory T cells, albeit at a much lower level (Yamaguchi et al., 2007). JUNO is highly expressed in unfertilized eggs but upon fusion with sperm is rapidly shed from the cell surface into extracellular vesicles (Bianchi et al., 2014). By the anaphase II stage, which takes place 30–40 min after fertilization, JUNO is barely detectable at the cell surface (Bianchi et al., 2014). The rapid removal of JUNO from the egg surface may help prevent the entry of more than one sperm into an oocyte.JUNO is a glycoprotein of 250 residues with a C-terminal glycosylphosphatidylinositol anchor. The crystal structures of human and mouse JUNO, both alone and in complex with human IZUMO1, were determined in 2016 (Aydin et al., 2016; Kato et al., 2016; Ohto et al., 2016; Han et al., 2016). The overall structure of human JUNO resembles structures of FOLR-α and FOLR-β with RMSDs of 1.1 Å and 1.0 Å, respectively. Like the folate receptors, JUNO has a compact, globular shape with five α helices, three 310 helices, and four short β strands stabilized by eight conserved disulfide bonds (Fig. 2 B). Despite its structural homology to folate receptors, five key residues in JUNO (A93, G121, Q122, R154, and G155) are not conserved compared with the folate-binding sites of FOLR-α and FOLR-β (Aydin et al., 2016). The aromatic and charged residues that in FOLR-α and FOLR-β anchor folate in the binding site through hydrogen bonds are replaced by alanine or glycine in JUNO, resulting in a larger cavity that cannot bind folate. Recombinant IZUMO1 binds to oocytes (and to nongamete human cells transfected with JUNO) but does not bind to oocytes that have been preincubated with an anti-JUNO antibody (Bianchi et al., 2014). The cocrystal structure of IZUMO1 in complex with JUNO reveals a 1:1 stoichiometry with a binding interface of ∼910 Å2 (Aydin et al., 2016). Biolayer interferometry, surface plasmon resonance, and isothermal titration calorimetry revealed that the complex of JUNO and IZUMO1 has a dissociation constant between 48 and 91 nM (Aydin et al., 2016; Ohto et al., 2016). The tight binding affinity results from an additive effect of extensive van der Waals, hydrophobic, and aromatic interactions, as well as two salt bridges. IZUMO1 binds to JUNO primarily via the β-hairpin hinge, with four residues from the 4HB domain and two from the IgSF domain also contributing to the binding (Fig. 2 C). In JUNO, the binding site is an elongated surface formed by the flanking regions of helices α1–α3 and loops L1–L3. The IZUMO1–JUNO interaction is not strictly species specific, as there is cross-species interaction between human IZUMO1 and hamster JUNO (see text box).Perspectives: Cross-species interactionsFertilization is a species-specific event, as sperm typically cannot fertilize eggs from a different species. The ZP provides an effective barrier against cross-species fertilization, but beyond this glycoprotein layer, IZUMO1-JUNO recognition is promiscuous. Human sperm cannot penetrate the hamster ZP, but they can fuse with zona-free hamster eggs (Inoue et al., 2005). Indeed, zona-free hamster eggs have been used to assess human sperm quality in fertility treatments. Using the ELISA-based AVEXIS platform, human IZUMO1 was confirmed to bind to hamster JUNO in solution (Bianchi and Wright, 2015). Like human IZUMO1, mouse and pig IZUMO1 also bind to hamster JUNO in solution (Bianchi and Wright, 2015). The results are consistent with the ability of human, mouse, and pig sperm to fuse with zona-free hamster eggs (Creighton and Houghton, 1987; Hanada and Chang, 1972).Hamster and human JUNO are highly similar, with a sequence identity of 73%; however, eight residues at the IZUMO1–JUNO interface are not conserved. To understand the cross-species specificity, a homology model of hamster JUNO was generated based on the crystal structure of human JUNO. Despite key substitutions in hamster JUNO, the IZUMO1-binding site preserves the same structural architecture and physiochemical characteristics as human JUNO, with the exception of E45. In human JUNO, E45 forms a key salt bridge at the IZUMO1–JUNO interface. How an E45L substitution in hamster JUNO is able to maintain binding remains unclear. It was previously shown that E45 is critical for human JUNO recognition to IZUMO1, as E45A or E45K mutations severely reduced the interaction (Aydin et al., 2016). It may