Molecular Human Reproduction vol.3 no.7 pp.599–637, 1997 Carbohydrates and fertilization: an overview Susan Benoff1,2,3 1Department of Obstetrics and Gynecology, North Shore University Hospital, Manhasset, New York and 2Departments of Obstetrics and Gynecology and Cell Biology, New York University School of Medicine, New York, New York, USA 3To whom correspondence should be addressed at: North Shore University Hospital, 300 Community Drive, Boas–Marks Biomedical Science Research Center, Room 125, Manhasset, New York 11030, USA Initial sperm–egg binding in mammals involves recognition of glycosylated proteins of egg zonae by glycosylated proteins on sperm surfaces. Egg zona protein structure is relatively simple, and has been strongly conserved. Species specificity must reside in the carbohydrate modifications on the egg surface, and in the co-ordinated assembly of a unique cohort of sperm proteins at capacitation. Fruitful advances have been made along four lines. Oligosaccharide structures capable of binding spermatozoa have been dissected by in-vitro synthesis and binding experiments, informed by the general advance of knowledge of protein glycosylation processes. Site-specific mutagenesis of zona proteins and their expression in tissue culture have identified glycosylation sites involved in species-specific sperm binding. Antibody and lectin labelling studies show a continuing process of remodeling of glycosylated sperm surface epitopes within a set of stable compartments during epididymal transit and capacitation of spermatozoa. Characterization of sperm– egg binding proteins from a variety of mammalian species shows that a different set of effectors induce acrosome reactions in each species, with each set including one or more sugar-recognizing proteins. Sequencing of some of these effectors suggests that each group may form a supermolecular complex to induce a species-specific acrosome reaction, with the functional activities distributed in a species limited or non-limited manner among the individual proteins. Key words: fertilization/lectins/mammals/oligosaccharides/zona pellucida Introduction A cascade of cell–cell interactions is required to achieve fertilization in mammals. Upon ejaculation and residence in the female reproductive tract, mammalian spermatozoa then: (i) bind to oviductal cells; (ii) undergo capacitation; (iii) respond to chemoattractants; (iv) penetrate the cumulus complex; (v) bind to the zona pellucida; (vi) undergo an acrosome reaction; (vii) enter into the perivitelline space after passage through the zona; (viii) bind to the egg plasma membrane; (ix) fuse with the egg plasma membrane, and (x) be drawn into the ooplasm and undergo nuclear decondensation and activate zygote development (Yanagimachi, 1994; Snell and White, 1996; Myles and Primakoff, 1997). Although the sperm surface plays an active role in the regulation of this pathway, it is clear that carbohydrates of one of the three barriers that the spermatozoa must cross, i.e. the zona pellucida, strongly contribute to the modulation of steps (v) to (vii). The zona pellucida, which surrounds the plasma membrane of mammalian oocytes, is a relatively thick, highly glycosylated extracellular matrix. This matrix plays at least four important roles during development: (i) regulation of normal endocrine profiles during folliculogenesis; (ii) formation of a barrier to heterospecific fertilization; (iii) formation of a block to polyspermic fertilization in some mammals; and (iv) protection of the embryo during preimplantation development (Bronson and McLaren, 1970; Keenan et al., 1991; Yamagimachi, 1994; © European Society for Human Reproduction and Embryology Liu et al., 1996). This review will focus on the contribution of carbohydrates to the regulation of zona functions (ii) and (iii). Interactions between carbohydrates on the extracellular matrix of the egg and glycoproteins on the sperm surface provide the molecular basis for gamete recognition and adhesion in species as diverse as sea urchins (Garbers, 1989), ascidians (Rosati, 1985), amphibians (Gerton, 1985), algae (Callow et al., 1985), mouse (Wassarman, 1989, 1990) and man (Benoff et al., 1995a). In general, a compulsory order of initial interactions of mammalian spermatozoa and eggs is observed: (i) attachment or adhesion; (ii) tight binding, and (iii) secondary binding (Wassarman, 1987; Wolf, 1996). Attachment is defined as relatively loose, non-specific association of the sperm head plasma membrane with the zona pellucida. Attached spermatozoa can be readily dissociated from the zona by pipetting. Tight binding is the relatively tenacious interaction between acrosome-intact, capacitated spermatozoa and the zona pellucida which is species-specific and regulated by glycoproteins. Spermatozoa which are tightly bound to the zona can not be removed by vigorous pipetting. The physiological acrosome reaction occurs after tight binding. Secondary binding is thought to maintain zona binding by acrosome-reacted spermatozoa. Herein, the molecules which participate in the complementary, carbohydrate-based binding processes which regulate initial specific interactions between mammalian spermatozoa and ovum, and thus fertilization, will be reviewed. 599 S.Benoff The first section of this review summarizes the recent advances in glycobiology that have permitted the dissection of the mechanisms by which carbohydrates regulate sperm– zona pellucida binding and induction of the acrosome reaction. In the second section, the structure of the zona pellucida in a variety of mammalian species is compared and contrasted. In this discussion, the role of molecular biological techniques in the elucidation of both the primary structure of zona glycoproteins and the identification of potential glycosylation sites which contribute to the formation of the primary sperm binding site is emphasized. Recent data concerning the structure of the oligosaccharide chains recognized by receptors on the sperm surface are also presented. The third section outlines a large number of studies in which carbohydrate moieties on the sperm surface have been examined. Changes in glycosylation of membrane-associated sperm proteins and lipids accompanies epididymal maturation and the acquisition of sperm fertilizing potential. Although these are thought to play a role in generation of a periacrosomal plasma membrane capable of calciumstimulated fusion with the outer acrosomal membranes, the argument is made that the primary role of sperm surface glycan structures is as antigenic determinants in natural immunoregulation of gamete function. The fourth section is of particular interest to me as it presents results of informed research on the identification of sperm receptors for zona carbohydrate ligands. Recent data from my laboratory on how expression of sperm surface carbohydrate binding sites can affect the outcome of human in-vitro fertilization (IVF) is included in this review. From the findings presented in this section, it will become evident that fertilization mechanisms in man cannot necessarily be equated with those in animal models. The concluding section summarizes the complex roles of carbohydrates in the sequence of sperm–egg interactions leading to fertilization and addresses issues that remain to be resolved. Recent advances in carbohydrate biochemistry The last 10 years have been witness to major advances in our understanding of the function of glycoproteins (Lis and Sharon, 1993; Varki, 1993). This is the result of the development of a variety of methodologies to study the structure and function of carbohydrate chains (‘glycans’) (see Table I). The addition of carbohydrates, ‘glycosylation’, is the most common form of covalent modification of newly synthesized proteins (Kobata, 1992). Glycans, oligosaccharides [composed of 2–10 monosaccharide units (simple sugars)] and polysaccharides (long linear or branched chains of many monosaccharide units), are secondary gene products. It must thus be emphasized that, while the polypeptide backbone sequence of a glycoprotein is directly correlated with the sequence of bases in DNA, the addition and structure of carbohydrate chains is cell-, tissueand species-specific (Kinloch et al., 1991; Kobata, 1992; Lis and Sharon, 1993; Varki, 1993). Biological or functional diversity is generated by the combination of monosaccharides present: by their sequence, chain length, branching points, linkages, anomery (α, β) and/or covalent attachment of modifying groups (sulphate, phosphate, acetyl, methyl). However, extant data suggest that terminal sequences, unusual structures 600 Table I. Advanced methods which have aided in the study of the structure and function of glycoproteins (Adapted from Noguchi et al., 1992; Lis and Sharon, 1993; Varki, 1993; Seppo et al., 1995) Chemical methods Smith degradation acetolysis methylation analysis Physiochemical methods mass spectroscopy 1H-nuclear magnetic resonance (NMR) 13C-NMR Enzymatic methodologies endoglycosylases to release oligosaccharides en bloc monosaccharide sequence analysis by sequential exoglycosidase digestion Non-enzymatic separations gel filtration chromatography reaction with lectins Biological systems mutant mammalian cell lines specific inhibitors of transferases competition studies in-vitro expression of mutant proteins transgenic animals ‘knockout’ animals Analytical techniques for the structural study of oligosaccharide chains compositional analyses in-vitro synthesis competition studies and modifications are most likely to mediate specific biological roles (Varki, 1993). Further, most glycoproteins contain more than one glycan, and glycan heterogeneity at a single attachment site is often observed (e.g. Mori et al., 1991). Intra-species polymorphisms in glycan structure have been noted (Varki, 1993). As a result, a variety of functions can be ascribed to the carbohydrate modifications of proteins. The common features of the varied functions is that they either mediate specific recognition events (e.g. cell–cell; cell–matrix; immune responses; binding of noxious agents; facilitators of infectious and malignant disease) or that they modulate protein function (e.g. ligand binding; intra- and inter-cellular trafficking; stability/clearance/turnover; storage depot for biologically important molecules) (Varki, 1993). Glycans can be linked to proteins through the nitrogen (‘Nlinked’) of asparagine, through the β-hydroxyl groups (‘Olinked’) of serine and threonine and, to lesser extent, of tyrosine, 5-hydroxylysine and 4-hydroxyproline, through a glycosyl-phosphatidylinositol (‘GlPtdIns’) anchor or through a carbohydrate–peptide bond (Table II). In the latter two linkages, the carbohydrate chain is attached respectively, to the carboxyl terminal amino acid of the protein by ethanolamine phosphate or by covalent linkage of 3-hydroxyl of monomeric or polymeric ADP-ribose to a variety of amino acids (‘ADPribosylation’). As these do not represent glycosylation in the strict sense, as a direct glycosidic linkage between protein and carbohydrate is absent, these novel linkages will not be further addressed in this review. Finally, sugars also form links nonenzymatically in vivo to the epsilon amino group of lysine. These adducts further react to form advanced glycation products, which have clinical significance in diabetes. These modifications are also excluded from this review. Carbohydrates and fertilization: an overview Table II. Protein–carbohydrate linkages (adapted from Lis and Sharon, 1993) Amino acid Linkage Carbohydrate residue Asparagine N Serine/Threonine O Serine O Threonine 5-Hydroxylysine 4-Hydroxyproline O O O Tyrosine Carboxy-terminus Amino Acid (Xaa) Glutamic acid Aspartic acid Arginine Cysteine Glucose N-Acetylglucosamine N-Acetylgalactosamine L-Rhamnose Mannose N-Acetylglucosamine N-Acetylgalactosamine Xylose Glucose Galactose L-Fucose Galactose L-Arabinose Galactose Glucose Glycosyl-Phosphatidylinositol O Ethanolamine Phosphate Covalent ADP-ribose (monomeric or polymeric) (3-hydroxyl of Ribose) N-linked glycans N-linked glycans share a common biosynthetic pathway (Kornfeld and Kornfeld, 1985; Elbein, 1991; Goochee et al., 1991). A pre-synthesized oligosaccharide moiety (Glc3Man9GlcNAc2; see abbreviations in legend to Figure 1) is co-translationally transferred from a lipid (dolichyldiphosphate) to the nascent polypeptide chain in the endoplasmic reticulum. Further processing (trimming and addition) in the endoplasmic reticulum and Golgi apparatus generates high mannose (Figure 1A), hybrid or complex (Figure 1B) oligosaccharide structures which contain a common pentasaccharide core (Kobata, 1992). Potential glycosylation sites along a polypeptide backbone for N-linked glycans can be identified from the consensus sequence Asn-Xaa-Ser/Thr (Struck and Lennarz, 1980). Whether or not a particular asparagine residue associated with this consensus sequence is glycosylated is dependent, in part, on protein conformation (Marshall, 1972). The asparagine must be in an exposed position to be available for glycosylation, such as in a β-loop. N-linked glycans modulate the physiochemical properties of the proteins to which they are attached by altering solubility, surface charge and adhesive properties. N-linked glycans can stimulate the immune response by acting as immunodeterminants, as in human blood group antigens (Watkins, 1987), or by activating the complement system (Sim and Malhotra, 1994). In the context of reproduction, it has been well demonstrated that carbohydrate moieties on the sperm surface play an important role in immune infertility (see below), a problem which affects .10% of all couples seeking fertility evaluation (Bronson et al., 1984; Isojima, 1989). Plasma membrane antigens provide potential immunogens for the development of contraceptive vaccines (e.g. Anderson and Alexander, 1983; Liang et al., 1994). Likewise, it is not unexpected that the zona pellucida, a glycosylated structure which is highly antigenic (e.g. Sacco, 1977; Sacco et al., 1986; Yurewicz et al., 1987), is now a target for immunocontraception (Paterson and Aitken, 1990; Epifano and Dean, 1994). In addition, carbohydrates can participate in protection of the embryo from maternal rejection. Glycodelin is a glycoprotein synthesized by the secretory and decidual endometrium, with maximal expression observed at the time of embryo implantation, which manifests immunosuppressive activities in in-vitro assays (Clark et al., 1996b). Glycodelin contains fucosylated glycans (Dell et al., 1995) which block lymphocyte homing (Yednock and Rosen, 1989). It has therefore recently been suggested that the N-linked oligosaccharides of glycodelin, which potentially could compete or block primary adhesive interactions between the embryo and maternal leukocytes, contribute to the highly region-specific immunosuppression of the maternal immune system to the fetus (Clark et al., 1995, 1996a,b). The most important role of N-linked glycans, however, is in the regulation of protein conformation. N-linked glycans aid in folding of nascent polypeptide chains and stabilization of mature glycoproteins. As a result, N- linked glycans modulate the biological activity and immunological properties of a wide variety of enzymes, hormones, receptors and lectins. Three examples are particularly relevant to a discussion of reproductive biology. Firstly, glycan structure regulates the expression of functional human chorionic gonadotrophin (HCG). N-linked glycans facilitate correct disulphide bond formation and thereby assist in the folding of the β-subunit of HCG; the unglycosylated protein folds more slowly and is more easily degraded (Feng et al., 1995). The glycan structure of the α subunit of HCG regulates its ability to interact with the β subunit to form the mature hormone (Blithe, 1990). In addition, N-linked glycans on the α subunit modulate signal transduction, leading to the production of cAMP and steriodogenesis, after receptor binding (Matzuk et al., 1989; Sairam, 1989). The native hormone and its deglycosylated analogue apparently interact with a different domains of the HCG receptor (Ji and Ji, 1990). Secondly, examination of wild-type and mutant rat luteinizing hormone receptor (LHR) expressed in insect cells indicates that N-linked oligosaccharide chains regulate intramolecular folding of the LHR required to achieve high affinity ligand binding (Zhang et al., 1995). Thirdly, N-linked glycans regulate the enzymatic activity of plasminogen activator by modulating the conversion of single chain to double chain forms and also limit their susceptibility to serum and tissue-derived protease inhibitors (Wittwer et al., 1989; Wittwer and Howard, 1990; Howard et al., 1991). Plasminogen activator is a key enzyme in the proteolytic cascade leading to follicular rupture (Canipari and Strickland, 1986; Sappino et al., 1989). O-linked glycans O-linked glycosylation, in contrast to the N-linked process, occurs exclusively in the Golgi apparatus. A preformed, lipidcoupled oligosaccharide precursor does not exist. Instead, the covalent attachment of N-acetylgalactosamine to an acceptor amino acid is through an α1 linkage in the O-linked glycosylation pathway (Figure 1C). Nucleotide sugars serve as substrates for all steps in O-linked glycan processing. The specific polypeptide chain structural requirements for the addition of O-linked oligosaccharides remains to be determined (Sadler, 1984; Carraway and Hull, 1989; Wilson et al., 1991). 