be possible that other interactions between hamster JUNO and human IZUMO1 compensate for the loss of the critical salt bridge. A crystal structure of hamster JUNO in complex with human IZUMO1 would provide important insight into the molecular basis of cross-species specificity in IZUMO1–JUNO recognition.The binding sites on both IZUMO1 and JUNO have been verified by alanine-substitution experiments (Ohto et al., 2016). W62 and L81 in JUNO and W148 in IZUMO1 play critical roles at the interface, as substitution of these residues by alanine dramatically reduces binding affinity (Ohto et al., 2016). These residues are strictly conserved across mammalian species. To verify the biological relevance of the IZUMO1–JUNO interface, the binding of oocytes to COS-7 cells expressing wild-type IZUMO1 or to mutants with one or more mutations to residues proposed to be important in JUNO binding was tested. Mutating W148, K154, H157, I158, R160, or L163 in IZUMO1 significantly reduced oocyte binding. COS-7 cells that expressed IZUMO1 with multiple mutations at the JUNO-binding interface showed a complete lack of binding to oocytes (Ohto et al., 2016). These results confirmed the JUNO-binding residues identified in the crystal structures and biophysical studies (Ohto et al., 2016).Oocyte CD9The importance of CD9 in sperm–egg fusion was first described in 1999 (Chen et al., 1999) and confirmed in 2000 (Miyado et al., 2000; Le Naour et al., 2000; Kaji et al., 2000; Chen et al., 1999). CD9 is expressed on the plasma membrane of oocytes, and an anti-CD9 antibody inhibits sperm–egg fusion in a dose-dependent manner (Chen et al., 1999). Interestingly, anti-CD9 antibodies do not block sperm from binding to oocytes but instead prevent the fusion of sperm and egg membranes (Miyado et al., 2000; Le Naour et al., 2000). These findings are consistent with mouse studies, which showed that CD9−/− mice develop normally and that male mice are fertile but female mice have dramatically reduced fertility (Miyado et al., 2000; Le Naour et al., 2000; Kaji et al., 2000). When the sperm–egg fusion step is bypassed by injecting capacitated sperm into the cytoplasm of CD9−/− oocytes, the fertilized eggs show normal implantation efficiencies, and embryos develop normally.CD9 belongs to the tetraspanin superfamily and is 228 amino acids long. It has four membrane-spanning domains (TM1–TM4) linked by a short extracellular loop (SEL) between TM1 and TM2 and a large extracellular loop (LEL) between TM3 and TM4 (Fig. 2 D). The transmembrane regions are highly conserved among tetraspanins, with sequence divergence only in the extracellular loops. The first structure of CD9 was recently determined to 2.7 Å resolution (Umeda et al., 2020). The four transmembrane helices tilt toward the cytoplasmic membrane interface to form a cone-shaped structure that creates a spacious cavity in the intramembranous region (Fig. 2 D). This is reminiscent of the CD81 structure; CD9 and CD81 have ∼60% sequence similarity and the same overall fold (RMSD of 1.9 Å; Umeda et al., 2020; Zimmerman et al., 2016). Previous studies have demonstrated the localization of tetraspanins in curved regions of cell membranes (Zhang and Huang, 2012; Dahmane et al., 2019). As CD9 clusters at the contact region of the egg and sperm membranes, the tight array of cone-shaped CD9 may increase the curvature of the oocyte membrane, effectively causing it to protrude. CD9-knockout oocytes produce short and sparse microvillus structures with a large radius of curvature of microvillar tips, which results in impaired fusion ability with spermatozoa (Runge et al., 2007).The relative lengths of the SEL and LEL of tetraspanins control access to the intramembranous cavity. In silico analysis revealed that the LEL undergoes a conformational change between the open and closed states during binding partner recognition (Umeda et al., 2020). In the closed-state CD9 structure, the LEL is weakly associated with the SEL. In the open state, the LEL moves away from the SEL, thereby allowing access to the intramembranous cavity. The importance of the LEL in fertilization was demonstrated in a domain-swapping experiment, in which the LEL of a fertilization-incompetent tetraspanin CD53 was swapped with its equivalent section from CD9. The CD9–CD53LEL chimera had dramatically reduced fertilization competency, whereas the chimera with the LEL from CD9, CD53–CD9LEL, was ∼50% competent (Umeda et al., 2020). This suggests that additional regions such as the SEL and transmembrane domains are also important in fertilization. Alanine-substitution experiments on residues within the LEL have produced conflicting results. Mutation of the 173SFQ175 motif in the murine CD9 LEL suggests that these residues are essential for fertilization (Zhu et al., 2002). However, the murine 173SFQ175 LEL region is not conserved in human CD9 (175TFT177). A triple-alanine mutations of this region in both murine and human CD9 revealed that contrary to previous findings, both mutants rescued fertilization in CD9−/− oocytes (Umeda et al., 2020). Further studies are required to probe the roles of specific CD9 LEL, SEL, and transmembrane residues in fertilization.Like other tetraspanins, CD9 can act as a scaffolding protein to bring together multiple protein partners to execute a biological function. For instance, CD9 associates with Igs, integrins, and other adhesion receptors and proteins (reviewed in Charrin et al., 2014). Recently, interaction studies using human sperm and mouse oocytes revealed that IZUMO1 and JUNO colocalize with CD9 on the surface of the egg during sperm–egg attachment. Along with sperm IZUMO1, egg CD9 accumulates at adhesion area surroundings, suggestive of a cis interaction with egg JUNO (Chalbi et al., 2014; Ravaux et al., 2018). In the same context, by measuring the force necessary to break contact between one sperm and an egg, it was suggested that CD9 induces the clustering of sperm receptors on the oocyte membrane, generating fusion-competent sites (Jégou et al., 2011).Single-particle cryoelectron microscopy studies of CD9 in complex with EWI-2 provide insights into how CD9 engages with its targets. EWI-2 belongs to the IgSF, with four to eight predicted IgSF domains and a single-pass transmembrane anchor. EWI-2 is a major binding partner to both CD9 and CD81 (Runge et al., 2007; Umeda et al., 2020; Rubinstein et al., 2006). An anti-IgSF8 antibody had moderate inhibitory effects on sperm–egg binding, suggesting that mouse EWI-2 participates in gamete interactions (Glazar and Evans, 2009). Cryoelectron microscopy revealed that a 2:2 heterotetrameric arrangement of the extracellular domains of two EWI-2 molecules forms a tight dimer and that the EWI-2 transmembrane helix is sandwiched by two CD9 molecules. The transmembrane helix of EWI-2 interacts with TM3 and TM4 of CD9 via hydrophobic residues (Umeda et al., 2020). The nonspecific nature of the transmembrane hydrophobic interactions in the CD9–EWI-2 complex may explain the promiscuous nature of tetraspanins.Newly identified players in mammalian fertilizationThe use of CRISPR-Cas9 technology has led to the recent identification of six new factors essential for mammalian fertilization: SPACA6, TMEM95, SOF1, FIMP, and DCST1/DCST2.Sperm SPACA6In 2014, Lorenzetti et al. characterized a mutant mouse line that had a deletion removing Spaca6 (Lorenzetti et al., 2014). Male homozygous knockout mice were infertile with a phenotype that closely resembles that of Izumo1-deficient mice. Subsequent studies by two other groups confirmed that Spaca6 deletion in male mice results in infertility, although mating behavior is normal and sperm are motile and morphologically normal (Noda et al., 2020; Barbaux et al., 2020). Fertility could be restored by a transgene (Noda et al., 2020). In a human zona-free in vitro fertilization assay, an anti-SPACA6 antibody reduced fertilization rates by threefold (Barbaux et al., 2020).Recovery of oocytes from female mice that were mated with Spaca6−/− male mice revealed that the spermatozoa were trapped in the PVS. This indicates that knockout spermatozoa migrate through the female genital tract to the oocyte and penetrate the ZP but fail to fuse with the oocyte membrane. When Spaca6−/− sperm was injected into the cytoplasm of oocytes to bypass the membrane fusion step, fertilization was successful, and the fertilized eggs showed normal embryonic development, suggesting that SPACA6 does not play a critical role downstream of sperm–egg fusion.SPACA6 is primarily expressed in testis, with low levels of expression in the epididymis, seminal vesicle, and ovary (Noda et al., 2020; Lorenzetti et al., 2014). Orthologues of SPACA6 have been annotated in