601 S.Benoff Figure 1. Typical key saccharide structures*. The arrows in the figure represent glycosidic bonds and the numbers correspond to the type of linkage. (A) N-linked high mannose oligosaccharide (adapted from Elbein, 1991). (B) Consensus sequence of related O-linked and N-linked, complex type oligosaccharides of porcine zona pellucida glycoproteins (adapted from Mori et al., 1991; Hokke et al., 1994). (C) O-linked glycan common to proteins expressed by many mammalian somatic cells (adapted from Goochee et al., 1991). (D) O-linked glycan from mouse ZP3 (adapted from Nagdas et al., 1994). (E) Consensus sequence of blood group-related oligosaccharides that inhibit binding of mouse spermatozoa to mouse ova (adapted from Litscher et al., 1995). *Asn 5 asparagine; Ser 5 serine; Thr 5 threonine; Fuc 5 fucose; Gal 5 galactose; Glc 5 glucose; Man 5 mannose; GlcNAc 5 N-acetylglucosamine; GalNac 5 N-acetylgalactosamine; NeuAc 5 N-acetylneuraminic (sialic) acid. O-linked glycans alone account for the zona’s sperm receptor activity in the mouse (Florman and Wassarman, 1985) and, where information on sugar chain composition is available, other mammals (e.g. Yurewicz et al., 1991). Carbohydrates as cellular recognition determinants Both O-linked and N-linked glycans serve as recognition determinants in targeting proteins to a cellular or tissue compartment (e.g. Dahms et al., 1989), binding of pathogenic organisms to animal cells (e.g. Gambaryan et al., 1995), neural adhesion, including cell interactions during development, modification of synaptic activity and regeneration of nerve connections after damage in the adult (e.g. Schachner and Martini, 602 1995), and leukocyte homing to activated platelets or endothelium (Gahmberg et al., 1992; McEver, 1994). Fertilization is a specialized form of cell–cell recognition. Although the mouse is the best studied system with regard to the role of carbohydrates in fertilization mechanisms (Wassarman, 1989, 1990, 1992; Dean, 1992; Epifano and Dean, 1994), fragmentary data from other mammalian systems is also available. As such, this review will touch upon information gathered from different species as appropriate. Where possible, correlations between saccharide moieties, glycan receptors and outcome of human IVF will be examined in order to help evaluate the precise carbohydrate binding criteria necessary for human fertility. Carbohydrates and fertilization: an overview (Loret de Mola et al., 1996), it will be of interest to determine how differential alterations in these two layers might affect female fertility potential. Thus, polarization microscopy can potentially have an important impact upon clinical practice. In this review, however, the discussion will focus on the outer layer of the zona pellucida as it provides the ligands required for initial sperm binding and induction of the acrosome reaction. Structure of the zona pellucida Figure 2. Hamster oocyte imaged with a polarized light microscope and digital image processing (‘pol-scope’). In the polscope image the zona pellucida is divided into outer and inner layers, each layer displaying birefringence. The orientation of the birefringence axis is indicated by the black lines drawn on a regular grid throughout the image. The measured birefringence axis indicates the orientation of filament arrays which are likely to induce the birefringence in the zona. The slow axis of birefringence (filaments) in the outer layer is oriented tangentially while in the inner layer the slow axis orientation is radial (Keefe et al., 1996). (Photograph generously provided by Dr David Keefe, Brown University School of Medicine, Providence, R.I., and Dr Rudolf Oldenbourg, Marine Biological Laboratory, Woodshole, MA, USA). Zona pellucida Spermatozoa, particularly those of man, do not readily adhere to the oocytes of foreign species (Austin and Bradin, 1956; Bedford, 1977; Yanagimachi, 1977; Oehninger et al., 1993). The zona pellucida, an extracellular glycoprotein matrix surrounding the oocyte, thus provides a barrier to cross fertilization. During evolution, as the oocyte has become smaller, its extracellular vestments have disproportionally increased in size, e.g. compare those of monotreme, marsupial and eutherian mammals (Bedford, 1991). Concomitantly, novel sperm head features have been developed, including decreased enzymic content of the acrosome, so that for the spermatozoa to achieve zona penetration of oocytes of their own species both lysis and thrust (hyperactivated motility) are required. In some limited cases where interspecies sperm–egg interactions have been studied in vitro, however, sperm motility has been sufficient to allow penetration of the zona without prior acrosome exocytosis (Wakayama et al., 1996). Nevertheless, it is safe to assume that such obviation of zona pellucida specificity was largely artefactual as attempts at interspecies breeding are not likely to occur in vivo. The zona pellucida of most nonrodent species can be resolved into a bilaminar structure consisting of an inner (oocyte–proximal) layer and an outer (oocyte–distal) layer, by lectin binding (Shalgi et al., 1991; Maymon et al., 1994) and electron microscopy (Dunbar et al., 1991) and more recently by non-invasive polarization microscopy (Keefe et al., 1996; see Figure 2). As abnormal zona pellucida quality (thickness) may contribute to infertility at the level of embryo implantation (Cohen et al., 1989, 1992), particularly in the older woman The mammalian zona pellucida is composed of three secreted sulphated glycoproteins, zona protein 1 (‘ZP1’), zona protein 2 (‘ZP29) and zona protein 3 (‘ZP39) (Shimizu et al., 1983; Wassarman, 1988), which form heteropolymer filaments of ZP2/ZP3 which are cross-linked by ZP1 (Greve and Wassarman, 1985). Each contains a signal peptide, which directs their secretion and is absent from the amino terminal sequence of the mature molecule, and possesses very little αhelical structure. In addition, each polypeptide contains a highly hydrophobic region at its carboxy terminus which may modulate the formation of the recognizable structural repeat (Liang et al., 1990). The importance of this region is discussed below. Early evidence for conservation of zona pellucida components across species was obtained from the examination of antigenic epitopes (Sacco, 1977; Sacco et al., 1981). Antihuman ovary sera prepared in rabbits cross-reacted with isolated pig zona pellucida. Similarly, the cross-reactivity of anti-pig ovary sera for human zona pellucida was demonstrated by agar-gel immunodiffusion, by formation of a precipitation layer on the zona, and by indirect immunofluorescence cytochemistry. The advent of cDNA technology has permitted the detailed characterization of the structure and sequence of the genes encoding these proteins (Harris et al., 1994). The high degree of homology of their DNA and deduced amino acid sequences in mouse, human, rabbit, cat, dog and pig indicate the existence of three highly conserved gene families. The regions which have been most highly conserved are postulated to be important in maintenance of the three-dimensional structure of the zona (Epifano and Dean, 1994). It is also interesting to note that the amino acid sequences of these proteins are not significantly similar to characterized proteins from somatic cells. However, as mentioned above, both ZP2 and ZP3 have a very hydrophobic region near their respective carboxyl termini (Liang et al., 1990; Liang and Dean, 1993a). The three-dimensional structure of this region (based on the pattern of cysteine, hydrophobic, polar and/or turn-forming residues and not on amino acid sequence conservation) appears to be shared with regions in the extracellular domain of somatic cell proteins such as uromodulin, βglycan [transforming growth factor (TGF)-β type III receptor] and the major zymogen granule membrane protein (Bork and Sander, 1992). It must be noted that the common biological property of the somatic cell membrane-bound proteins which contain the shared domain is that all have been detected in soluble form (Bork and Sander, 1992; Lopez-Casillas et al., 1994). Although the function of this shared structure remains unknown, it has been postulated that it may regulate protein stability or interaction of these 603 S.Benoff proteins in the pericellular environment (Liang and Dean, 1993a; Lopez-Casillas et al., 1994). Using [35S]-methionine, [3H]-fucose, specific antibodies, gel and in-situ hybridization, it has been ascertained that the genes encoding ZP1, ZP2 and ZP3 are transcribed and translated during the 2 week oocyte growth phase, where they represent 7–8% of total protein synthesis (Bleil and Wassarman, 1980b; Salzmann et al., 1983; Ringuette et al., 1986, 1988; Philpott et al., 1987; Roller et al., 1989; Liang et al., 1990). While expression of ZP2 and ZP3 is restricted to the oocyte proper, there is some evidence that ZP1 transcripts may be transiently expressed in granulosa cells (Lee and Dunbar, 1993). Although once secreted, these proteins have a long half-life (.100 h; Bleil and Wassarman, 1980c; Shimizu et al., 1983), zona pellucida protein synthesis is strictly correlated with mRNA levels and both decline in fully grown oocytes. ZP2 and ZP3 transcripts in ovulated eggs are 200 nucleotides shorter than those of growing oocytes, presumably due to deadenylation (Liang et al., 1990). As a result of both low copy number and an inability to translate the shortened transcripts, zona pellucida protein synthesis is not detectable in ovulated eggs. In the mouse, ZP1, ZP2, and ZP3 exhibit apparent molecular weights of 185–200 kDa, 120–140 kDa and 83 kDa respectively (Greve and Wassarman, 1985). As ZP1 is thought to serve primarily as a structural protein and may not contain N-linked oligosaccharide side chains (Tulsiani et al., 1992), glycan modifications have been better studied in ZP2 and ZP3. These polypeptides contain high percentages of proline, serine and threonine residues. Although the sequence for ZP2 encodes seven sites which could direct N-linked glycosylation and more than 100 amino acid residues that could serve as receptors for O-linked glycosylation, the core secreted protein contains six complex-type N-linked glycans and an undetermined number of O-linked glycans (Greve et al., 1982; Liang et al., 1990). ZP3 contains six potential sites for N-linked glycosylation, of which 3–4 are used, and, similar to ZP2, an undetermined number of O-linked sugar side chains attached at .70 potential sites (Salzmann et al., 1983; Harris et al., 1994). Differences detected between species with regard to sizes of homologous zona proteins (e.g. Shabanowitz and O’Rand, 1988a) are attributed to differences in glycosylation. As a result, human zona pellucida glycoproteins appear to be of a lower molecular weight than their mouse counterparts (Liang and Dean, 1993b). In addition, although only a single ZP3 polypeptide is synthesized in mice, at least two ZP3 isoforms are expressed by pig and human oocytes (Hedrick and Wardip, 1986; Yurewicz et al., 1987; Shabanowitz, 1991; Bercegeay et al., 1995). Both N-linked and O-linked oligosaccharides are added prior to secretion by synthetic pathways similar to those utilized in somatic cells (Roller and Wassarman, 1983; Salzmann et al., 1983; Wassarman, 1988). For N-linked glycans, high mannose oligosaccharides added to nascent chains are processed to complex-type moieties in the Golgi. The N-linked glycans of mouse ZP2 and ZP3 have very similar structures (Noguchi and Nakano, 1993). The N-linked oligosaccharides of mouse ZP3 are complex type (bi-, tri-, and tetra-antennary) containing linear N-acetyllactosamine repeats formed on a fucosylated trimannosyl core which may or may not contain significant 604 amounts of sialic acid (Figure 1B; Mori et al., 1991; Tulsiani et al., 1992; Nagdas et al., 1994). In the mouse and in the pig, these are highly heterogeneous in structure (Mori et al., 1991; Yurewicz et al., 1991; Noguchi et al., 1992; Noguchi and Nakano, 1993; Hokke et al., 1994). As these sulphated glycoproteins behave as acidic species (Sacco et al., 1981), Yurewicz et al. (1991) have postulated that their N-linked glycans represent the mammalian analogue of sulphated fucan anionic polysaccharides found in the jelly coat of sea urchin eggs (DeAngelis and Glabe, 1987). Asparagine-linked sugar moieties are, however, not required for secretion (Roller and Wassarman, 1983). Periodic acid Schiff staining and antibody and lectin binding studies suggest that additional modifications in the content and distribution of sugar residues in the secreted proteins occur following ovulation (Sakai et al., 1988; Kaufman et al., 1989; Shalgi et al., 1991). As a result of their extracellular location, ‘structural’, ‘barrier’ and ‘recognition’ functions can be ascribed to the carbohydrates of the zona pellucida. Primary role in sperm–egg interactions A considerable body of evidence indicates that ZP3 serves as the primary sperm receptor regulating tight binding to the zona pellucida. For example, purified mouse ZP3 competitively inhibits binding of mouse spermatozoa to mouse eggs in vitro (Bleil and Wassarman, 1980a; Florman et al., 1984; Florman and Wassarman, 1985; Endo et al., 1987; Leyton and Saling, 1989a). ZP1 and ZP2 lack this inhibitory activity. Mouse spermatozoa bind by their heads to the zona matrix (Drobnis and Katz, 1991). Similarly, binding of purified mouse ZP3 is restricted to the plasma membrane overlying the acrosome (Wassarman et al., 1985b; Bleil and Wassarman, 1986). Only mouse spermatozoa with intact acrosomes have the capability of binding to purified zona pellucida proteins from unfertilized mouse eggs. Consistent with this, binding of purified mouse ZP3 is restricted to acrosome-intact mouse spermatozoa and is not observed on spermatozoa which have undergone an acrosome reaction (Bleil and Wassarman, 1986). As would be expected, ZP3 purified from mouse embryos fails to bind to mouse spermatozoa or inhibit sperm–egg binding in vitro. As a consequence of fertilization, the zona pellucida is modified such that mouse sperm receptor activity is lost (Bleil and Wassarman, 1983, 1988). The link between zona pellucida fine structure and sperm binding capability has been investigated both biochemically and by electron microscopy. Superficial pores very similar in diameter to size of the sperm head have been detected on zona pellucida of unfertilized human oocytes (Nikas et al., 1994); these pores may be essential to fertilization. The anterior portion of the human sperm head has been observed to initially penetrate the zona pellucida through these pores in the zona surface (Tsuiki et al., 1986; Familiari et al., 1988). Treatment of human zonae in vitro which increase the number of these pores is associated with a dramatic increase in sperm binding (Henkel et al., 1995). Fertilization, atresia or artifical egg activation with ionophores results in oocyte cortical granule dehiscence which mediates chemical modifications of the zona (Gulyas, 1980). The proteolytic activity of the cortical granule exudate (Gwatkin et al., 1973) may, in part, be responsible for the loss of sperm binding Carbohydrates and fertilization: an overview activity and the block to polyspermy. In man, these functional changes are correlated with changes in the three-dimensional structure of the zona, from one of numerous elastic ring shaped hoops with superficial pores to an amorphous mass without pores (Familiari et al., 1988; Nikas et al., 1994). It has been postulated that the condensation of the outer layer of the zona pellucida causes disorientation of sperm binding sites (Familiari et al., 1988; Shabanowitz and O’Rand, 1988b). The changes in structure associated with fertilization are more dramatic at the level of the inner zona, and consistent with prior evidence, the cortical granule release block to polyspermy acts at the level of the inner zona in mouse (Sathananthan et al., 1982; Sathananthan and Trounson, 1982, 1985). The migration of ZP3 on sodium dodecyl sulphate (SDS)– polyacrylamide gels is unchanged by the cortical granule reaction (Miller et al., 1992). Loss of ZP3 sperm binding activity and the contribution of ZP3 to the block to polyspermy after fertilization has been attributed to the removal of terminal carbohydrate residues from ZP3-linked oligosaccharide chains by glycosidases in the cortical granule exudate (Miller et al., 1992, 1993). In contrast, the electrophoretic mobilities of ZP1 and ZP2 are dramatically altered following fertilization. In the mouse, ZP2 is modified by limited proteolysis via a proteinase associated with the cortical granules (Bleil et al., 1981; Moller and Wassarman, 1989; Moos et al., 1994; Ducibella et al., 1995). Although preliminary data suggest the human ZP2 may be similarly modified (Moos et al., 1995), a significant body of evidence indicates that species specific changes occur in ZP1 such that ZP1 of human oocytes, but not that of mouse oocytes, is modified by proteolysis or deglycosylation after fertilization (Shabanowitz and O’Rand, 1988a; Bercegeay et al., 1995; Ducibella et al., 1995; Moos et al., 1995). O-linked oligosaccharide chains regulate sperm binding Although the ability of monosaccharides, various complex moieties and lectins to effectively block tight binding of spermatozoa to homologous oocytes in vitro has been examined in a number of mammalian species including man (Oikawa et al., 1974; Huang et al., 1982; Shur and Hall, 1982b; Ahuja, 1985; Shalgi et al., 1986; Mori et al., 1989; Oehninger et al., 1990, 1991), the gamete molecules involved in this binding during fertilization have been best characterized in the mouse. In mice, four lines of evidence strongly suggest that O-linked oligosaccharide ligands on the ZP3 polypeptide backbone act as the primary sperm receptor. Firstly, sperm receptor activity is unaffected by exposure to elevated temperature, reducing agents, denaturants or detergents: treatments which destroy protein-based active sites (Bleil and Wassarman, 1980a; Wassarman et al., 1985b; Wassarman, 1988). Secondly, extensive digestion with pronase similarly fails to affect sperm receptor activity (Florman et al., 1984; Leyton and Saling, 1989a). Thirdly, complete deglycosylation with trifluoromethanesulphonic acid or selective removal of O-linked oligosaccharides following mild alkaline hydrolysis abrogates the ability of ZP3 to inhibit sperm–egg binding (Florman and Wassarman, 1985). The O-linked oligosaccharides released by these procedures were demonstrated to bind directly to spermatozoa and reduce sperm–zona binding. Fourthly, ZP3 treated with endoglycosidase F to selectively remove N-linked glycans and untreated, control ZP3 were equally effective in reducing the number of spermatozoa bound to mouse egg zonae pellucidae in an in-vitro competition binding assay (Florman and Wassarman, 1985). All of the above data were obtained using native ZP3 polypeptides. Recently, recombinant ZP3 (rZP3) has been synthesized in tissue culture cells (Kinloch et al., 1991; Beebe et al., 1992). When expressed in cell lines with the capacity to synthesize a wide variety of oligosaccharides, the ligand activity of rZP3 in sperm–zona pellucida binding competition assays is indistinguishable from that of the native molecule. These arguments are further supported by studies indicating that: (i) human rZP3 must be appropriately glycosylated to be biologically active (Barratt et al., 1994; Van Duin et al., 1994); (ii) O-linked glycosylation is an important step in the processing of human rZP3 (Van Duin et al., 1994); (iii) O-linked oligosaccharide ligands mediate pig sperm adhesion to ZP3 (Yurewicz et al., 1991; Dostalova et al., 1995b); and (iv) deglycosylated pig ZP3 lacks sperm receptor activity and is unable to elicit the production of antibodies which can block sperm–zona binding in vitro (Berger et al., 1989a; Sacco et al., 1989). Evidence has been presented indicating N-linked neutral carbohydrate chains of porcine ZP3 may also contribute to its sperm receptor activity (Noguchi et al., 1992). Competition of sperm–zona binding by such N-linked oligosaccharides is, however, only observed at or above concentrations equivalent to 500 zonae pellucidae. The high concentrations required to effect inhibition of sperm binding suggest that N-linked glycans play a minor role at best in regulating species-specific gamete interactions. In addition, sperm binding activity was retained by porcine ZP3 digested with endo-β-galactosidase, which removes peripheral sulphated polylactosamines of N-linked glycans (Sacco et al., 1989). Further, sperm receptor activity was retained by ZP3 containing truncated O-glycans and by truncated O-glycans liberated by alkaline-borohydride treatment (Sacco et al., 1989; Yurewicz et al., 1991). Similarly truncated (trisaccharide) O-linked chains of mouse ZP3 can inhibit mouse sperm–zona attachment (Nagdas et al., 1994). The latter findings are consistent with the hypothesis that sperm binding activities of the zona pellucida can be associated with internal saccharide core structures (Yurewicz et al., 1991) as well as with terminal residues (Bleil and Wassarman, 1988). This hypothesis is supported by examination of the carbohydrate binding affinities of at least one human sperm surface receptor for zona carbohydrate ligands (see below, sperm receptors for zona carbohydrate ligands). The carbohydrate moieties of ZP3 are also important from a contraceptive standpoint, with the idea of producing reversible interference of sperm–egg interactions (Epifano and Dean, 1994). The antigenicity of ZP3 is directly related to its carbohydrate content (Sacco et al., 1986; Yurewicz et al., 1987). Although the role of carbohydrate moieties as targets for specific antibody production is addressed in the next section, it must be noted that in marmosets, baboons and pigs immununized with zona glycoproteins, glycosylated ZP3 induced immediate production of antibodies whereas only a 605 S.Benoff weak or no response was