bull, hamster, human, mouse, rat, and zebrafish (Noda et al., 2020). In fresh spermatozoa, SPACA6 is not detected on the plasma membrane; rather, it is localized underneath the membrane of the sperm head. After the acrosomal reaction, SPACA6 relocates to the equatorial segment of the sperm head, with reduced levels detected in the midpiece, and completely diminishes from the neck region (Barbaux et al., 2020). Immunofluorescence staining revealed that the localization of IZUMO1 is unaffected in Spaca6−/− sperm before and after the acrosomal reaction (Barbaux et al., 2020). To verify this result, Spaca6−/− male mice were mated with female mice and oocytes were extracted and immunostained with an anti-IZUMO1 antibody revealing that IZUMO1 distribution in Spaca6−/− spermatozoa was identical to that in wild-type spermatozoa (Barbaux et al., 2020). This confirmed that SPACA6 does not affect IZUMO1 localization.Both IZUMO1 and SPACA6 belong to the IgSF and are expressed in sperm and localized to the equatorial segment upon the acrosomal reaction (Noda et al., 2020). The phenotypes of knockout mutants are highly similar as well. The similarities also extend into their domain organization. Both proteins have an N-terminal signal peptide, followed by a helical domain, a single IgSF domain, a single N-linked glycosylation site, a monotopic transmembrane helix, and a short cytoplasmic tail (Noda et al., 2020). Despite these similarities, the proteins are not redundant, as both cell-based and mouse studies show that one cannot compensate for the lack of the other (Barbaux et al., 2020). Moreover, SPACA6 does not accumulate at the interface with IZUMO1, and SPACA6-expressing COS-7 or HEK293T cells do not bind to the surface of the oocyte (Inoue et al., 2015; Noda et al., 2020). No interaction was detected between SPACA6 and IZUMO1 by coimmunoprecipitation from testis extracts (Noda et al., 2020). How SPACA6 interacts with other sperm and oocyte proteins to mediate sperm–egg adhesion and fusion remains to be determined.Sperm TMEM95A genome-wide analysis designed to reveal genetic associations with infertility in bulls revealed the essential role of TMEM95 in fertility (Pausch et al., 2014). A nonsense mutation that introduces a premature stop codon in Tmem95 diminishes male fertility, although it does not significantly affect sperm morphology or motility (Pausch et al., 2014). TMEM95 is conserved in primary sequence among bull, hamster, mouse, rat, and humans (Zhang et al., 2016; Noda et al., 2020). In bulls, TMEM95 is expressed in spermatozoa and is localized on the acrosome, on the equatorial segment and on the connecting piece (Pausch et al., 2014). Bull spermatozoa with a knockout mutation in Tmem95 are unable to fuse with oocytes, suggesting that TMEM95 is required for sperm–oocyte fusion (Fernandez-Fuertes et al., 2017).RT-PCR analysis revealed that in mice Tmem95 is expressed exclusively in testis. Expression begins on day 21 postpartum when spermiogenesis begins. Tmem95−/− mice that carry a 1,919-bp deletion in the Tmem95 locus have normal mating behavior but males are infertile (Noda et al., 2020). The Tmem95−/− spermatozoa have normal morphology and motility and bind to oocytes; however, the mutant sperm have impaired ability to fuse with oocytes and accumulate in the PVS. Expression of a Tmem95 transgene in Tmem95−/− male mice restored fertility (Noda et al., 2020). The fertility of Tmem95−/− female mice is unaffected (Noda et al., 2020). Consistent with the study by Noda et al., Lamas-Toranzo et al. also found that Tmem95−/− male mice were infertile (Lamas-Toranzo et al., 2020).In silico analysis suggested that TMEM95 shares organizational similarities with IZUMO1 (Zhang et al., 2016). Like IZUMO1, it is a type I single-pass transmembrane protein with a signal peptide at the N terminus, a helix-rich N-terminal region, and a transmembrane helix. TMEM95 has an additional leucine-rich cytoplasmic domain compared with IZUMO1. Examination of the localization of IZUMO1 in acrosome-reacted wild-type and Tmem95−/− sperm revealed no difference in IZUMO1 translocation (Lamas-Toranzo et al., 2020). Moreover, TMEM95 disappears after the acrosomal reaction (Fernandez-Fuertes et al., 2017). Since IZUMO1 relocates to the equatorial segment only after the acrosomal reaction, this suggests that TMEM95 and IZUMO1 function independently. Experiments using the AVEXIS platform showed that TMEM95 does not bind to JUNO or IZUMO1 (Lamas-Toranzo et al., 2020). In contrast, coimmunoprecipitation studies using HEK293T cells coexpressing IZUMO1 and TMEM95 suggested that IZUMO1 does bind to TMEM95 (Noda et al., 2020). Further studies are required to verify whether or not IZUMO1 and TMEM95 interact.Sperm SOF1Another new molecular player important for male fertility identified by Noda and collaborators using CRISPR-Cas9–mediated gene knockout was a gene called 1700034O15Rik (also known as Llcfc1; Noda et al., 2020). The gene was aptly renamed SOF1 (sperm–oocyte fusion required 1). SOF1 is widely conserved in mammals and is highly expressed in the testis. SOF1 is predicted to be a 147-residue secreted protein with conserved LLLL and CFN(L/S)AS motifs. These motifs are observed in the DUF4717 family of proteins that have an unknown function but are exclusively found in eukaryotes. SOF1 reportedly undergoes posttranslational modifications during sperm maturation, and it was detected as a protein singlet in testicular germ cells but a doublet in acrosome-intact spermatozoa (Noda et al., 2020).The sterility of Sof1−/− male mice is likely due to defective membrane fusion (Noda et al., 2020). The morphology and motility of Sof1−/− spermatozoa are similar to wild type; however, when used for in vitro fertilization with cumulus-intact oocytes, Sof1−/− spermatozoa do not fuse or fertilize oocytes and accumulate in the PVS. This inability to fuse was also observed using zona-free oocytes, but sperm continued to bind to the oocyte membrane, suggesting a role in either sperm–egg fusion or in the control of sperm–egg adhesion properties. Expression levels and localization of IZUMO1 are not affected in Sof1−/− spermatozoa before or after the acrosome reaction. Thus, sterility in Sof1−/− male mice is not due to a disruption of IZUMO1.Sperm FIMPCRISPR-Cas9–mediated deletion of Fimp (also known in mice as 4930451I11Rik) results in failure in sperm–egg fusion in mice (Fujihara et al., 2020). Similar to SOF1, FIMP is also a small protein of 132 amino acids that is highly expressed in the testis. The expression of this testis-specific gene is first observed 20 d after birth. The protein is detected in two distinct isoforms: membrane anchored and secreted. Only the transmembrane form appears to be critical for sperm-oocyte fusion in mice. Fimp−/− mice have normal testicular and sperm morphologies, and Fimp−/− spermatozoa penetrate the ZP but fail to fuse with oocytes (Fujihara et al., 2020). In an in vitro fertilization assay using zona-free oocytes, Fimp−/− sperm ability to fuse are severely reduced. IZUMO1 localization and expression levels in Fimp−/− mice are similar to the wild type. FIMP localizes to the equatorial segment membrane, but the FIMP-mCherry signal disappeared in 40% of the acrosome-reacted sperm (Fujihara et al., 2020). In contrast to IZUMO1, it does not appear FIMP is involved in the initial attachment step, as FIMP-expressing cells do not bind oocytes. The precise role of FIMP in sperm–egg fusion remains unclear.Sperm DCST1/DCST2As identified by gene disruption and complementation experiments, the evolutionarily conserved factors dendrocyte expressed seven transmembrane protein (DC-STAMP) domain-containing 1 and 2 (DCST1/DCST2) are required for gamete fusion (Inoue et al., 2021). Individual or double gene deletion results in male sterility with the same phenotype as that of Izumo1−/− or Spaca6−/− knockouts. Although their molecular mechanism of action is still unknown, DCST1 and DCST2 function might be intrinsically related to SPACA6. Surprisingly, while the rescued double transgenic males had normal fertility, SPACA6 was not detected. The protein stability of SPACA6 may be differently regulated by DCST1/DCST2 and IZUMO1 (Inoue et al., 2021).Molecular mechanism of sperm–egg fusionWhen the acrosome-reacted spermatozoon reaches the PVS, it is primed to interact and fuse with the egg. The molecular mechanism of mammalian sperm–egg attachment and fusion requires a complex sequence of events that for the most part remain a mystery (Fig. 3). However, new insights into some of the steps have now been obtained through structural, biophysical, and genetic analyses of the proteins essential to the adhesion fusion process.Open in a separate windowFigure 