obtained with deglycosylated ZP3 as immunogen (Paterson and Aitken, 1990; Paterson et al., 1992). Consistent with these findings, immunization of mice and rabbits with their cognate native ZP3 results in autoimmune disease with significant pathological ovarian damage, whereas immunization with deglycosylated ZP3 does not (Keenan et al., 1991; Rhim et al., 1992; Lou et al., 1995; Bagavant et al., 1997). However, although it has been postulated that immunoregulation of gamete function is effected by direct antibody coating of the zona pellucida which inhibits sperm–zona binding and/or penetration (East et al., 1985; Millar et al., 1989; Bagavant et al., 1997), the endocrine changes (elevation of LH and FSH; inhibition of gonadotrophin-induced progesterone secretion; Skinner et al., 1984; Mahi-Brown et al., 1985; Keenan et al., 1991) which occurred in animals immunized with glycosylated ZP3 indicate that inhibition of normal ovarian function also contributes to the production of the infertile state (Keenan et al., 1991). A decrease in the number of primordial follicles, developing follicles and corpora lutea was observed in association with the endocrine changes. When deglycosylated ZP3 was used as the zona antigen, normal endocrine profiles were observed and folliculogenesis was unaffected (Skinner et al., 1984; Mahi-Brown et al., 1985; Keenan et al., 1991). Nevertheless, it must be noted that use of deglycosylated ZP3 as immunogen results in a reduction, and not a complete loss, of fertility (Millar et al., 1989; Lou et al., 1995; Bagavant et al., 1997). In the mouse, oligosaccharide side chains comprise ~50% of the mass of ZP3 (Salzmann et al., 1983). O-linked glycans are found attached to both serine and threonine residues (Florman and Wassarman, 1985). A specific size class of these glycans, ~4 kDa or equivalent to 20–25 monosaccharide residues, inhibit sperm–zona binding and also bind directly to spermatozoa (Florman and Wassarman, 1985), potentially via α-galactose residues at their non-reducing termini (Bleil and Wassarman, 1988). That the sperm receptor activity of these glycans is sensitive to very small changes in carbohydrate structure is indicated by the fact that activity is lost if the terminal galactose residue is removed or if its C-6 hydroxyl is converted to an aldehyde (Bleil and Wassarman, 1988). The amino terminal and carboxyl terminal domains of ZP3 are separated by a flexible (‘hinge’) region similar to that of immunoglobulins (Wassarman and Litscher, 1995). The presence of this hinge region allows proteolytic cleavage of ZP3 (Rosiere and Wassarman, 1992). Separate analysis of the bioactivities of the amino and carboxyl halves of the ZP3 molecule indicate that the carboxyl terminal half of ZP3 is sufficient to inhibit sperm binding to oocytes in vitro and to induce the acrosome reaction (Rosiere and Wassarman, 1992; Litscher and Wassarman, 1996). Complete removal of N-linked oligosaccharides with N-glycanase, release of lactosaminoglycan units from N-linked glycans and removal of sialic acid from non-reducing termini with neuraminidase had no effect on the bioactivities of the carboxyl terminal peptide (Litscher and Wassarman, 1996). Thus, the O-linked glycans responsible for the sperm binding activity of ZP3 are derived from the carboxyl terminal half of ZP3 (Rosiere and Wassarman, 1992). Results from peptide cleavage experiments indicate that the 606 relevant O-linked glycans appear concentrated in a small region of 16 amino acids (328 to 343; exon 7) containing five serine residues. Mapping of the mouse ZP3 combining site for spermatozoa by exon swapping supports these findings (Kinloch et al., 1995). In addition, site directed mutagenesis has produced a mutant gene which converted the five serine residues of exon 7 to non-hydroxyl amino acids. When this mutant gene (mZP3[ser]) is expressed in vitro, its protein product is less extensively glycosylated than wild-type mouse rZP3 and does not exhibit sperm binding or acrosome reaction inducing activities (Kinloch et al., 1995; Liu et al., 1995). The amino acid sequence in and around this region is less highly conserved than other portions of the molecule (Kinloch et al., 1995), supporting the concept that the regions of the ZP3 molecule which potentiate species specific high affinity interactions with spermatozoa are hydrophilic regions lying on the surface of the protein which exhibit maximal difference in sequence between species (Chamberlin and Dean, 1990). Thus, it is not unexpected that the amino acid sequence of mouse ZP3 exon 7 is also highly antigenic. Vaccination of mice with a version of this 16 amino acid peptide synthesized in vitro elicits the production of antibodies that bind to the zona pellucida (Millar et al., 1989). These antibodies produce longterm, reversible contraception in female mice. Although antigenic epitopes in this region of mouse ZP3 can not be detected in hamster, guinea pig, cat or dog ovaries (East et al., 1985), a comparable domain containing O-linked oligosaccharides has been identified in the porcine and human homologues of ZP3 (Yurewicz et al., 1992; Bagavant et al., 1997). Results from the site directed mutagenesis studies described above also provide insight into whether all or only some of the ZP3 molecules in the mouse zona pellucida are inactivated following the cortical granule reaction at fertilization which ablates zonae sperm receptor activity. The zona pellucida of the ovulated mouse egg contains .109 copies of ZP3 (Greve and Wassarman, 1985; Wassarman and Mortillo, 1991). This is in great excess of the estimated 50–500 ZP3 molecules which participate in the initial adhesive event with a spermatozoon (Liu et al., 1995). Thus, it may be postulated that the composition of the zona pellucida can be substantially altered without loss of ability to bind spermatozoa. Although there is a .50% reduction in the number of sperm receptors in the zona pellucida of mice transfected with mZP3[ser], these transgenic mice are fertile and produce litters of normal size (Liu et al., 1995). These findings, when taken together with the fact that embryo ZP3 exhibits a low level of sperm binding activity when tested at concentrations 3–5 times higher than those commonly employed for egg ZP3 (Wassarman et al., 1985a), argue that complete inactivation of a large percentage (.65%) of ZP3 following fertilization will inhibit subsequent sperm–egg binding and polyspermy. Inhibition of ZP3 protein synthesis in mouse oocytes by injection of anti-sense oligonucleotides to degrade ZP3 mRNA (Tong et al., 1995) or by targeted disruption of the ZP3 gene (Liu et al., 1996; Rankin et al., 1996) results in female infertility. Adult ovaries from females homozygous for the disrupted ZP3 gene contain growing and fully grown oocytes. Although ZP1 and ZP2 protein synthesis is unaffected by the anti-sense oligonucle- Carbohydrates and fertilization: an overview otides or disruption of the ZP3 gene, all oocytes lack a zona pellucida. Directed degradation of ZP2 mRNA similarly results in the absence of the zona (Tong et al., 1995). Collectively, these data indicate that assembly of the zona pellucida requires the coordinate synthesis of at least ZP2 and ZP3. However, infertility in female mice homozygous for the targeted ZP3 gene disruption is probably not due to the loss of the extracellular carbohydrates which regulate sperm binding activity. It has been demonstrated that spontaneously acrosome-reacted mouse spermatozoa can fertilize zona-free mouse oocytes in vitro (Naito et al., 1992). Rather, infertility is more likely attributed to adherence of zonafree embryos to the oviductual epithelium which inhibits passage through the oviduct and subsequent implantation (Bronson and McLaren, 1970; Modlinski, 1970). It must be stressed, however, that a naturally occurring genetic defect causing loss of ZP3 expression or absence of the zona pellucida has never been observed to be an underlying cause of mammalian infertility. The latter is consistent with data indicating that genetic defects in glycosylation are rarely observed in intact organisms (Varki, 1993). O-linked glycan structure is complex Analysis of the structure of the O-linked oligosaccharide ligands responsible for sperm receptor activity has been complicated by glycan heterogeneity, which in the hamster occurs on a poly-N-acetyllactosamine backbone (Yurewicz et al., 1991; Hokke et al., 1994). The ability to synthesize in vitro multiply branched oligosaccharide constructs (Seppo et al., 1995) has helped to define structures that inhibit binding of mouse spermatozoa to unfertilized eggs in vitro (Litscher et al., 1995). Constructs tested were rich in N-acetyllactosamine repeats for three reasons. Firstly, multiple polylactosamine units are commonly associated with developmentally regulated glycoconjugates (Fukada et al., 1985; Yurewicz et al., 1987) and mouse zona glycoproteins possess at least two polylactosamine structures per molecule (Cahova and Draber, 1992). Secondly, antibodies to these structures inhibit fertilization (Cahova and Draber, 1992). Thirdly, rZP3 synthesized in mouse embryonal carcinoma cells known to be particularly rich in poly-Nacetyllactosamine glycans (Muramatsu et al., 1978; Renkonen, 1983) exhibited sperm receptor and acrosome reaction inducing activities while that produced in other transfected cell lines with low polylactosamine glycan production did not (Kinloch et al., 1991), Preincubation of capacitated mouse spermatozoa with either blood group β-related linear tri- and penta-saccharides and biantennary blood group I-related oligosaccharides (Figure 1E) resulted in a significant inhibition of sperm–egg binding. In the latter case, increasing chain length also increased the affinity of the branched oligosaccharide for spermatozoa. These data are supported by the direct compositional analysis of Olinked oligosaccharides derived from mouse ZP3 (Nagdas et al., 1994). Treatment of de-N-glycosylated mouse ZP3 with O-glycanase in the presence of exoglycosidases provides evidence for the presence of O-linked trisaccharide chains (Figure 1D), as well as polylactosaminylated glycans (Nagdas et al., 1994). That glycan size and branching pattern play a role in the sperm receptor activity of mouse zona ligands containing terminal α-galactose residues (Litscher et al., 1995) and that O-linked sugar moieties purified from mouse or pig ZP3 are 1–3 orders of magnitude less effective than the native molecule in inhibiting mouse sperm–zona attachment (Yurewicz et al., 1991; Litscher et al., 1995) are consistent with the concept that the nature of the clustering regulates the binding affinity of carbohydrate ligands to their cognate receptor proteins (Lee et al., 1983; Lee, 1989, 1992). Further, the binding to the sperm surface of synthetic glycans which mimic those of blood groups may help to explain why human spermatozoa accumulate seminal blood group antigens on their surface (Cooper and Bronson, 1990; Gilbert et al., 1991; see below, carbohydrates on the sperm surface). Additional evidence that α-galactose-terminated polylactosamines on the mouse zona pellucida regulate species specific tight binding of mouse spermatozoa comes from a study of mouse sperm–erythrocyte interactions (for review see Clark et al., 1996a). Capacitated mouse spermatozoa bind to rabbit erythrocytes but not erythrocytes from other species. An intact acrosome is a requirement for rabbit erythrocyte binding. Pretreatment of rabbit erythrocytes with α-galactosidase reduced mouse sperm binding by a maximum of 40% in comparison with untreated controls. The residual 60% binding of mouse spermatozoa to α-galactosidase-treated rabbit erythrocytes has been attributed to an interaction with internal polylactosamine sequences. On rabbit erythrocytes, α-galactosyl residues are primarily associated with branched polylactosamines. Finally, it must be recognized that the identification of the specific zona carbohydrate ligands regulating tight binding of spermatozoa is complicated by the fact that the glycosyl modifications presumably exhibit intra-individual variation. Oocytes from the same donor exhibit variations in their sperm binding capacity (e.g. Fazeli et al., 1993). Species-specific glycosylation patterns The size and structural features of the active oligosaccharide site of ZP3 which regulates sperm binding appears to differ between species (Noguchi et al., 1992). For example, while terminal carbohydrate residues apparently form the sperm receptor site in the mouse zona (Bleil and Wassarman, 1988; Miller et al., 1992, 1993), larger internal carbohydrate units regulate sperm binding to pig zonae (Yurewicz et al., 1991). It is therefore interesting to note that mouse embryonal carcinoma cells which constitutively express mouse or hamster ZP3 genes can discriminate between these highly related polypeptides (Kinloch et al., 1991). Although these cells secrete a biologically active form of mouse rZP3 which is glycosylated to an extent similar to that of the native mouse ZP3, the mature hamster srZP3 failed to exhibit mouse sperm receptor activity. That hamster rZP3 secreted by mouse cells fails to inhibit hamster sperm binding to ovulated eggs in vitro may be due to: (i) the fact that it is glycosylated to a lesser extent than the authentic protein; (ii) the possibility that the oligosaccharide structure of the recombinant molecule differs from that of hamster oocytes, thus altering the sperm binding site; and/or (iii) the possibility that the oligosaccharide structure of the recombinant molecule differs from that of hamster 607 S.Benoff oocytes such that protein folding is altered. In addition, the lectin binding affinities of zonae pellucidae differ between species (Shalgi et al., 1991; Maymon et al., 1994). All of these arguments support the concept of species-specific glycosylation patterns for ZP3. The majority of evidence that glycosyl modification of ZP3 is species specific will, however, be discussed in detail below in regard to the sperm molecules participating in zona adhesion (see below, sperm receptors for zona carbohydrate ligands). Carbohydrates on the sperm surface The surface of the mature mammalian spermatozoon can be divided into five domains [anterior head (acrosomal segment), equatorial segment, posterior head (post-acrosomal segment), midpiece, and principal piece] that originate during elongation of round cells and are conserved across species (for review see Bearer and Friend, 1990). These structural domains are thought to reflect the presence of ‘functional domains’ involved in the process of fertilization. Studies employing specific saccharides in hapten inhibition tests, exoglycosidases to remove functional saccharides from the cell surface and inhibitors of glycoprotein synthesis all suggest that a large variety of oligosaccharides synthesized by the lipid-linked pathway (N-linkage) are expressed on the sperm surface and serve as carbohydrate determinants involved in mammalian fertilization (review, Ahuja, 1985). Therefore, although glycoproteins are involved in intercellular adhesion and information transfer during spermatogenesis (Lee and Damjanov, 1985; Malmi et al., 1987; Wollina et al., 1989a,b), the majority of studies have focused on the role of surface carbohydrates in the evolution of membrane domains during epididymal maturation and capacitation post-ejaculation. Distribution Fluorescent dye-, peroxidase-, biotin-, ferritin-, colloidal gold-, and radio-labelled phytoagglutinins (plant ‘lectins’) have been used as probes to study carbohydrates on the sperm surface (Edelman and Millette, 1971; Nicolson and Yanagimachi, 1972, 1974; Nicolson et al., 1972; Nicolson, 1974; Koehler, 1981; Kallojoki et al., 1985; Singer et al., 1985; Cross and Overstreet, 1987; Kumar et al., 1990; Ahluwalia et al., 1990; Vasquez et al., 1990; Bains et al., 1992; Mladenovic et al., 1993; Gabriel et al., 1994; Fierro et al., 1996). Surface charge differences, potentially attributable to the negative charge associated with terminal sialic acid residues, have been assessed with colloidal iron hydroxide or by electrophoresis of immobilized live spermatozoa (Nevo et al., 1961; Bedford, 1963; Cooper and Bedford, 1971; Yanagimachi et al., 1972). To date, the results of these studies indicate that a wide variety of saccharide groups are accessible on the surface plasma membrane, both on the termini and within the cores of polysaccharide complexes. Their distribution is non-uniform, e.g. concanavalin A (‘ConA’; specific for D-mannose and Dglucose residues) binding is limited to the region overlying the acrosome in human, mouse, hamster, guinea pig, rat, rabbit and ram spermatozoa. Significant differences exist between species with regard to 608 lectin binding affinities. Most important are observations indicating reproducible changes in lectin binding affinities during epididymal maturation and capacitation, e.g. ConA labelling of the sperm head is decreased while that of wheat germ agglutinin (‘WGA’; specific for N-acetylglucosamine and/or sialic acid) is increased. Some of these changes are correlated with the acquisition of the ability to bind zona pellucida ligands (Lassalle and Testart, 1994). The overall loss of lectin binding sites from the periacrosomal plasma membrane correlates with changes in the distribution of intramembranous particles prior to the acrosome reaction first described by freeze fracture/electron microscopy (Friend et al., 1977; Yanagimachi and Suzuki, 1985; Aguas and Pinto da Silva, 1989; Suzuki and Yanagimachi, 1989). This allows calcium to induce phase transitions which create disorganized lipid arrays necessary for the initiation of fusion (Papahadjopoulos, 1978) of plasma and outer acrosomal membranes at particle-deficient interfaces leading to acrosome exocytosis. Membrane contraints restrict diffusion of bound lectins in the anterior head and the tail regions but not in the posterior head. This is indicative of sperm membrane compartments where lipid is immobile and is not free to diffuse laterally as in most somatic cells (Wolf et al., 1988), and is consistent with observations indicating that alterations in such lipid domains are associated with decreased fertility potential of human spermatozoa (Sinha et al., 1994). The best studied lipid diffusion barrier is that in the equatorial segment of the spermatozoon (Arts et al., 1994); as (i) this is the primary domain of mammalian acrosome-reacted spermatozoa that harbour in vitro fusogenic capacity for artificial membranes (Arts et al., 1993) and (ii) it is this domain which initiates fusion of the spermatozoa with the oolemma (Bedford et al., 1979; Talbot and Chacon, 1982; Clark and Koehler, 1990; Yanagimachi, 1994). It is unlikely that transmembrane protein barriers to lipid diffusion exist in this region. Intramembranous particles are cleared from the equatorial segment after the acrosome reaction (Friend et al., 1977; Aguas and Pinto da Silva, 1989). The stability of the equatorial segment barrier to lipid diffusion appears to be dependent on the presence of calcium ions (Arts et al., 1994). Annexins have been detected in the equatorial segment of human spermatozoa (Berruti, 1988; Feinberg et al., 1991). Annexins, which require calcium for their functioning, have been implicated in the attachment of cytoskeletal elements to membranes (Creutz, 1992). Thus, the possibility that annexins contribute to the barrier to lipid diffusion in the equatorial segment cannot be excluded. Calcium also regulates the physical state of lipids. Although calcium can trigger some lipids to partially ‘melt’ (e.g. become more fluid), it can also trigger the transition of other lipids from a liquid-crystalline to a gel state (‘freezing’) (Lee, 1977; Papahadjopoulos, 1978; Cullis and DeKruijff, 1979). Lipid factors, such as the sperm-specific glycolipid sulphatoxygalactosyl-glycerolipid localized to the equatorial segment, can form gel-like areas in sperm plasma membranes where molecular motion is very slow (Wolf et al., 1988; Benoff et al., 1993a; Gadella, 1993). Regulation of the formation of these gel Carbohydrates and fertilization: an overview domains by calcium offers a plausible alternative to annexins as an equatorial diffusion barrier. In addition, the existence of glycolipids associated with the sperm plasma membrane suggests that glycoprotein turnover and remodelling may not be the exclusive agents regulating sperm surface lectin binding affinities during epididymal transit and sperm capacitation. Sperm plasma membrane lipid/sterol ratios change markedly during capacitating incubations and the extent of spontaneous and induced acrosome reactions is critically dependent on these ratios (e.g. Go and Wolf, 1983; Hoshi et al., 1990; Benoff, 1993; Benoff et al., 1993a,c,d, 1996a). The changes that occur in membrane composition are specific to a particular domain of the mature spermatozoa (Bearer and Friend, 1982). Thus, modification of sulphatoxygalactosylglycerolipid and similar glycosylated lipids may have a biological role in the regulation of fertilization processes occurring after induction of the acrosome reaction. Type of association In the study of sperm surface glycans, a distinction needs to be made between those which are part of integral plasma membrane proteins and those which are derived from proteins and lipids which are only peripherally attached. There are additions to the sperm surface after testicular differentiation (review, Yanagimachi, 1994). Spermatozoa dramatically alter their antigenicity during epididymal transit for two reasons. Firstly, the epididymis secretes a variety of glycoproteins, some of which bind tightly to spermatozoa. Secondly, active glycosylation of membrane-integrated components has also been noted and is thought to be mediated by galactosyltransferase and sialyltransferase in epididymal fluid. Thus, the storage function of the epididymis may be responsible for the need for capacitation in order to undergo the zona pellucida-induced acrosome reaction and to fertilize metaphase II oocytes in vitro (Yanagimachi, 1994; Visconti et al., 1995a). However, as epididymal modulation of expression of adhesion molecules, such as β-integrin and fibronectin, has been observed (Glander et al., 1994), the glycoproteins added in the epididymis are thought to help stabilize the sperm plasma membrane and prevent premature capacitation or acrosome reactions. This is particularly important because spermatozoa differ significantly from somatic secretory cells which are able to undergo repeated cycles of exocytosis. The acrosome reaction can occur but once and the lifetime of an acrosome-reacted spermatozoa is extremely limited (Drisdel et al., 1995). Seminal plasma from many mammalian species also exerts antifertility activities (Robertson et al., 1971; Eng et al., 1978; Reddy et al., 1978; Fraser, 1984; Cross, 1996). Glycoproteins in seminal plasma can readily bind to the sperm plasma membrane (e.g. Han et al., 1990; Drisdel et al., 1995). This may help explain why ejaculated spermatozoa from many species are more difficult to capacitate than spermatozoa from the cauda epididymis (Yanagimachi, 1994). At least one of these seminal plasma glycoproteins has been subjected to a detailed characterization (Drisdel et al., 1995). A 74 kDa glycoprotein from human seminal plasma specifically blocks the induction of the human sperm acrosome reaction by a variety of agonists. The 74 kDa glycoprotein contains both complex N-linked and O-linked oligosaccharide chains with high concentrations of galactose, galactosamine, glucosamine and mannose. Biological activity is lost if either the Nlinked or O-linked oligosaccharides are removed. The 74 kDa protein is either inactivated or detached from the sperm surface during capacitation. These data suggest that interaction of this glycoprotein with human spermatozoa is mediated by carbohydrate-binding proteins on the surface plasma membrane. Further, the carbohydrate composition of the glycans associated with the 74 kDa protein suggests that sperm surface proteins which bind this glycoprotein may also be involved in zona pellucida binding (see below, sperm receptors for zona carbohydrate ligands). The peripherally attached glycocalyx is removed during capacitation in the female tract. Sperm-coating blood groupspecific carbohydrates can be similarly removed in vitro from ejaculated spermatozoa by Percoll density gradient centrifugation (Gilbert et al., 1991). Surface glycans, in the form of glycosaminoglycans (GAGs), absorbed from uterine, oviductal and follicular fluid, in contrast to those from epididymal fluid, promote sperm capacitation and the acrosome reaction, possibly by increasing intracellular pH and calcium ion uptake (review, Miller and Ax, 1990). The GAG content of these fluids is highest at oestrus (Parrish et al., 1989), consistent with prior observations on oestrus cycle dependency of capacitation (Viriyapanich and Bedford, 1981). The ability of GAGs to promote sperm fertilizing potential is correlated with their content of sulphate and amino sugars, with heparin having relatively high potency. Radiolabelled heparin is disproportionately bound to the head of human and other mammalian spermatozoa in comparison with other membrane compartments. At least nine proteins isolated from bovine sperm plasma membranes are capable of binding heparin. The most important of these may be fibronectin and fibronectin-related peptides. Fibronectin contains the Arg-Gly-Asp sequence implicated as a ligand in sperm adhesion to oolemmal integrins (Bronson and Fusi, 1990; Fusi and Bronson, 1992; Schaller et al., 1993; Fusi et al., 1992, 1993). Anti-fibronectin antibodies inhibit binding and penetration of zona-free hamster oocytes by human and hamster spermatozoa. Fibronectin contains two heparin-binding domains (for review see Hynes, 1985). Fibronectin has been found overlying the head region of human spermatozoa, and expression of fibronectin in the equatorial segment is specifically correlated with normal sperm morphology (Vuento et al., 1984; Glander et al., 1987). The percentage of human spermatozoa exhibiting surface binding of antifibronectin antibodies is increased in association with capacitation (Fusi and Bronson, 1992). That neither the proportion of spermatozoa exhibiting anti-fibronectin antibody reactivity nor the distribution of fibronectin on spermatozoa is altered by induction of the acrosome reaction with progesterone (Fusi and Bronson, 1992) suggests that fibronectin is located on the surface plasma membrane, not the inner acrosomal membrane. This localization would be consistent with its ability to react with GAGs during capacitation. Detection The most convenient method for detecting sperm glycans is that of lectin binding. Results from lectin binding studies must, 609 S.Benoff however, be interpreted with caution for two reasons. Firstly, lectin specificities are dependent on technical conditions such as fixation, temperature and concentrations of ligands and/or competing sugars (Kochibe and Furukawa, 1980; Gallagher, 1984; Hindsgaul and Ballou, 1984; Sato and Muramatsu, 1985; Yamashita et al., 1985, 1987, 1988). Secondly, the method of sperm preparation greatly influences analytical outcome, e.g. whether or not internal carbohydrate residues are exposed by drying and/or fixation (Bronson et al., 1984; Haas et al., 1988; Bains et al., 1992). Nevertheless, glycan expression in extracellular and intracellular compartments, as determined by lectin binding, can be used to predict the functional state of human spermatozoa. Glycan expression may be used as a marker for assessment of sperm morphology (Gabriel et al., 1994), i.e. by identifying spermatozoa with normal oval head morphology and an intact acrosome, requirements for binding and penetration through the human zona pellucida (Liu and Baker, 1994). Lectin binding properties can also distinguish subpopulations within normospermic semen (Singer et al., 1986). Lectin binding affinities have been observed to be altered in spermatozoa from infertility patients (Vasquez et al., 1990; Bains et al., 1992). For example, decreased expression of WGA receptors has been suggested as one cause of human male infertility (Fei and Hao, 1990; Zhixing and Yifei, 1990). N-acetylglucosamine and/or sialic acid receptors for WGA localized to the equatorial segment and posterior head of normal human spermatozoa have been postulated to play an important role during fertilization (Cerezo et al., 1982; Kallojoki et al., 1985; Franken et al., 1990). The intensity of sperm surface WGA labelling is high in normal spermatozoa and decreases with increasing degree of teratozoospermia (Gabriel et al., 1994), suggesting that either WGA receptors play a quasi-structural role in sperm surface organization or that the boundaries of diffusible/nondiffusible lipid domains change with morphology. Peanut agglutinin (‘PNA’) is specific for β-galactosyl residues whereas ConA and Pisum sativum agglutinin (‘PSA’) are specific for D-mannosyl and D-glucosyl residues. These lectins can be used to differentiate between human sperm head compartments. PNA binds to the plasma membrane/outer acrosomal membrane (Mortimer et al., 1990; Bains et al., 1992), PSA to acrosome content (Cross et al., 1986) and ConA to the inner acrosomal membrane (Holden et al., 1990). Thus, these lectins can be used to assess differences between control and patient populations with regard to the ability to undergo induced acrosome loss (Mladenovic et al., 1993). That lectins with similar monosaccharide specificities, e.g. PSA and ConA, label different sperm compartments emphasizes the diversity of sperm glycan structure. Lectin binding is not merely the result of a stoichiometric relationship to respective monosaccharides, but also reflects the secondary and tertiary structure of the glycan as a whole (e.g. Lee et al., 1983, 1984; Kery et al., 1992; Lee, 1992). Immune infertility The most important role ascribed to sperm surface carbohydrate moieties is that of antigens stimulating the production of naturally occurring anti-sperm antibodies (ASAs). Cell surface 610 carbohydrates are highly antigenic (Sela et al., 1975). Use of monoclonal antibodies raised against human spermatozoa demonstrated that different surface antigens are confined to regional domains similar to those originally described by electron microscopy (Villarroya and Scholler, 1984, 1986). As above, however, surface coat antigens which are peripherally associated with the sperm membrane must be distinguished from those linked to integral plasma membrane proteins. Freshly ejaculated and capacitated human spermatozoa have been reported to react with different types of ASAs (Fusi and Bronson, 1990; Monroe et al., 1990; Margalioth et al., 1992). The alterations in antigenicity associated with capacitation are consistent with changes in sperm surface lectin binding affinities described above and may reflect surface modifications associated with changes in functional state of spermatozoa. Loss of peripherally associated surface coating molecules, immunoglobulin (Ig) M isotypic antibody reactive blood group antigens and non-blood group seminal antigens reacting with IgG isotypic antibodies, may to a large part, account for changes in sperm antibody test results when spermatozoa are subjected to capacitating conditions (Cooper and Bronson, 1990; Gilbert et al., 1991). Antibodies to blood group antigens do not appear to influence sperm function. In contrast, ASAs directed against intrinsic sperm surface antigens originating in the testis can dramatically reduce the potential for gamete interaction (Bronson et al., 1984; Esponda and Bedford, 1985; Patrizio et al., 1992). Conception rates are markedly reduced in couples where immunity to spermatozoa is detected (Ayvaliotis et al., 1985; Busacca et al., 1989). Only a limited number of glycans on lipids (Kurpisz et al., 1989) and proteins (Primakoff et al., 1990; Snow and Ball, 1992; D’Cruz et al., 1993; Benoff et al., 1995b) are immunodominant. Their distribution on the sperm surface correlates with effects of ASAs on sperm function. Results from IVF indicate that sperm head-directed antibodies may prevent or reduce fertilization, whereas tail-directed antibodies do not (e.g. Zouari et al., 1993). Such data are consistent with in-vitro studies following passive transfer of anti-sperm immunoglobulins to antibody-free spermatozoa of fertile donors: tail-directed ASAs are associated with motility loss in the presence of complement while head-directed ASAs primarily impede zona pellucida binding or subsequent acrosome loss (e.g. Bronson et al., 1982a,b; Lansford et al., 1990). The anti-fertility effect of head-directed ASAs is produced by at least two different mechanisms: (i) inhibition of membrane cholesterol efflux, an important part of the capacitation process (Benoff et al., 1993a; see Figure 3); and (ii) inhibition of progesterone binding to the non-nuclear progesterone receptor (Benoff et al., 1995b). Evidence has been presented that immune infertility is the result of local inflammation which augments natural anticarbohydrate immunities against infectious agents (Witkin and Toth, 1983; Cooper et al., 1991; Kurpisz and Alexander, 1995). The same simple sugars terminate bacterial and sperm surface antigens (Tsang et al., 1987; Cooper et al., 1991; Kurpisz and Alexander, 1995). Antisperm antibodies from human males cross-react with bacteria (Tung, 1975). Autoimmunity to spermatozoa is observed following chronic prostatitis (Fjallbrant and Nilsson, 1977). Isoimmunity to spermatozoa Carbohydrates and fertilization: an overview often develops in women following genital tract infections (Cunningham et al., 1991). Although no information is available concerning the nature of the carbohydrate structures recognized by head-directed ASAs, those recognized by ASAs participating in complementmediated sperm immobilization have been partially characterized using anti-sperm human monoclonal antibodies (Tsuji et al., 1988; Isojima, 1989). The internally repetitive, unbranched N-acetyllactosamine (heteropolymers of galactose and N-acetylglucosamine; blood type I antigen) and sialyl branched internally repetitive N-acetyllactosamine (sialyl blood type I antigen) on the surface of ejaculated human spermatozoa are reactive with this class of ASAs. The ASAs recognize an internal sequence of the glycans, as terminal substitutions of galactose with sialic acid have no effect on monoclonal antibody binding. Nevertheless, sialylation levels appear critical to overall antibody recognition. If removed, other ASAs will bind to the glycan. Biological role in reproduction The findings described above indicate that the sperm surface glycans serve mainly as ‘traitors’ in generating an immune response and an infertile state. Sperm surface glycans and their remodeling during capacitation also serve in a ‘protective’ or ‘stabilizing’ role, e.g. to prevent premature loss of acrosome content. It is therefore unlikely that oligosaccharides on the sperm surface directly participate in species-specific tight binding to the zona. Thus, sperm surface glycans are of lesser interest than are sperm plasma membrane receptors for zona pellucida glycoproteins. Figure 3. Comparison of the sugar ligand binding affinities of human, cynomologus macaque and rabbit spermatozoa. Ligand binding was performed using the calcium-supplemented buffer system previously described (Benoff et al., 1993b–d). Data are presented as mean 6 SEM for fresh isolates and duplicate aliquots subjected to capacitating incubations. (A) The sugar ligand binding affinities of motile spermatozoa populations from fertile donors (n 5 14) were assessed at time of swim-up and after overnight incubation at 37°C in 5% CO2 in air in Ham’s F-10 supplemented with 30 mg/ml delipidated human serum albumin. (B) Cynomologus macaque spermatozoa (n 5 6) was obtained by rectal electroejaculation under ketamine anaesthesia by Dr Mary C.Mahony (Jones Institute, Norfolk, VA, USA). Each ejaculate was collected into TALP–Hepes medium, diluted to a concentration of ~53106/ml and transported overnight to NSUH at ambient temperature. Surface sugar ligand receptor affinities were measured in aliquots of untreated spermatozoa and in duplicate aliquots after activation for 30 min at 37°C in 5% CO2 in air in TALP medium supplemented with 1 mM each of caffeine and dibutyrl cyclic AMP (conditions for monkey IVF; VandeVoort et al., 1992). (C) Rabbit spermatozoa (n 5 6) were collected using an artificial vagina. X-FITC-LIGAND binding was assessed at time of collection and after overnight capacitation in high salt media (Breitenstein and Schrader, 1996). (Data generously provided by Dr Steven M.Schrader and Michael Breitenstein, NIOSH, Cincinnati, OH, USA). α-MAN 5 α-D-mannosylated serum albumin (Sigma No. A7790); α-GAL 5 α-D-galactosylated serum albumin (Sigma No. A2420); β-GAL 5 β-D-galactosylated serum albumin (Sigma No. A8165); α-FUC 5 α-L-fucosylated serum albumin (Sigma No. A5793); α-GLU 5 α-D-glucosylated serum albumin (Sigma No. A5543); β-GLU 5 β-D-glucosylated serum albumin (Sigma No. A7172); β-LAC 5 β-D-lactosylated serum albumin (Sigma No. A8040). FITC 5 fluorescein isothiocyanate. Spermatozoa receptors for zona carbohydrate ligands The conserved structure and simple composition of the mammalian zona pellucida has facilitated identification of surface components which interact with spermatozoa. In contrast, it has been postulated that the sperm receptors for the zona pellucida have not been conserved and that, in fact, such variation drives speciation (Snell and White, 1996). However, the complexities of the patterns of expression of proteins in the sperm plasma membrane (Gillis et al., 1978; Russell et al., 1983; Mack et al., 1987; Naaby-Hansen, 1990; Xu et al., 1994), the regional specificities of sperm protein distributions (Peterson et al., 1987; Bearer and Friend, 1990) and difficulties in isolating plasma membrane fractions free of cytoplasmic contamination (e.g. Snow and Ball, 1992) have made the identification of components which bind to zona glycoproteins a daunting task. Even where abnormal sperm plasma membrane protein expression has been correlated with human male infertility (e.g. Morgentaler et al., 1990), the large number of proteins to be studied has precluded identification of protein function. At least three different sperm head membrane compartments interact with unfertilized eggs at different stages of the fertilization process in mice (Mortillo and Wassarman, 1991). In the mouse, competition analysis has demonstrated selective binding of acrosome-intact spermatozoa to the zona in the 611 S.Benoff Table III. Sperm surface proteins with affinity for the extracellular matrix of the egg* Species Sperm receptor Mouse p95 Rat Boar Guinea pig Rabbit Human Reference Leyton et al. (1992); Kalab et al. (1994) β-1,4-galactosyl transferase Miller et al. (1992) trypsin-like enzyme Saling, 1981; Benau and Storey (1987) α-D-mannosidase Cornwall et al. (1991) SLIP1 Tanphaichitr et al. (1993) 15 kDa protein Poirier et al. (1986); Aarons et al. (1991) sp56 Cheng et al. (1994) α-D-mannosidase Tulsiani et al. (1995) galactose receptor Abdullah et al. (1991); Mertz et al. (1995) APz Peterson and Hunt (1989); Peterson et al. (1991) Zonadhesin (p105/45) Hardy and Garbers (1994, 1995) AWN-1 (spermadhesin) Calvete et al. (1995); Dostalova et al. (1995b) L-fucose receptor Huang et al. (1982) PH-20 Primakoff