3.Current model of sperm–egg attachment and fusion. (A) Acrosome reaction. After the acrosome reaction, IZUMO1 (blue), SPACA6 (purple), and TMEM95 (violet) colocalize to the equatorial regions of sperm. FIMP (pink) appears to function before the acrosome reaction. There are conflicting data on whether or not TMEM95 interacts with IZUMO1. SOF1 (turquoise) is a secreted sperm protein. DCST1 (green) and DCST2 (orange) are transmembrane proteins implicated in regulating the protein stability of SPACA6. (B) Initial attachment. After the sperm reaches the PVS, it attaches to the egg. IZUMO1 is localized on the equatorial segment of acrosome-reacted sperm and its counterpart receptor, JUNO (yellow), on the oocyte membrane. JUNO specifically recognizes and binds to IZUMO1 in a monomeric conformation. IZUMO1 binding to JUNO drives the accumulation of CD9 (pink) at the sperm–egg interface to form a physical anchor that holds the sperm and oocyte membranes in proximity. (C) IZUMO1 multimerization. After the initial IZUMO1–JUNO attachment, the complex undergoes a dimerization event. The trigger for IZUMO1 oligomerization is not fully understood; however, colocalization analysis revealed the presence of PDI (gray) on the sperm surface. JUNO is thought to be shed from the oolemma and into the PVS after fertilization. (D) Fusogen recruitment. The bona fide sperm–egg fusogen remains a mystery. However, data suggest that IZUMO1 forms a scaffold to recruit the gamete fusion complex. The roles of SPACA6, TMEM95, and SOF1 remain unclear, but these proteins likely play roles in fusion. (E) Fusion pore formation. The merger of the egg and sperm membranes requires modulation of the membrane architecture. The fusogen is thought to catalyze the formation of a hemifusion intermediate, which is a stalk-like structure where the outer leaflets of the sperm and egg membrane bilayers mix. Subsequently, the inner bilayer leaflets mix to form the fusion pore. The precise mechanism of this step will require the identification of the sperm–egg fusogen. Created with BioRender.Initial attachmentAt the molecular level, the first step in the attachment of the sperm to the egg involves binding of IZUMO1, which is localized on the equatorial segment of acrosome-reacted sperm (Satouh et al., 2012), to its counterpart receptor JUNO on the oocyte membrane (Fig. 3 A). JUNO is monomeric when it binds to IZUMO1 (Fig. 3 B; Inoue et al., 2015). Comparison of the crystal structures of IZUMO1 with and without JUNO suggests that there is a binding-induced conformational change in IZUMO1 whereby the 4HB domain moves ∼20° to adopt an upright conformation (Aydin et al., 2016). IZUMO1 binding to JUNO drives the accumulation and local membrane organization of CD9. The accumulation of CD9 at the sperm–egg interface causes the egg membrane to protrude toward the sperm membrane (Chalbi et al., 2014).IZUMO1 multimerizationAfter the initial IZUMO1–JUNO attachment, Inoue et al. suggested that the IZUMO1–JUNO complex undergoes a multimerization event that is critical for sperm–egg fusion (Fig. 3 C; Inoue et al., 2015). Bimolecular fluorescence complementation and photon-counting histogram analyses were used on a cultured cell–oocyte system to reveal that IZUMO1 forms a multimer at the cell–oocyte interface, but not on the rest of the cell surface. After IZUMO1 oligomerization, JUNO is not detected at the cell surface and presumably is shed (Inoue et al., 2015).Inoue et al. suggested that the trigger for IZUMO1 multimerization involves a protein disulfide isomerase (PDI). Localization studies revealed the presence of PDIs on the sperm surface (Fig. 3 C; Ellerman et al., 2006). PDIs are responsible for proper folding of extracellular or membrane proteins during the maturation process in the endoplasmic reticulum. Western blot and proteomic analysis detected at least four PDI members on the sperm surface, PDI, ERp57, ERp72, and P5. Interestingly, PDI inhibitors reduce sperm–egg fusion in vitro in a dose-dependent manner (Ellerman et al., 2006), and a membrane-impermeable thiol-reactive reagent significantly reduces cell–oocyte binding (Inoue et al., 2015). To identify those PDIs that function in gamete fusion, sperm were preincubated with antibodies that specifically blocked each PDI member, and the ability of the spermatozoon to fuse with the oocyte was assessed (Ellerman