et al. (1985) RSA (54 kDa) Abdullah et al. (1991) RSA (Sp17) O’Rand et al. (1988); Richardson et al. (1994) mannose receptor Benoff et al. (1993b,c, 1995a) α-D-mannosidase Tulsiani et al. (1990) hu9 Burks et al. (1995) β-1,4-galactosyl transferase Humphreys-Beher and Blackwell (1989); Humphreys-Beher et al. (1990) trypsin-like enzyme Pillai and Meizel (1991); Llanos et al. (1993) 15 kDa prtoein Boettger-Tong et al. (1993) galactose lectin Goluboff et al. (1995) selectin Lucas et al. (1995); Clark et al. (1995) FA-1 Naz and Ahmad (1994) *Although every attempt was made to ensure that the numerous alternative candidates have been included in this review, this table may still be incomplete. SLIP 5 sulphoglycolipid immobilizing protein; AP 5 adhesion protein; RSA 5 rabbit sperm antigen; FA 5 fertilization antigen; PH 5 posterior head antigen. presence of an excess of acrosome-reacted spermatozoa (Saling and Storey, 1979). Electron microscopy has demonstrated that mouse spermatozoa first establish contact with the zona pellucida via the region of the head plasma membrane overlying the acrosome (Wassarman, 1990; Drobnis and Katz, 1991). These findings have been confirmed by binding of radioiodinated or colloidal gold-labelled ZP3 to mouse spermatozoa (Bleil and Wassarman, 1986; Mortillo and Wassarman, 1991). Autoradiographic grains were specifically restricted to the sperm head; none were observed along the midpiece or tail principal piece. Further, it was established that ZP3 bound to acrosome-intact, but not acrosome-reacted spermatozoa. Taken together, these data indicate that the molecules distributed over the periacrosomal plasma membrane were responsible for the initial tight binding of spermatozoa to zona. A large body of data indicates that this group of proteins exhibits speciesrestricted expression (Table III; Figure 3). Spermatozoa that undergo an acrosome reaction after zona binding remain bound to oocytes, as a result of secondary binding to ZP2 (Bleil and Wassarman, 1983, 1986; Bleil et al., 1988). As the acrosome 612 reaction results in the exposure of a new membrane compartment, it is argued that molecules responsible for maintenance of binding of acrosome-reacted spermatozoa to oocytes are located on the inner acrosomal membrane. Examination of the distribution of colloidal gold-labelled ZP2 binding to acrosomeintact and acrosome-reacted mouse spermatozoa by transmission electron microscopy supports this argument (Mortillo and Wassarman, 1991). Acrosin is the primary candidate for a ZP2 binding protein (Jones, 1989, 1991). Other acrosomal antigens may also be involved, e.g. HSA-63 (Liu et al., 1992) and SP10 (Wright et al., 1990). Preliminary findings suggest that secondary binding to ZP2 is not species limited as [125I]labelled hamster zona glycoproteins bind to sperm proacrosin from boar, ram, bull, rat and mouse (Williams and Jones, 1993) and as the fucoidin ligand for acrosin blocks sperm– zona binding in a variety of species (Yanagimachi, 1994). In addition, the acrosome reaction also generates a restricted and modified segment of the sperm plasma membrane overlying the equatorial region which is capable of fusion with the oolemma (Moore and Bedford, 1978; Bedford et al., 1979; Yanagimachi, 1994). At least three sperm surface molecules in this final membrane compartment have been postulated to participate in oolemmal binding and gamete fusion: vitronectin/ fibronectin (Bronson and Fusi, 1990; Fusi and Bronson, 1992; Schaller et al., 1993), β integrin (Schaller et al., 1993; Klenteris et al., 1995) and PH-30/‘fertilin’ (Blobel et al., 1992; Myles et al., 1994). In this review, the discussion will concentrate on the zona binding proteins expressed on the surface of acrosomeintact spermatozoa. Six major approaches have been employed in the identification of sperm surface components which exhibit zona binding activity: (i) immobilized, size-separated sperm proteins were probed with radiolabelled or biotinylated zona glycoproteins (e.g. O’Rand et al., 1985; Topfer-Petersen and Henschen, 1987; Jones, 1989); (ii) sperm membrane components obtained following detergent solubilization were fractionated by a variety of chromatographic techniques and then tested for their ability to block sperm–zona binding (e.g. Peterson et al., 1983; Jonakova et al., 1991); (iii) affinity chromatography of detergent-extracted spermatozoa by passage through columns of zona glycoproteins bound to a solid support (e.g. Hanqing et al., 1991; Hardy and Garbers, 1994); (iv) specific polyclonal and monoclonal antisera which inhibit sperm–zona binding (e.g. Peterson and Hunt, 1989); (v) cross-linking of sperm surface proteins to zona ligands (O’Rand et al., 1988; Bleil and Wassarman, 1990); and in man, (vi) correlation with the infertile state (e.g. Benoff et al., 1993b,c). The individual molecules which have been identified by these techniques have been divided by species and described in detail below. Mouse Five integral plasma membrane components [p95, β-1,4galactosyl transferase, a trypsin-like proteinase, α-D-mannosidase and sulphoglycolipid immobilizing protein (SLIP)1] and two peripherally associated proteins (15 kDa and sp56) have been implicated in zona recognition and tight binding by mouse spermatozoa. Carbohydrates and fertilization: an overview p95 A 95 kDa protein, ‘p959, has been identified on the head surface of motile, acrosome-intact mouse spermatozoa, which binds to whole mouse zonae or purified mouse ZP3 and also serves as a tyrosine kinase substrate (Leyton and Saling, 1989b). The anti-phosphotyrosine antibody reactivity of this protein is increased by interaction with zona glycoproteins (Leyton and Saling, 1989b). Oligomerization of p95 by immunoaggregation induces the acrosome reaction (Leyton et al., 1995). p95 is an autophosphorylating protein (Leyton et al., 1992). Spermatozoa express voltage-dependent calcium channels (Florman, 1994; Goodwin et al., 1997; see Figure 10) and tyrosine phosphorylation regulates ion fluxes through voltage-dependent calcium channels (Sieglebaum, 1994; A.Jacob, L.O.Goodwin and S.Benoff, unpublished observations). As inhibition of tyrosine kinase activity inhibits acrosome exocytosis (Leyton et al., 1992), p95 has been identified as a sperm surface protein which is aggregated by ZP3 (Leyton and Saling, 1989b; Leyton et al., 1995). There are conflicting reports as to whether p95 tyrosine phosphorylation is increased during mouse sperm capacitation (Leyton and Saling, 1989b; Ducan and Fraser, 1993; Visconti et al., 1995a,b). Whether or not the capacitation associated increase in the phosphotyrosine content of p95 is regulated by a cyclic AMP/protein kinase A signalling pathway is similarly debated (Duncan and Fraser, 1993; Visconti et al., 1995b). Leyton et al. (1995) provide no sequence data for the p95 mouse sperm antigen that reacts with anti-phosphotyrosine antibodies. Therefore, there are conflicting reports in the literature as to whether p95 is in fact a unique tyrosinephosphorylated form of hexokinase (Kalab et al., 1994). This leaves open the possibility that there are two phosphotyrosinecontaining proteins that participate in the initial events of mouse gamete binding. It is interesting to note that four different isoforms of p95/ hexokinase are transcribed in mouse male germ cells and spermatozoa (Visconti et al., 1996). At least one of these isoforms is an integral membrane protein present on the sperm head and has an extracellular domain. This isoform becomes phosphorylated on tyrosine residues during spermatogenesis. This protein is not phosphorylated in spermatozoa from mice carrying two t haplotypes (Olds-Clarke et al., 1996). Spermatozoa from such mice are also unable to penetrate the mouse zona pellucida (Johnson et al., 1995). These recent findings suggest that it is this p95/hexokinase isoform which binds zona ligands. However, a direct demonstration that p95/ hexokinase binds zona carbohydrates has yet to be presented. β-1,4-Galactosyltransferase β-1,4-Galactosyl transferase (GalTase) has been postulated to be an externally oriented, integral plasma membrane component of the sperm surface that is necessary for sperm binding to the zona pellucida (Shur and Hall, 1982b; Shur and Neely, 1988; Shur, 1991; Miller et al., 1992). GalTase participates in the synthesis of complex carbohydrates by specifically transferring galactose from its sugar nucleotide substrate, UDP-galactose (UDPGal), to N-acetylglucosamine (GlcNAc), its sugar acceptor substrate. GalTase can transfer galactose from UDPGal to ZP3, but not ZP1 or ZP2. In the absence of manganese, however, GalTase is thought to act as a lectin, not an enzyme. That a lectin with GlcNAc specificity exists on the mouse sperm surface is supported by the fact that mouse sperm agglutination of unfixed rabbit erythrocytes is inhibited by GlcNAc (Kawai et al., 1991). In its capacity as a lectin, it is thought that GalTase reacts with terminal β-linked GlcNAc or N-acetyl galactosaime (GalNAc) residues on the mouse zona pellucida (Benau et al., 1990). The murine zona pellucida has in fact been demonstrated to contain O-linked oligosaccharides with terminal β-linked GlcNAc residues (Nadgas et al., 1994). The latter observation is of particular importance. Although terminal α-galactose residues on mouse ZP3 oligosaccharides have been implicated as being critical in the species-specific binding of mouse gametes (Bleil and Wassarman, 1988), recent data indicate that female mice lacking terminal α-galactosyl residues due to a deficiency in α-glycosyltransferase are fully fertile (Thall et al., 1995). These suggest that β-linked GlcNAc is a critical residue in the regulation of tight binding of mouse spermatozoa to the mouse zona. In competition studies, both purified GalTase and GalTase substrates inhibit sperm–zona binding. Agents which perturb GalTase activity also perturb mouse sperm–egg binding. For example, UDP-dialdehyde, a substrate analogue, and α-lactalbumin, which stimulates GalTase activity towards glucose, produce a dose-dependent inhibition of sperm–egg binding. In addition, UDPGal, which would force GalTAse enzymatic reaction to completion, and enzymatic pretreatment of the zonae, either to galactosylate GlcNAc or to hydrolyse terminal GlcNAc residues, inhibits sperm binding to the zona. Finally, loss of GlcNAc at fertilization appears responsible for loss of sperm binding activity by ZP3. Recent data suggest that an egg cortical granule N-acetylglucosaminidase is released at fertilization, thereby inactivating the sperm GalTase-binding site and producing the block to polyspermy (Miller et al., 1993). GalTase is first observed on primary spermatocytes, where it functions during adhesion to Sertoli cells (Pratt et al., 1993). A surface redistribution of immunodetectable GalTase is observed after meiosis. In freshly ejaculated spermatozoa, GalTase sugar binding sites are masked by epididymally derived glycoside ‘decapacitation factors’ which are then shed during in-vitro capacitation (Shur and Hall, 1982a). GalTase is then localized by indirect immunofluorescence to the anterior sperm head plasma membrane overlying the acrosome. Acid solubilized mouse zonae trigger both induction of the acrosome reaction and redistribution of GalTase to the lateral surface of spermatozoa (Lopez and Shur, 1987). That GalTase–zona binding contributes to the induction of the physiological acrosome reaction is demonstrated by the facts that: (i) aggregation of GalTAse on mouse spermatozoa by bivalent anti-GalTase antibody or by anti-GalTase Fab fragments crosslinked with secondary antibody induces the acrosome reaction (Macek et al., 1991); and (ii) once redistributed to its new membrane domain, GalTase is no longer able to bind ZP3 (Miller et al., 1992). Sperm surface GalTase expression, like phosphorylation of p95/hexokinase (see above), is regulated by the T/t locus (Shur 613 S.Benoff and Bennett, 1979). A significant increase in GalTase activity is observed on t-bearing spermatozoa (Shur and Scully, 1990), which is the result of elevation of mRNA expression during meiotic division (Pratt et al., 1993) and is associated with premature acrosome loss (Brown et al., 1989). That overexpression of GalTase is associated with decreased fertility is directly demonstrated in transgenic mice (Youakim et al., 1994). Spermatozoa from these animals exhibited two defects: (i) increased binding of decapacitation factors, resulting in masking of presumptive zona binding sites, and (ii) premature acrosome loss, also limiting tight binding to the zona. These data suggest that an ‘optimal’ level of zona binding sites on the sperm surface is required to achieve fertilization. There is presumptive evidence that the cytoplasmic tail of sperm surface GalTAse is associated with the cytoskeleton. mRNAs encoding Golgi and cell surface forms of GalTAse are distinguishable by size (Lopez et al., 1991). Mouse spermatozoa preferentially express the long form of GalTase mRNA (Pratt et al., 1993), which targets the encoded protein to the cell surface in somatic cells. In somatic cells, movement of GalTase on the surface membrane is regulated by its attachment to the cytoskeleton (Appeddu and Shur, 1994). As cytoskeletal elements participate in antigen redistribution in association with sperm maturation and the acrosome reaction in other species (Feuchter et al., 1986; Ochs et al., 1986; Saxena et al., 1986a,b), it is reasonable to postulate that translocation of GalTase on developing spermatids and following the acrosome reaction is regulated by GalTase–cytoskeletal interactions. The cytoplasmic domain of GalTase is also linked (through the cytoskeleton?) to a pertussis toxin-sensitive guanine nucleotide binding protein (G protein) (Gong et al., 1995). Ligandstimulated aggregation of GalTase results in activation of a G protein complex containing the Gi α subunit. These observations are consistent with findings indicating that G-protein activation participates in signal transduction leading to acrosome exocytosis following ZP3 binding and that spermassociated Gi subunits are critical to this second messenger system (Ward and Kopf, 1993; Ward et al., 1994). Voltagedependent calcium channels, which regulate calcium influx necessary for acrosome exocytosis (Florman, 1994), are activated by class Gi G-proteins (Ward et al., 1994). Finally, redistribution of GalTase to the equatorial segment and inner acrosomal membrane in association with acrosome exocytosis has been postulated to allow GalTase to play a role in subsequent binding events with the zona pellucida and/or the oolemma (Lopez and Shur, 1987). The relative importance of the above data is somewhat diminished by two findings. First, a body of data now indicates that GalTase activity is associated with plasma membranes derived from bovine, equine and human spermatozoa (Sullivan et al., 1989; Humphreys-Beher and Blackwell, 1989; Humphreys-Beher et al., 1990; Fayrer-Hosken et al., 1991). These data suggest that GalTase may participate in sperm– zona binding and/or acrosome exocytosis in a variety of species. Although GalTase exhibits all characteristics expected of a mouse sperm protein with primary zona recognition activity, it cannot determine species specificity of expression. GlcNAc is a ubiquitous component of cell surface glycopro614 teins. GalTase is likely to be a generalized receptor that functions during initial sperm–egg attachment as part of a complex with molecules regulating species-specific gamete binding. A second and more disturbing finding suggests that the GlcNAc-specific lectin may not be directly concerned with sperm–egg interaction. The results of sugar competition experiments examining sperm–egg binding and sperm–erythrocyte agglutination are not correlated (Kawai et al., 1991). Sialic acid and asialo-transferrin are considerably more effective than GlcNAc in blocking species-specific sperm–zona binding. In addition, the localization of GalTase to the anterior sperm head may be an artefact of antibody-induced cross-linking (Cardullo and Wolf, 1995). Using monovalent fluorescent probes specific for GalTase and the technique of fluorescence recovery after photobleaching, data have been presented indicating that GalTase resides on the posterior sperm head and migrates to the anterior region of the sperm head after cross-linking with polyvalent antibodies. More importantly, this migration does not lead to the release of acrosin. However, induction of the acrosome reaction with the ionophore A23187 leads to the loss of GlcNAc-specific lectin activity from the sperm surface; it is then found in released membrane vesicles (Kawai et al., 1991; Cardullo and Wolf, 1995). This suggests that migration of GalTase may be an early event in the induction of the acrosome reaction and that binding between the zona and other sperm surface molecules are required to complete acrosome exocytosis. An electron microscopic time study has suggested that inital mouse sperm–zona pellucida contact is random (Saling et al., 1979). These findings raise the question of whether the posterior head region of mouse spermatozoon also participates in tight binding to the mouse zona pellucida, as previously described for hamster sperm–zona interactions (Drobnis et al., 1988). Trypsin-like enzyme A trypsin-like enzyme on the surface of the plasma membrane has been postulated to be involved in binding of mouse spermatozoa to the mouse zona pellucida, as spermatozoon binding occurs prior to the acrosome reaction (Florman and Storey, 1982) and trypsin inhibitors competitively block this binding (Saling, 1981). A unique property of this potential zona binding site is that it reacts with trypsin inhibitors but not with trypsin substrates (Benau and Storey, 1987). Although this trypsin inhibitor binding site is separate from that responsible for GalTase activity, the two zona binding sites are apparently located in close proximity on the sperm surface (Benau and Storey, 1988). Binding of ligand at one site elicits a steric hindrance effect which limits ligand binding at the other site. The component(s) of this trypsin inhibitor-sensitive site which mediates zona binding have not been subjected to detailed biochemical and molecular characterization. Thus, the possibility remains that this zona binding site is one and the same with the mouse sperm 15 kDa peripheral protein with proteinase inhibitor binding capabilities described below. It is unlikely, however, that this mouse protein is directly related to trypsin inhibitor binding sites present on the surface of hamster and human spermatozoa (Dravland et al., 1984; Pillai Carbohydrates and fertilization: an overview and Meizel, 1991; Llanos et al., 1993). In these species, trypsin inhibitors block the acrosome reaction (see below). α-D-Mannosidase Given that an α-D-mannosidase activity has been detected on the surface of rat, mouse, hamster and human spermatozoa (Tulsiani et al., 1989, 1990), mouse in-vitro sperm–zona binding assays were performed in the presence of D-mannose, mannose analogues and mannose-containing oligosaccharides (Cornwall et al., 1991). In all cases a dose-dependent decrease in the number of spermatozoa bound per egg and in sperm mannosidase activity was observed. Mannose-containing ligands were not, however, observed to induce the acrosome reaction. Whether this enzyme actually participates in the initial binding of mouse spermatozoa to mouse zonae pellucidae is thus in doubt for two reasons. First, in the studies of Cornwall et al. (1991), galactose, the sugar identified as the mouse sperm receptor on O-linked zona oligosaccharides (Bleil and Wassarman, 1988), had no effect on sperm–egg binding. Second, and more importantly, high mannose/hybrid oligosaccharide chains are found on mouse ZP2 and ZP3 as Nlinked, not O-linked, oligosaccharide chains (Tulsiani et al., 1992). Only O-linked zona oligosaccharides have been convincingly demonstrated to exhibit sperm receptor activity (Florman and Wassarman, 1985; Rossiere and Wassarman, 1992; Kinloch et al., 1995). Sulphoglycolipid immobilizing protein (SLIP)1 SLIP1 is a 68 kDa sulphoglycolipid immobilizing protein which is observed to be distributed on the mouse sperm head in two distinct patterns: (i) over the whole acrosomal region except the convex ridge, and (ii) concentrated in the posterior acrosomal area which borders on the post-acrosome region (Tanphaichitr et al., 1992, 1993). SLIP1 is of testicular origin and binds in vitro to the major sulphoglycolipid found only in mammalian testes and spermatozoa (Lingwood, 1985). Two populations of SLIP1 are detected in ejaculated mouse spermatozoa: one which is peripherally attached and readily released in the absence of detergent and a second which is released after exposure to Triton X-100, suggesting that it is an integral membrane component (Tanphaichitr et al., 1993). Although recognition of ZP3 by SLIP1 has never been directly demonstrated, in in-vitro competition studies, purified SLIP1 inhibits mouse sperm–egg binding (Tanphaichitr et al., 1993). Pre-exposure of mouse eggs to SLIP1 reduces the number of spermatozoa subsequently bound. Exposure of spermatozoa to anti-SLIP1 antibodies also decreases the number of spermatozoa bound per egg. These observations are difficult to reconcile with the fact that SLIP1 is not detected on the convex ridge of the mouse sperm head (Tanphaichitr et al., 1993) which participates in the initial binding of spermatozoa to the mouse zona pellucida (Bleil and Wassarman, 1986). Further, bivalent anti-SLIP1 antibodies do not induce capacitated mouse spermatozoa to undergo acrosome exocytosis. No cDNA or amino acid sequence information is available, but it is clear from the fact that heat inactivates the inhibitory activity of SLIP1 that its native conformation plays an important role in sperm–egg binding. These data indicate that it is unlikely that SLIP1 is the primary mouse sperm zona receptor and/or is associated with signal transduction activity. Rather, it has been suggested that SLIP1 may act synergistically to facilitate zona binding with other sperm proteins (Tanphaichitr et al., 1993). 15 kDa protein A 15 kDa component of the mouse sperm apical plasma membrane has been identified which binds a proteinase inhibitor of seminal plasma origin and, in spermatozoa capacitated in vitro, participates in binding to homologous zonae (Poirier et al., 1986). This protein becomes expressed on the sperm surface during epididymal maturation and is also detected on the luminal surface of epididymal epithelium (Richardson et al., 1987), suggesting that the 15 kDa protein becomes adsorbed to the sperm surface during epididymal transit. Proteinase inhibitor is apparently bound by spermatozoa at ejaculation but the level of bound proteinase inhibitor is significantly reduced following incubation in media which support sperm capacitation but not in media which do not support capacitation (Boettger et al., 1989). Zona glycopeptides and purified seminal plasma inhibitor compete for sperm binding (Boettger-Tong et al., 1992). Competition studies also demonstrate that the seminal plasma inhibitor blocks sperm– zona binding only if added during the first 15 min of coincubation, indicating that the site which binds the proteinase inhibitor participates in the early events of gamete adhesion but not in tight binding (Robinson et al., 1987). Supporting this are observations indicating that immunoaggregation of bound proteinase inhibitor and translocation to the equatorial segment induces the acrosome reaction in a manner paralleling induction of acrosome loss by immunoaggregation of zona peptides produced by pronase digestion (Aarons et al., 1991). Induction of the acrosome reaction was observed only in capacitated spermatozoa; in fresh spermatozoa, aggregation and translocation were without effect on acrosome status (Aarons et al., 1992). Although this 15 kDa sperm protein exhibits many characteristics which would be expected of a receptor for zona pellucida ligands, no information is available as to whether or not this protein specifically binds carbohydrates. sp56 Of all mouse sperm proteins postulated to be involved in the initial stages of binding to the zona, only sp56 exhibits all characteristics predicted for a mouse sperm protein with primary zona recognition activity: (i) highly specific affinity for mouse ZP3; (ii) high affinity for 3.9 kDa O-linked oligosaccharide with sperm receptor activity (‘functional domain oligosaccharide’ 5 FD oligo); (iii) binding to galactose affinity columns, and not N-acetylglucosamine affinity columns; (iv) presence on the surface of acrosome-intact sperm heads, but not on acrosome-reacted spermatozoa; (v) distribution on the convex ridge of the dorsal surface of the sperm head colocalizing with ZP3 binding sites; and (vi) aggregation by FD oligos (Bleil and Wassarman, 1986, 1988, 1990; Cheng et al., 1994; Suzuki-Toyota et al., 1995). sp56 can be purified by cross-linking to zona ligands or by ZP3 affinity chromatography. Post-translational truncation from 62 to 56 kDa results in 615 S.Benoff the production of the active zona receptor (Bookbinder et al., 1995). This truncation results in the deletion of an aminoterminal sequence with high homology to the α subunit of complement 4B-binding protein, suggesting that sp56 may belong to a superfamily of protein receptors (Reid et al., 1986; Cheng et al., 1994; Bookbinder et al., 1995). The purified native protein is a homomultimer (Mr 110 kDa); each monomer is stabilized by disulphide bonds. sp56 is a peripheral membrane protein that can be released from the sperm surface in the absence of detergents by agents that disrupt hydrogen bonds (Cheng et al., 1994). This is consistent with the absence of a transmembrane-spanning domain in the deduced amino acid sequence of this protein (Bookbinder et al., 1995). cDNA sequence analysis also fails to reveal a sequence with homology to the carbohydrate recognition domain of previously characterized galactose lectins (e.g. ‘galectins’, Barondes et al., 1994; Bookbinder et al., 1995). Expression of sp56 mRNA is restricted to mouse spermatids. That sp56 protein epitopes are not detected on guinea pig or human spermatozoa and that sp56 mRNA is not detected on Northern blots of testicular mRNA from guinea pig and man suggests that sp56 participates in species-specific gamete adhesion in the mouse (Bookbinder et al., 1995). Rat Lectin binding studies indicate that the zona pellucida of rat oocytes is particularly rich in mannose and GlcNAc residues (Shalgi et al., 1991). Other sugar residues which are in abundance on the rat zona include glucose, α and β galactose, GalNAc and fucose (Shalgi et al., 1991). α-D-Mannosidase Sugar competition for rat sperm–egg binding and pretreatment of rat oocytes with specific enzymes indicates that sulphated glycans containing α-methyl-mannoside are critical determinants of the sperm binding site on the rat zona pellucida (Shalgi et al., 1986). Consistent with these observations, an α-Dmannosidase is found associated with rat sperm plasma membranes (Tulsiani et al., 1989). Sperm mannosidase exhibits activity mainly towards mannose-containing oligosaccharides. This activity and its pH optimum of 6.5 distinguishes the sperm surface enzyme from ‘acidic’ (pH optimum 5 4.4) α-D-mannosidase of the acrosomal matrix which shows a preference for synthetic substrates. Rat sperm surface mannosidase is proteolytically processed by trypsin-like enzymes during epididymal maturation from a 135 kDa inactive precursor to a 115 kDa enzymatically active mature form (Tulsiani et al., 1995). It is localized on the periacrosomal region of the sperm head (Tulsiani et al., 1995), with its catalytic domain externally oriented (Tulsiani et al., 1989). Both surface mannosidase activity and zona binding ability increase as rat spermatozoa move from the caput to cauda region of the epididymis (Yanagimachi, 1994; Tulsiani, 1995). Thus, it has been proposed that rat sperm surface mannosidase participates in tight binding of spermatozoa to the rat zona pellucida. Whether, however, this novel α-D-mannosidase regulates species-specific gamete binding is in doubt for two reasons. Firstly, the catalytic domain of this protein is purported to interact with high-mannose/hybrid oligosaccharide chains 616 from N-linked, not O-linked, glycoproteins (Tulsiani et al., 1992, 1995). Secondly, surface α-mannosidase activity is also observed in spermatozoa from mouse and men (Cornwall et al., 1990; Tulsiani et al., 1990). Galactosyl receptor Given that lectins play an important role in cellular adhesion processes (Sharon and Lis, 1989), a rat testis and sperm surface calcium-dependent galactosyl receptor (RTG-r), capable of binding galactose and/or GalNAc, has been postulated to be a zona binding protein of rat spermatozoa. RTG-r shares antigenic and sequence homology with a calcium-dependent asialoglycoprotein lectin of liver (Abdullah and Kierszenbaum, 1989; Abdullah et al., 1991; Mertz et al., 1995). The hypothesis that RTG-r is involved in rat sperm–zona interaction is supported by immunolocalization of RTG-r to the dorsal surface of the sperm head overlying the acrosome and loss of immunoreactivity after induction of the acrosome reaction (Mertz et al., 1991). It is important to note that no function has been ascribed to RTG-r. Whether this molecule participates in species-specific binding of spermatozoa to the rat zona pellucida is in doubt as galactosyl residues are predominantly found on the inner, and not the predicted outer layer, of the rat zona pellucida (Shalgi et al., 1991), and because related molecules with similar sperm head topography have been detected on rabbit, human, pig and mouse spermatozoa (Abdullah et al., 1991; Goluboff et al., 1995). Boar Confocal laser-scanning microscopy (CLSM) of boar spermatozoa in hemizona assays has recently suggested that a gradient of sperm binding sites exists on the porcine zona pellucida, with the number of sperm binding sites decreasing as spermatozoa move from the outside to the inside of the zona (Fazeli et al., 1997). The same study also indicates that acrosomeintact boar spermatozoa initiate binding to the homologous zona pellucida in vitro. This is consistent with findings in other species that lectin binding sites are more abundant on the outer as compared with the inner side of the zona (Ahuja and Bolwell, 1983). Prior findings that boar spermatozoa must be at least partially acrosome reacted to bind to the porcine zona pellucida (Jones et al., 1988; Yonezawa et al., 1995) can be explained by the rapidity with which the acrosome reaction is induced by solubilized zona glycoproteins (Berger et al., 1989b). Peterson et al. (1980, 1981) were the first to report that boar spermatozoa bind to homologous zonae via receptors located on the sperm head plasma membrane. More recently, the binding of zona glycoproteins to the acrosomal cap region of the boar sperm surface has been demonstrated by immunofluorescence microscopy (Hanqing et al., 1991). At least three classes of sperm proteins (APz, p105/45, and spermadhesins) have been proposed to be involved in the initial binding of boar spermatozoa to porcine eggs. Zona adhesion protein (APz) Zona adhesion protein (APz) is an Mr 55 kDa integral plasma membrane protein in boar spermatozoa that does not exhibit enzymatic activity, or react with antibodies to proacrosin or bind to fucose, a sugar which binds proacrosin (Petersen and Carbohydrates and fertilization: an overview Hunt, 1989; Peterson et al., 1985, 1991). APz is a minor sperm membrane component, representing ,1% of the total membrane protein and showing microheterogeneity, which exhibits high affinity for sulphated polysaccharides. Antigenically active APz is first detected on epididymal spermatozoa and the amount of APz is increased after ejaculation in association with capacitation. That APz is associated with the membrane cytoskeleton in boar spermatozoa is consistent with the hypothesis that actin–plasma membrane protein interactions play a major role in sperm capacitation (Saxena et al., 1986a,b). Monovalent antibodies to APz completely block sperm–egg binding. Competition studies suggest APz zona and antibody binding sites overlap. Such studies also indicate that APz specifically binds to ZP3α, the component of the porcine zona pellucida with sperm binding activity (Sacco et al., 1989; Yurewicz et al., 1993). Ligand binding is unaffected by GalTase substrates and inhibitors and anti-proteases suggesting that APz is a unique protein. APz tends to aggregate on standing which suggests that hydrophobic interactions are important in carbohydrate binding and indicates that APz exhibits lectinlike properties. APz appears to be a major plasma membrane protein regulating tight binding of acrosome-intact boar spermatozoa to porcine ova. p105/45 p105/45 and p56-62 were isolated after disulphide bond reduction of sperm membrane protein complexes of respectively, Mr 130–170 and 56 000 kDa. The original complexes were purified from detergent solubilized membrane proteins of boar spermatozoa by chromatography over native, particulate porcine zona pellucida glycoproteins (Hardy and Garbers, 1994). The proteins in these complexes represent only a small fraction of total membrane protein and bind with high affinity to the porcine zona pellucida at physiological ionic strength. p56-62 exhibits only limited species specificity with regard to zona binding. On the basis of size, p56-62 could represent GalTase, sp56 or PH-20. Species limited expression of these proteins has not been reported. In contrast, binding of p105/ 45 to heterologous zona (murine, bovine) and Xenopus laevis oocyte envelopes is not observed. p105/45 is likely to be a unique protein, as sequences from tryptic peptides do not match peptide sequences deposited in existing databases. Although the carbohydrate specificities of p105/45 remain to be investigated, these data suggest that p105/45 is of physiological importance. p105/45 has recently been cloned and renamed ‘zonadhesin’ (Hardy and Garbers, 1995). Analysis of the deduced amino acid sequence of its precursor molecule indicates a multidomain structure. A single transmembrane segment separates a short basic intracellular carboxy terminal segment from a large extracellular region containing a mucin-like domain and 5 tandem repeated sequences which are homologous to Ddomains of prepro-von Willebrand factor (vWF). Zonadhesin is apparently synthesized in haploid spermatids and undergoes proteolytic processing during epididymal transit or sperm capacitation to generate p105 and p45 as well as to remove the amino terminal and mucin-like domains. The remaining vWF D-domains are postulated to mediate covalent oligomeriz- ation, promoting membrane protein aggregation and induction of the acrosome reaction, as well as heparin binding. It is therefore interesting to note that heparin-binding proteins are thought to act as extrinsic regulatory capacitation factors that enhance the zona pellucida-induced acrosome reaction (Florman and First, 1988). Thus, it appears that zonadhesin is a multifunctional protein. Spermadhesins Spermadhesins are a family of four low molecular weight proteins (AQN-1, AQN- 2, AQN-3, AWN; 111-113 amino acids), with further diversity regulated by post-translational modification (AWN-1, AWN-2). Post-translation modification also regulates zona binding ability; addition of a single carbohydrate chain blocks their zona pellucida binding sites. The conformation of spermadhesins is regulated by two disulphide bonds which modulate the binding of polysulphated ligands (review, Calvete et al., 1995). Amino acid compositional analyses suggest that spermadhesins may be weakly related to the sea urchin sperm protein bindin which binds to the egg vitelline membrane (Jonakova et al., 1991). Spermadhesins are located on the apical portion of the sperm head plasma membrane overlying the acrosome, consistent with prior findings on the topography of the porcine egg receptor on spermatozoa (Yurewicz et al., 1993). They are derived from secretions of the seminal vesicle epithelium and are added to the sperm surface at the time of ejaculation. These absorbed proteins are retained on the sperm surface as a result of their interaction with a major phospholipid component, phosphorylethanolamine (Dostalova et al., 1995a). The structural determinants for lipid and carbohydrate binding lie on non-overlapping domains of the molecules. Spermadhesins appear to be multifunctional proteins in that they exhibit both lectin-like binding of carbohydrates and heparin-binding activity or serine protease inhibitor binding capabilities (Sanz et al., 1992, 1993; Calvete et al., 1995). Heparin-binding proteins have been discussed above and inhibitors of serine proteases are released from spermatozoa in the female reproductive tract exposing surface zona binding sites (Robinson et al., 1987). This suggests that spermadhesins participate both in sperm–zona binding and earlier events regulating sperm capacitation. Spermadhesins can be purified by affinity chromatography on isolated zona glycoproteins (Hanqing et al., 1991). AQN1, -2 and -3 bind fucoidin (Jonakova et al., 1991; Sanz et al., 1991). AQN-1 also binds acrosin inhibitors (Sanz et al., 1992). By analogy with the proposed function of acrosin as a ZP2 binding protein (Jones, 1989, 1991), these data suggest that AQN proteins are active in secondary binding to the zona pellucida. In contrast, the following characteristics of AWN suggest that this protein participates in the initial steps of sperm–egg recognition. AWN-1 synthesized by the rete testis is the only spermadhesin found on the surface of epididymal spermatozoa and may be a factor contributing to the fertilizing ability of such spermatozoa. Additional AWN molecules become absorbed to the sperm surface on ejaculation and AWN-1 remains preferentially bound to capacitated spermatozoa when other spermadhesins are released (Dostalova et al., 1994). AWN-1 can be detected on capacitated, zona-bound 617 S.Benoff boar spermatozoa. Interaction of capacitated boar spermatozoa with isolated zona pellucidae can be inhibited by preincubation of zonae with purified AWN protein. Consistent with findings in the mouse (Florman and Wassarman, 1985), AWN-1 binds to a minor fraction of O-linked carbohydrate chains of porcine zona pellucida glycoproteins (Hirano et al., 1993; Hokke et al., 1993, 1994; Dostalova et al., 1995b). The basic structure of the preferred ligand is Gal-GalNAc 6 terminal NeuAc (sialic acid) (Dostalova et al., 1995b). AWN-1 molecules are present on boar spermatozoa which bind to zona-encased oocytes (Dostalova et al., 1995b). That these spermatozoa have the capability to penetrate the zona suggests, again by analogy with findings from other species, that AWN-1 participates in the acrosome reaction. Alternatively, AWN-1 could serve simply to anchor spermatozoa to the zona. However, AWN-1 is also capable of binding acrosin inhibitors (Sanz et al., 1992), which suggests that this sperm surface molecule may have more than one function. Further, that AQN-1 and AWN-1 both bind acrosin inhibitors suggests that a complex of molecules is required to effect species specific fertilization. It is interesting to note that similar low molecular weight zona binding proteins have been identified in a variety of species including man (O’Rand et al., 1985). However, although these proteins share antigenic epitopes with