et al., 2006). This revealed ERp57 is critical for gamete fusion. On the IZUMO1 ectodomain, 10 cysteines form five disulfide bonds. Four of the five disulfide bonds are located on the surface and are solvent accessible. The N-terminal helical domain of IZUMO1 was proposed to undergo a collapse or becomes buried at the oligomeric interface (Inoue et al., 2015). ERp57 and/or other PDIs may catalyze a thiol-disulfide exchange during this conformational rearrangement.Fusogen recruitmentAfter IZUMO1 multimerization, the next step is thought to involve the recruitment of the bona fide human sperm–egg fusogen. Dimerized IZUMO1 was suggested to directly recruit a tight binding unidentified oocyte receptor (Fig. 3 D; Inoue et al., 2015). The identity of the gamete fusion complex remains unknown. SPACA6 was proposed to interact with IZUMO1 to mediate the binding of an oocyte receptor (Lorenzetti et al., 2014). However, coimmunoprecipitation studies of testis extracts did not show an interaction between SPACA6 and IZUMO1, whereas coimmunoprecipitation analysis using HEK293T cells showed interactions between IZUMO1 and SPACA6, FIMP, TMEM95, and SOF1 (Noda et al., 2020). Expression of all five proteins on HEK293T cells did not lead to fusion with zona-free oocytes, suggesting the recruitment of a yet-to-be discovered fusogen is required (Noda et al., 2020). Intriguingly, IZUMO1 and SPACA6 both contain an IgSF domain. While IgSF domains are known to facilitate protein–protein interactions, their role and importance in sperm–egg fusion is unknown (see text box). A complete understanding of the interplay of these proteins and the composition of the human gamete fusion machinery will require additional biochemical and functional experimentation.Perspectives: Role of the IgSF in sperm–egg fusionMany of the essential protein players involved in sperm–egg attachment and fusion contain IgSF domains (Bork et al., 1994; Harpaz and Chothia, 1994). The IgSF fold displays a common core composed of four anti-parallel β strands sandwiched by a second set of three to five β strands. Based on the number of strands and relative locations, several distinct subtypes have been defined. Most common are the variable (V) and constant (C) Ig domains, while a third type (I) is an intermediate structure between the V and C types.Human IZUMO1, SPACA6, EWI-2, and EWI-F (CD9P-1) all contain at least one IgSF domain. The importance of IgSF domains in sperm–egg fusion is not limited to humans. In lower eukaryotes, the HAP2/GCS1 sperm–egg fusogen contains a C-type IgSF domain. Moreover, Caenorhabditis elegans encodes a transmembrane protein with an IgSF domain, termed SPE45, that is required for gamete fusion (Nishimura et al., 2015; Singaravelu et al., 2015). Interestingly, the IgSF domains of SPE45 and IZUMO1 share a common function, despite only 8.7% sequence identity (Nishimura et al., 2015, 2016). This was demonstrated by the finding that a chimeric SPE45 that contains the murine IZUMO1 IgSF domain had ∼77% of the activity of wild-type SPE45 (Nishimura et al., 2015). The IgSF domain may act as a scaffold to recruit binding partners in cis and/or in trans. Various organisms have IgSF proteins that act during gamete interactions, indicating the widespread utility of IgSF domains in fertilization.The importance of the IgSF domains in human IZUMO1 and C. elegans SPE45 raises the question of whether there are special features unique to these IgSF domains. The crystal structure of human IZUMO1 revealed a novel Ig domain fold with a 2+5 β-sheet arrangement. The IZUMO1 IgSF domain is the only known member to adopt a 2+5 arrangement, thus suggesting a new IgSF subtype. It is not clear whether the IgSF domain in SPE45 adopts a similar structure to the IZUMO1 IgSF domain. The precise roles played by these IgSF domains in IZUMO1, SPACA6, EWI-2, and SPE45 are currently unknown. However, the IgSF domains in other cell surface proteins often form homo- or heterodimers. Structures of IgSF oligomers reveal that all regions of the domain surface can be used for interaction with other molecules. Noncovalent association of IgSF-type domains usually occurs through the exposed faces of the β sheets. Continued studies will undoubtedly reveal important functions of IgSF proteins in gamete fusion.Fusion pore formationThe merger of the egg and sperm membranes is an