AWN-1, their topology on the sperm head differs in human and stallion spermatozoa. Thus, whether or not such proteins play a role in the binding of human and stallion spermatozoa to their homologous zonae remains to be determined. Guinea pig Both acrosome-intact and acrosome-reacted guinea pig spermatozoa can bind to the homologous zona pellucida in vitro (Huang et al., 1981; Myles et al., 1987; Myles and Primakoff, 1997). Only one of the two candidates for the guinea pig sperm zona receptor has been demonstrated to bind carbohydrates and neither candidate shows species restricted expression. Nevertheless, the changing patterns of protein localization during spermatogenesis, epididymal maturation and the acrosome reaction have been well studied (Myles and Primakoff, 1984; Cowan et al., 1986, 1987, 1991; Phelps et al., 1988; Myles and Primakoff, 1997). These studies provide strong evidence for barriers to protein/lipid diffusion in the guinea pig sperm plasma membrane. These studies also emphasize that when a sperm membrane protein is relocated it comes in contact with different lipids or proteins that could alter its function. L-fucose receptor Both L-fucose and fucoidin, an algal molecule containing sulphated L-fucose, are very strong inhibitors of binding of guinea pig spermatozoa to the guinea pig zona pellucida in comparison with a variety of other monosaccharides (Huang et al., 1982). Binding of human spermatozoa to the human zona pellucida and of hamster spermatozoa to the hamster zona pellucida is also inhibited by fucoidin (Huang et al., 1982). Therefore, it is likely that fucose is a part of a common recognition signal regulating binding of mammalian gametes. Whether or not the hamster L-fucose receptor is a homologue 618 of the fucose-binding human sperm L-selectin molecule (see below) remains to be determined. As acrosin from a variety of species binds fucose ligands (Yanagimachi, 1994), an alternative possibility is that the L-fucose receptor of guinea pig spermatozoa is acrosin. The latter would be consistent with data indicating that only acrosome-reacted guinea pig spermatozoa bind to the zona pellucida (Huang et al., 1984). Posterior head antigen (PH)-20 Posterior head antigen (PH)-20 is a surface protein which is inserted into the guinea pig sperm plasma membrane by at least two separate mechanisms (Cowan et al., 1986). Proteolytic processing during epididymal maturation and sperm capacitation is required for specific surface localization and formation of the active protein that comprise two disulphide-linked fragments (Primakoff et al., 1988; Phelps et al., 1990). In mature spermatozoa, PH-20 is located both on the sperm plasma membrane and on the inner acrosomal membrane. PH20 is anchored via a GlPtdIns linkage (Phelps et al., 1988). A role for PH-20 in guinea pig sperm binding to the egg zona pellucida was first derived from findings that a monoclonal antibody to this sperm protein inhibited binding of spermatozoa of undetermined acrosomal status (Primakoff et al., 1985). Studies with antibodies to other sperm surface antigens indicated that they failed to elicit this inhibition and that PH-20 may be complexed on the sperm surface with at least one other protein (Primakoff et al., 1985). PH-20, however, is not required for the binding of acrosome-intact spermatozoa to the zona pellucida (Myles et al., 1987). Competition studies with purified PH-20 indicate that PH-20 inhibits the binding to the zona pellucida of acrosome-reacted, but not acrosome-intact, guinea pig spermatozoa (Myles et al., 1987). Therefore, it must be noted that PH-20 migrates from one surface domain (posterior head) to another (inner acrosomal membrane) in association with the acrosome reaction. At the same time, proteolytic cleavage occurs and the number of PH-20 molecules on the sperm surface is increased. The latter is consistent with findings in somatic cells that cell adhesion requires a threshold density of functional surface molecules (e.g. Doherty et al., 1990). No homology is detected between PH-20 and any other previously identified sperm surface egg binding protein (Lathrop et al., 1990; Lin et al., 1993). PH-20 has a hyaluronidase activity (Lin et al., 1994) which is homologous to that found in bee venom (Gmachl and Kreil, 1993). In addition to zona binding, one suggested function for PH-20 on the posterior sperm head is thus to enable acrosome-intact spermatozoa to penetrate the cumulus layer in order to reach the zona pellucida (Hunnicutt et al., 1996). Cumulus penetration and zona binding are likely to be independent functions as antibodies which block sperm–zona binding have no effect on the hyaluronidase activity of PH-20 and inhibition of hyaluronidase activity with apigenin has no effect on sperm–zona binding (Hunnicutt et al., 1996). This suggestion is consistent with the structure of PH-20. The hyaluronidase activity of PH-20 has been localized to the N-terminal domain of the protein while the active site involved in secondary binding to the zona is localized to the C-terminal end. These domains are separated Carbohydrates and fertilization: an overview by proteolytic cleavage in association with the acrosome reaction (Primakoff et al., 1988; Myles and Primakoff, 1997). The gene encoding PH-20 is found by Southern blotting to be present in genomic DNA from mouse, rat, hamster, rabbit, cow, monkey and man (Lathrop et al., 1990). PH-20 is found on the plasma membrane of spermatozoa from a variety of species including man (Lin et al., 1993, 1994). However, in contrast to the guinea pig, PH-20 is localized to the anterior or whole head in the mouse, macaque and man (Lin et al., 1994; Overstreet et al., 1995) and to the tail in the rat (Jones et al., 1990). Expression of human PH-20 mRNA is restricted to the testis (Lin et al., 1993). In mouse spermatozoa, PH-20 is expressed as both a soluble form with the acrosomal matrix which is released during the acrosome reaction and GlPtdInslinked isoform which is present on the surface of acrosomeintact and acrosome-reacted spermatozoa (Thaler and Cardullo, 1995). Similarly, cynomologus macaque spermatozoa exhibits two forms of PH-20 hyaluronidase (Cherr et al., 1996). In addition, results from immunolabelling of fixed macaque spermatozoa indicate that PH-20 is distributed over most of the head of acrosome-intact spermatozoa and migrates to the inner acrosomal membrane after the acrosome reaction. This suggests that this protein plays a role in primary or secondary binding of macaque spermatozoa to macaque zona (Overstreet et al., 1995). Thus, it may be hypothesized that PH-20 participates in a complex which regulates the species specificity of sperm–zona binding, but by itself, is not responsible for species limited tight binding of gametes. Rabbit Rabbit sperm–zona interaction is unusual for two reasons. Firstly, although as in other species, acrosome-intact rabbit spermatozoa bind to the rabbit zona pellucida and then acrosome react at the zona surface (Bedford, 1968, 1972), acrosome-reacted spermatozoa recovered from the perivitelline space can also bind and penetrate additional zonae (Kusan et al., 1984). Secondly, the numbers of spermatozoa penetrating the zona and entering the perivitelline space is considerably higher than that reported for other species, and is sometimes in the hundreds (Bedford, 1971; Overstreet and Bedford, 1974; Berrios and Bedford, 1979). Galactose residues on the rabbit zona pellucida may be the primary determinant regulating tight binding of rabbit spermatozoa. Activation of rabbit spermatozoa is associated with a larger increase in binding for galactose ligands in comparison with all other sugar ligands tested (see Figure 3C). Both rabbit sperm surface proteins identified as exhibiting zona ligand binding affinity specifically bind galactose. Rabbit sperm autoantigen (54 kDa) Rabbit sperm autoantigen (RSA) is a membrane glycoprotein which exhibits lectin-like binding of sulphated, complex carbohydrates with high affinity (O’Rand et al., 1988). Lower affinity binding of monosaccharides, e.g. galactose, is also observed. RSA is thought to be a 54 kDa protein which is capable of zona binding as identified by zona blotting (O’Rand et al., 1985). In acrosome-intact rabbit spermatozoa, RSA is localized over the post-acrosomal–equatorial region border (Esaguy et al., 1988). RSA is relocalized after the acrosome reaction in large aggregates in the medial aspect of the equatorial region (Esaguy et al., 1988). High affinity zona binding is effected by the aggregated form of RSA (O’Rand et al., 1988). This protein shares antigenic epitopes with the rat sperm galactose lectin (RTG-r) (Abdullah et al., 1991). Purified rabbit testis galactose receptor is composed of two subunits, 54 and 49 kDa (Abdullah et al., 1991). Purified rabbit testis galactose receptor readily binds heat solubilized rabbit zonae (Abdullah et al., 1991). These data indicate that the 54 kDa RSA zona binding protein is one component of the rabbit sperm galactose receptor. RSA (Sp17) A 17 kDa rabbit sperm protein (Sp17), which is a member of the rabbit sperm autoantigen (RSA) family, binds with high affinity to rabbit zonae pellucidae and sulphated carbohydrates when in an aggregated state (O’Rand et al., 1988). Native Sp17, identified on Western blots following cross-linking of zona glycoproteins, effectively competes with capacitated rat spermatozoa for zona pellucida binding and Sp17–zona binding is strongly inhibited by complex sulphated carbohydrates including fucoidin, dextran sulphate, chondroitin-β sulphate and dextran. Recombinant Sp17 exhibits similar properties (Richardson et al., 1994; Yamasaki et al., 1995). These data, in conjunction with observations that cholesterol-3-sulphate has only limited effects on binding of Sp17 to rabbit zonae, indicate that Sp17 functions as a rabbit sperm zona pellucida binding protein. The deduced amino acid sequence of Sp17 exhibits three important homologies with other protein sequences: (i) the amino terminus is 43% identical to human testis cAMPdependent protein kinase type IIa regulatory subunit; (ii) the amino terminus also contains residues critical to the formation of the galactose binding domain of rat hepatic lectin (e.g. Halberg et al., 1987; Drickamer, 1992); and (iii) there is a domain which is extremely similar to the calmodulin binding site of neuromodulin (Richardson et al., 1994). Thus, as observed in other species, it is likely that the rabbit sperm receptor for the zona pellucida is a multi-functional, lectinlike protein. Ultrastructural mapping of fixed rabbit spermatozoa before and after the acrosome reaction demonstrates that Sp17 is localized and aligned along with actin under the plasma membrane (Welch and O’Rand, 1985; Esaguy et al., 1988) and suggests that cytoskeletal interactions are responsible for the limited lateral mobility of this antigen (O’Rand, 1982). A reversible plasma membrane association is possible via interaction with palmitic acid (Richardson et al., 1994). The Sp17 polypeptide appears closely associated with the inner aspect of the rabbit sperm apical plasma membrane and is only accessible to antibodies after the initiation of the acrosome reaction (Richardson et al., 1994). For Sp17 to function as a zona binding protein, it is argued that rabbit spermatozoa begin the acrosome reaction within the cumulus oophorus as the result of zona glycoprotein synthesis by granulosa cells (Richardson et al., 1994) and that acrosome-reacted rabbit spermatozoa can bind to the zona pellucida (O’Rand and Fisher, 1987). Two lines of evidence support this argument. Firstly, although zona 619 S.Benoff glycoprotein synthesis differs according to species of origin, cultured rabbit granulosa cells synthesize zona glycoproteins (Lee and Dunbar, 1993). Secondly, it has been demonstrated that fluorescein isothiocyanate (FITC)-labelled, heat-solubilized rabbit zonae bind in patches along the apical ridge of the rabbit sperm head, consistent with early fenestration, and that such solubilized zonae will induce the acrosome reaction (O’Rand and Fisher, 1987). Homologues for Sp17 can be found in mouse and human testis (O’Rand et al., 1985; Naz and Ahmad, 1994; Kong et al., 1995) and immunization of female mice with synthetic peptides derived from the rabbit Sp17 sequence results in infertility (O’Rand et al., 1993; O’Rand and Widgren, 1994), raising the question of whether this protein participates in the regulation of species-limited binding to the zona pellucida. Here, an argument can be put forth that the specificity of zona binding and induction of the acrosome reaction resides in the affinity of zona binding rather than binding itself (O’Rand and Fisher, 1987). Although rabbit spermatozoa are observed to bind to pig zonae, solubilized pig zonae will not induce acrosome loss by capacitated rabbit spermatozoa (O’Rand and Fisher, 1987). Human Although both acrosome-intact and acrosome-reacted human spermatozoa can initiate attachment to the human zona pellucida (Morales et al., 1989), a considerable body of evidence now indicates that, for fertilization to occur, human spermatozoa must be acrosome-intact when they bind to homologous zonae and must undergo an acrosome reaction after zona binding (for review see S.Benoff et al., manuscript submitted). Sugar competition of sperm–zona binding, lectin binding specificities of the human zona pellucida, and direct detection of carbohydrate ligand binding sites on the sperm plasma membrane have indicated that human spermatozoa express surface receptors for a variety of simple and complex saccharide moieties, including fucoidin (Oehninger et al., 1990, 1991), fucose (Tesarik et al., 1993b; Lucas et al., 1994), fructose (Mori et al., 1993), GlcNAc (Brandelli et al., 1994), and mannose (Mori et al., 1989; Oehninger et al., 1991; Benoff et al., 1993c; Brandelli et al., 1994; Maymon et al., 1994) (see Figure 3A). Of these, the best studied receptor is that involved in mannose binding and is the only sugar ligand receptor whose expression has been directly correlated with naturally occurring and drug-induced male infertility and with fertilization outcome following conventional IVF insemination (Tesarik et al., 1991; Benoff et al., 1993b–d, 1994b, 1995a, 1996a; S.Benoff et al., manuscript submitted; Hershlag et al., 1995, 1996). Mannose lectin Ejaculated human spermatozoa may require 3–5 h preincubation before they are capable of binding to native zonae pellucidae or recombinant ZP3 and undergoing an acrosome reaction after binding (White et al., 1990; Morales et al., 1994; Van Duin et al., 1994). This latency in response is consistent with the time course of appearance of mannose ligand receptors on the sperm surface during capacitation (Benoff et al., 1993b). The topographical distribution of mannose receptors on the 620 Figure 4. The topographical distribution of human spermatozoa surface mannose ligand binding sites and acrosome status are compared (Benoff et al., 1993b,d; S.Benoff et al., manuscript submitted). Mannose ligand binding sites are visualized on motile spermatozoa following reaction with fluorescein-conjugated, α-Dmannosylated serum albumin (Man–FITC–BSA; upper panel). Acrosome status is assessed following ethanol permeabilization and reaction with tetramethyl-rhodamine labelled Pisum sativum agglutinin (RITC–PSA; lower panel). Mannose receptors first appear over the entire head region of acrosome-intact spermatozoa during capacitation. Aggregation and translocation to the equatorial/ postacrosome region occurs in association with loss of acrosome content. Note that the sites in the acrosome content to which RITC–PSA binds are progressively lost following exocytosis (Tesarik et al., 1993a). human sperm surface strongly correlates with acrosome status (Figure 4; Benoff et al., 1993b,d, 1995c; S.Benoff et al., manuscript submitted). Double labelling experiments demonstrate that mannose receptors first appear uniformly distributed over the acrosome cap. Binding of polyvalent mannosecontaining ligands induces receptor aggregation and translocation to the equatorial/post-acrosome region of the spermatozoa (S.Benoff et al., manuscript submitted). This translocation is followed by induction of the acrosome reaction (Benoff et al., 1993c,d, 1996a). That the time of induction of the acrosome reaction is markedly decreased when polyvalent mannose ligands are mixed with D-mannose monosaccharide indicates that allosteric interactions between mannose carbohydrate recognition domains within a single mannose receptor contributes to receptor function (S.Benoff et al., manuscript submitted). In addition, mannose receptor apparently forms a complex with non-nuclear progesterone receptors on the surface of capacitated human spermatozoa (Benoff et al., 1995b), suggesting that in-vivo progesterone and zona pellucida ligand binding act in conjunction to produce the physiological acrosome reaction. The level of non-esterified cholesterol in the sperm plasma membrane regulates the surface appearance of functional mannose binding sites (Figure 5; Benoff, 1993; Benoff et al., 1993c,d). While it is recognized that numerous biochemical changes occur during sperm capacitation (Benoff, 1993; Drobnis, 1993; Shi and Roldan, 1995; Zeng et al., 1996), it is the efflux of cholesterol to external sterol acceptors which creates a state that thermodynamically favours the insertion of Carbohydrates and fertilization: an overview Figure 5. The diagram depicts the four stages in the expression of mannose lectins (MLs) on the surface of motile human spermatozoa. mannose receptors into the plasma membrane from subplasmalemmal stores. Anti-sperm antibodies and calcium ion channel blockers limit membrane cholesterol efflux and thus inhibit the appearance of mannose binding sites on the human sperm surface (Benoff et al., 1993a, 1994b, 1995a; Hershlag et al., 1996). In both retrospective, blinded analyses (Benoff et al., 1993b,c, 1994b, 1995a, 1996a) and for the first time in blinded, prospective analyses (e.g. see Figure 6; Benoff et al., 1996b; S.Benoff et al., manuscript submitted; Hershlag et al., 1996), my laboratory has demonstrated that the fractional increase, following capacitating incubation, in the percentages of acrosome-intact spermatozoa capable of binding mannose ligands, such as mannosylated bovine serum albumin (‘Man–FITC– BSA’); (Figure 7A), and the ability to undergo mannosestimulated acrosome loss are predictive of the rate of fertilization of mature oocytes following insemination in vitro. These data support the argument that human sperm mannose receptors contribute to the initial adhesion of spermatozoa to egg. That mannose-containing ligands on the human zona pellucida are involved in species-limited tight binding of gametes is further supported by the failure to detect naturally occurring antisperm antibodies to mannosyl-terminated antigens in the sera from female immune infertility patients (Cooper and Bronson, 1990; Cooper et al., 1991), as this could result in autoimmune disease. Human sperm mannose receptors require calcium to bind tightly to mannose-containing ligands (Benoff et al., 1993b) and surface expression of mannose receptors is highly correlated with the level of zinc in seminal plasma (1996b). Figure 6. Prospective analysis of 70 consecutive males undergoing an in-vitro fertilization (IVF) cycle. Two semen specimens were obtained from each patient prior to IVF. Motile spermatozoa populations were isolated and analyzed for mannose receptor expression both at the time of isolation and again after 18 h incubation in Ham’s F-10 1 30 mg/ml human serum albumin at 37°C in 5% CO2. Where sufficient numbers of spermatozoa were obtained (n 5 28), capacitated spermatozoa were assessed