energetically unfavorable process and must require modulation of the membrane architecture in order to form a fusion pore (Fig. 3 E). The formation of a fusion pore typically proceeds through one of two mechanisms, either via a hemifusion intermediate or via direct fusion (Chernomordik and Kozlov, 2005). In the case of hemifusion, the fusion of the two membranes occurs through the sequential mergers of each pair of bilayer leaflets. First, the outer membrane leaflets contact and mix to form the hemifusion stalk intermediate. This is followed by mixing of the inner leaflets to form the fusion pore. Enveloped viral-cell fusion proceeds through a hemifusion intermediate that is catalyzed by a viral fusion glycoprotein as previously discussed (Harrison, 2015; Sapir et al., 2008; White et al., 2008; Podbilewicz, 2014). The viral fusogens all contain distinctive hydrophobic fusion peptides that are inserted into the host target membrane when triggered. In direct fusion, proteins on both membranes arrange into complexes at the site of fusion and bind in trans to bring the two membranes together. This forms a continuous connection between the two protein-lined pores to allow for content mixing. In yeast vacuolar fusion, two proteolipid hexamers formed in trans by the V0 subunit of vacuolar H+-ATPase establishes a bridging channel/pore between the two membranes (Peters et al., 2001; Chernomordik and Kozlov, 2003). The pore is subsequently opened through a Ca2+-triggered conformational change that expands the proteolipid hexameric complex.It is thought that sperm–egg fusion also proceeds via a membrane hemifusion intermediate, similar to viral-cell fusion. This is at least true in lower eukaryotic cells, in which the sperm–egg fusogen HAP2/GCS1 has a similar overall structure to the class II viral fusogens (Fédry et al., 2017), such as tick-borne encephalitis virus E glycoprotein. No evidence for HAP2/GCS1 orthologues has been found in vertebrates or mammals; thus, in an early vertebrate ancestor, a new fusogen likely replaced HAP2/GCS1 (Vance and Lee, 2020). The structures of CD9, IZUMO1, and JUNO lack characteristics common to viral fusogens such as the prototypical hydrophobic fusion peptide (Aydin et al., 2016). Moreover, no readily identifiable fusion peptides were detected upon sequence analysis of SOF1, DCST1/DCST2, TMEM95, FIMP, or SPACA6. Furthermore, cell fusion assay experiments show that the sperm proteins alone or together are not able to trigger cell–cell fusion (Fujihara et al., 2020; Noda et al., 2020; Barbaux et al., 2020; Lamas-Toranzo et al., 2020). Thus, additional factors remain to be identified that are essential for fusion pore formation.Concluding remarksWhile significant advances have been made to fully understand the molecular mechanism of fertilization, many questions are still outstanding. The identification of proteins involved in the sperm–egg fusion process remains the holy grail in reproductive biology. Understanding the interplay of all the partners involved has the potential to impact multiple areas of biology. Identifying the full complement of proteins involved in sperm–egg attachment and fusion will allow the mapping of genotype–phenotype correlations and improve diagnostic tests for people suffering from infertility. Understanding the mechanisms of sperm–egg fusion will also reveal ways to improve assisted reproductive technologies for humans and animals. High and predictable fertility rates for cows, pigs, chickens, and sheep are essential for efficient food animal production. Finally, although generally safe and effective, current hormone-based contraceptives may lead to adverse side effects that discourage many from long-term use. The safety and acceptability of contraceptives are particularly important for women, since they bear the greatest burden of contraceptive side effects. It is important to innovate and develop new alternative contraceptives that better meet the reproductive needs and desires of women and couples. Molecules that disrupt sperm–egg protein–protein interactions by binding to the sperm or egg protein side of the axis should result in a potent contraceptive. These reasons underscore why understanding the mechanisms of sexual fertilization is one of the most crucial biological questions.  相似文献   

设为首页 | 免责声明 | 关于勤云 | 加入收藏

Copyright©北京勤云科技发展有限公司  京ICP备09084417号