for mannose-stimulated acrosome loss. Both the increase in percentages of acrosome-intact spermatozoa capable of binding mannose ligands on their heads after capacitating incubation and the ability of capacitated spermatozoa to undergo an acrosome reaction upon exposure to mannose ligands directly correlate with the rate of fertilization following conventional IVF inseminations (respectively, Pearson Product Moment Correlation, r 5 0.571, P ,0.0001, and t-test, P ,0.0016) (Benoff et al., 1996b; Hershlag et al., 1996; A.Hershlag and S.Benoff, unpublished observations). These observations suggest that the mannose receptor is a lectin, as transition metal ions help to position amino acid residues and regulate carbohydrate binding in a variety of plant and animal lectins (Sharon, 1993). My laboratory has investigated the structure of the human sperm surface mannoseligand binding site through analysis of its effector specificities, by reaction with specific polyclonal and monoclonal antisera, and by cDNA sequence analysis (Benoff et al., 1992a,b, 1993a,d–f, 1994a, 1996b; S.Benoff, unpublished observations). The results from these studies indicate that this putative zona acceptor shares the amino acid sequence and ligand binding characteristics of the conserved carbohydrate recognition domain (CRD) common to a family of membrane-bound and soluble mannose-specific lectins (Drickmer, 1989; Taylor et al., 1992). It is therefore unlikely that the CRD of the human sperm mannose lectin would behave as a ‘self’ antigen. The contention is supported by observations indicating that naturally occurring antisperm antibodies in the sera of immune infertility patients (both men and women) fail to react with the mannose lectins expressed on the surface of capacitated human spermatozoa (Benoff et al., 1993a, 1995b). These data on the structure of the human sperm mannose lectin also provide a basis for speculation on the nature of the specific oligosaccharides on the human zona pellucida with sperm binding activity that are recognized by this marker. Other C-type mannose lectins have been reported to be ‘corespecific’, recognizing not only terminal sugars but also those within the core region of an oligosaccharide (Summerfield and Taylor, 1986; Taylor and Summerfield, 1987; Childs et al., 621 S.Benoff Figure 7. The topographical distribution of the binding of fluorescein-conjugated polyvalent mannose ligands on the surface of motile human spermatozoa (A, 3600 original magnification) was compared with that detected by indirect immunofluorescence of anti-myosin (B, 3400 original magnification) and anti-actin (C, 3400 original magnification) antibodies. Antibody labelling was performed following permeabilization of unfixed human spermatozoa (Benoff et al., 1996a; S.Benoff et al., manuscript submitted). Labelled spermatozoa were photographed on 400 ASA film with a 50 s exposure time. Mannosylated BSA and anti-cytoskeletal antibodies exhibit similar distributions over the spermatozoa head: over the acrosome cap (pattern II; large arrow) and limited to the equatorial segment (pattern III: small arrow). Double labelling with rhodamine-conjugated Pisum sativum agglutinin, which stains acrosome content (Cross et al., 1986), has revealed that pattern II binding occurs on acrosome-intact human spermatozoa while pattern III is observed on spermatozoa which have undergone an acrosome reaction (Benoff et al., 1993b,d, 1996a, 1997a; S.Benoff et al., manuscript submitted). 1990). These findings suggest that mannose need not be a terminal residue of the glycoconjugates of human ZP3 in order to regulate the initial steps in species specific tight binding of human gametes. Analysis of the deduced amino acid sequence of the human sperm mannose lectin indicates that, as in other species, this putative zona binding molecule is multifunctional. Its cytoplasmic tail contains a binding/hydrolysis site for ATP and two actin binding motifs, all of which exhibit both antigenic and sequence homology to those found in the globular head of skeletal muscle myosin (Figure 7B; Benoff et al., 1994a, 1996a). A superfamily of human sperm plasma membrane proteins containing similar cytoplasmic ‘myosin motors’ was identified during the course of this analysis (Benoff et al., 1994a, 1996a). This discovery raises the possibility that directed movement and clearance of these molecules from the apical plasma membrane prior to the acrosome reaction is controlled by interaction of their myosin motors with underlying actin (Benoff et al., 1997a; S.Benoff, unpublished observations). Mannose lectins share characteristics with actin binding proteins of somatic cells (for review see Isenberg, 1991), in that their expression on the sperm surface is regulated by membrane fluidity changes (Benoff, 1993; Benoff et al., 1993a,c,d, 1994b, 1995a, 1996a). Actin is readily detected in the unfixed permeabilized human sperm head by indirect immunofluorescence cytochemistry (Figure 7C; Benoff et al., 1997a). Other laboratories have reported that cytoskeletal actin participates in antigen redistribution associated with induction of the acrosome reaction (Feuchter et al., 1996; Ochs et al., 1986) and disruption of actin filaments with cytochalasin D reduces the number of human spermatozoa capable of penetrating zona-free hamster ova (Rogers et al., 1989), an event requiring prior acrosome loss. Recent findings demonstrate that human spermatozoa which exhibit reduced expression of either plasma membrane proteins with anti-myosin antibody 622 reactivity or sperm head actin also exhibit a reduced ability to undergo spontaneous and induced acrosome loss (Benoff et al., 1996a, 1997a). These data support the argument that actin– myosin-like interactions are involved in the rearrangements of proteins in the human sperm head plasma membrane and induction of the acrosome reaction. α-D-Mannosidase Despite the findings described above, it is also possible that mannose-based human sperm–egg interactions could theoretically be mediated, at least in part, by carbohydrate binding enzymes. Another laboratory has described an α-D-mannosidase activity associated with the human sperm plasma membrane (Tulsiani et al., 1990) which is similar to those of rat and mouse sperm plasma membrane (Tulsiani et al., 1989; Cornwall et al., 1991) and can be differentiated from the acidic αmannosidase activity associated with the acrosome (Tulsiani et al., 1990). The effector specificities of this membrane-bound mannosidase activity and that of the mannose lectin described above differ significantly, such that one could not readily be confused for the other. This enzyme has only been studied in freshly isolated human spermatozoa. Despite generalizations from data obtained in the mouse (see above), there is no direct evidence that the sperm surface-associated α-mannosidase regulates interactions between human gametes. In fact, an argument against an important role for this enzyme in human sperm–zona binding rests in the fact that only 5% of its total activity is found tightly associated with sperm plasma membrane fractions (Tulsiani et al., 1990). hu9 A 95 kDa antigen is localized to the equatorial segment of acrosome-intact human spermatozoa and shares antigenic epitopes with hamster spermatozoa (Moore et al., 1987). A monoclonal antibody, mAB 97.25, that reacts with this antigen prevents human sperm binding to the zona pellucida of salt- Carbohydrates and fertilization: an overview stored human oocytes (Moore et al., 1987). It is possible that this antigen is the 94 6 3 kDa autophosphorylating protein of human spermatozoa which has been independently identified (Naz et al., 1991a) and which may be identical with fertilization antigen-2 (FA-2) of human spermatozoa (Naz et al., 1993). Both the 94 6 3 kDa antigen and FA-2 have been localized to the acrosomal region of the human sperm head using indirect immunofluorescence with an anti-phosphotyrosine monoclonal antibody. Phosphorylation of this protein is increased during human sperm capacitation (Naz et al., 1991a; Luconi et al., 1996; Emiliozzi and Fenichel, 1997) and is further stimulated by exposure to zona pellucida glycoproteins (Naz et al., 1991a). Binding of this 94 6 3 kDa protein to ZP3 apparently involves its phosphotyrosine residues, as it can be eluted from ZP3 in the presence of excess o-phospho-L-tyrosine (Naz and Ahmad, 1994). The 95 kDa/94 6 3 kDa/FA-2 protein described above may have been independently identified as hu9 (Burks et al., 1995). hu9 is a potential human sperm homologue of mouse spermatozoa p95 (see above). cDNA sequencing indicates that hu9 is a transmembrane protein which exhibits no homology with hexokinase. Indirect immunofluorescence analysis with mAB 97.25 (Moore et al., 1987; see above) localizes hu9 to the surface in the acrosomal region of capacitated human spermatozoa (Burks et al., 1995) and autophosphorylation of hu9 is stimulated by capacitation and by exposure to human ZP3, characteristics which are shared with the 94 6 3 kDa human sperm protein described above. In competition studies, synthetic peptides to its extracellular region inhibited human sperm–zona pellucida binding. These data suggest that p95 is an unlikely candidate for a receptor regulating species specific tight binding of gametes. The possibility exists, however, that hu9 is part of a complex which regulates sperm–zona interaction and/or subsequent signal transduction. Particularly relevant to this argument is the finding from cDNA sequence analysis that hu9 contains an intracellular tyrosine kinase domain that can participate in both autophosphorylation and transphosphorylation events (Burks et al., 1995). With regard to the latter, my laboratory has demonstrated that mannose ligand binding is associated with an increase in tyrosine phosphorylation on proteins in the human sperm head (Figures 8 and 9; Benoff et al., 1995c). It is likely that activation of a tyrosine kinase (Bielfeld et al., 1994) is ultimately responsible for calcium influx through voltage-dependent calcium channels in the post-acrosome region of the human sperm head (Figure 10; Goodwin et al., 1997; A.Jacob, L.O.Goodwin, N.B.Leeds and S.Benoff, unpublished observations) and the acrosome reaction. In men with iatrogenic infertility associated with the therapeutic administration of calcium ion channel blockers, mannose-stimulated tyrosine phosphorylation is markedly reduced (Figure 8; Benoff et al., 1995c). That mannose receptor aggregation and translocation precedes induction of this tyrosine kinase activity is demonstrated in the same infertile population by the fact that receptor movement can be observed in the absence of an acrosome reaction (Benoff et al., 1994b, 1995a). The human sperm mannose lectin is not itself phosphorylated during capacitation or the acrosome reaction and exhibits no sequence homology to catalytic domains from identified phosphotyrosine Figure 8. Examination of the affinity of motile human spermatozoa populations for anti-phosphotyrosine monoclonal antibodies following previously described protocols (Naz et al., 1991a). Briefly, motile spermatozoa populations were isolated by swim- up and divided in half. One half was used to measure phosphotyrosine accumulation in the fresh isolate (CONTROL). The other half was incubated 18 h under spermatozoa capacitating conditions (Ham’s F10 1 30 mg/ml human serum albumin, 37°C, 5% CO2; 5 INCUBATED). Half of each control and incubated aliquot was exposed to 100 µg/m mannosylated bovine serum albumin (BSA) in the presence of 75 mM D-mannose to induce acrosome exocytosis (Benoff et al., 1993c,d, 1995c, 1996a; S.Benoff et al., manuscript submitted) (1MANNOSE) and the other half was exposed to phospate-buffered saline. Spermatozoa were then air-dried onto glass microscope slides, fixed by immersion in 100% ethanol and then stained with anti-phosphotyrosine monoclonal antibody clone 1G2. Spermatozoa bound monoclonal was detected with fluorescein isothiocyanate-conjugated sheep anti-mouse immunoglobulin G. For each specimen analysed, .300 spermatozoa were scored by visual inspection at 3400 original magnification for anti-phosphotyrosine binding in the head versus tail regions. (A) The percentages of motile spermatozoa binding anti-phosphotyrosine monoclonal antibodies on their heads was quantitated in specimens from five fertile donors. There was a significance difference between the four conditions examined (repeated measures analysis of variance, P ,0.0001). In all cases, a specific increase in spermatozoa head labelling is observed following mannose exposure (CONTROL versus INCUBATED, not significant; CONTROL versus INCUBATED/1MANNOSE, P ,0.004; INCUBATED versus INCUBATED/1MANNOSE, P ,0.0038) (A.Jacob and S.Benoff, unpublished observations). (B) In spermatozoa from men taking nifedipine, a Ca21 channel blocker medication, signal transduction in response to mannose treatment is markedly reduced (A.Jacob and S.Benoff, unpublished observations). This reduction can be attributed to the inhibition of surface mannose receptor expression resulting from drug-related increases in membrane cholesterol content (Benoff et al., 1994,1995a; Hershlag et al., 1995). 623 S.Benoff Figure 9. Proteins in the head of capacitated human spermatozoa from fertile donors are phosphorylated on tyrosine residues following exposure to model zona ligands, polyvalent mannosylated bovine serum albumin (BSA) plus mannose monosaccharide. Phosphotyrosine residues were detected by immunocytochemical staining as described in the legend to Figure 8. Labelled spermatozoa were viewed at 3600 original magnification and photographed at 31500 on 400 ASA film with an exposure time of 50 s. (A) Phase-contrast image of unfixed, capacitated human spermatozoa. (B) Examination of the same field with UV-epifluorescence illumination shows that binding of antiphosphotyrosine monoclonal antibodies occurs over the spermatozoa head in two distinct topographical distributions: over the acrosome cap (small arrow) or limited to the equatorial segment (large arrow) (A.Jacob and S.Benoff, unpublished observations). kinases (S.Benoff, unpublished observations) raising the possibility that either this lectin is directly complexed with hu9 on the sperm surface or that cross-talk between mannose lectins and hu9 occurs via the cytoskeleton. In the mouse, cytoskeletal-associated sperm membrane proteins become phosphorylated during capacitation and induction of the acrosome reaction (Visconti et al., 1996). β-1,4-galactosyltransferase Whether or not human spermatozoa express a GalTase activity has been previously debated (Tulsiani et al., 1990). However, recent data suggest that a human sperm associated GalTase, with biochemical characteristics similar to those of mouse sperm GalTase, segregates with the plasma membrane when spermatozoa are vortex fractionated (Huszar et al., 1997). Its role in human sperm–zona pellucida binding has been assessed only indirectly (Huszar et al., 1997). Surprisingly, immature spermatozoa containing residual cytoplasm and which lack zona binding ability overexpress GalTase when compared with mature spermatozoa. Nevertheless, these findings still do not rule out a more generalized role for GalTase in human sperm– egg interactions. In man, expression of a GalTase allelic variant is observed in a population of human males whose spermatozoa fail to penetrate zona-free hamster ova (Humphreys-Beher and Blackwell, 1989), presumably due to the failure to undergo the acrosome reaction. Antibodies to sperm surface GalTase may play a role in the aetiology of human male immune infertility (Humphreys-Beher et al., 1990). Thus, human sperm surface GalTase may be part of a recognition system common to mammalian gametes. Trypsin-like enzyme The human sperm surface, like that of the mouse, is capable of specific binding of trypsin inhibitors. However, in contrast to findings in the mouse, trypsin inhibitors have no effect on human sperm–zona binding (Llanos et al., 1993). Rather, these agents block the zona- and follicular fluid-induced acrosome reactions by preventing the increase in intracellular calcium 624 required for acrosome exocytosis (Pillai and Meizel, 1991; Llanos et al., 1993). Therefore, a possible role for such trypsin inhibitor binding sites in human sperm–egg interactions lies in regulation of signal transduction after zona binding. 15 kDa A human sperm homologue of the mouse spermatozoa 15 kDa protein may exist. A 20 kDa component of the human sperm head plasma membrane apparently recognizes the murine trypsin–acrosin inhibitor and immunoaggregation of bound inhibitor can induce the acrosome reaction in capacitated human spermatozoa (Boettger-Tong et al., 1993). However, the ability of this protein to block human sperm–zona binding in a hemizona assay has not been tested and no reports have appeared documenting defective 15 kDa protein expression in human male infertility. Therefore, the role of this protein in tight binding of human spermatozoa to the human zona pellucida is questionable. Galactose lectin Human testis and spermatozoa express a 50 kDa protein which is antigenically related to galactosyl lectin found on rat (RTG-r) and rabbit spermatozoa (Goluboff et al., 1995). This lectin is postulated to be testis-specific as Northern blotting experiments indicate that expression of the RNA encoding the human homologue of RTG-r is restricted to the human testis (Goluboff et al., 1995). This galactosyl receptor is localized to the head plasma membranes overlying the acrosome (Goluboff et al., 1995). Its role in tight binding of human spermatozoa to the human zona pellucida and/or induction of the acrosome reaction after zona binding has not been directly assessed. Thus, there is no compelling evidence that this 50 kDa protein plays an important role in the species specific interaction of human gametes. The human sperm galactose receptor shares at least two biochemical characteristics with the mannose lectin described above and with selectins (see below): (i) a calcium-dependency of carbohydrate ligand binding; and (ii) a requirement for non- Carbohydrates and fertilization: an overview Figure 10. L-type calcium channels are detected in the postacrosome region of the spermatozoa head following reaction of unfixed, Triton-permeabilized motile human spermatozoa with monoclonal antibody IIF7, specific for the α1 subunit of the rabbit skeletal muscle calcium channel (Leung et al., 1987; a generous gift of Dr Kevin Campbell, University of Iowa, Iowa, USA) (Goodwin et al., 1997). Labelled spermatozoa were viewed at 3600 original magnification and photographed on 400 ASA film with an exposure time of 50 s for the fluorescein image. (Upper panel) Phase-contrast image. (Lower panel) Corresponding indirect immunocytochemical labelling patterns detected using UVepifluorescence optics. ionic detergents for release from sperm membranes (Goluboff et al., 1995). Mannose ligand binding by mannose lectins can be partially inhibited by high concentrations of D-galactose (Summerfield and Taylor, 1986). A few simple amino acid changes in the carbohydrate recognition domain of mannose lectins result in the acquisition of high affinity galactose binding (Drickamer, 1992; Iobst and Drickamer, 1994b). In the absence of amino acid sequence data, the possibility that the mannose lectin has been misidentified as a galactose lectin cannot be ruled out. L-selectin Solubilized human zona pellucida glycoproteins compete with a fucosylated–BSA ligand for binding sites on the human sperm surface, suggesting that human spermatozoa possess specific receptors for fucose (Tesarik et al., 1993b). Using fucosylated–BSA conjugated to a fluorescent tag, these receptors were localized to the acrosome region of the head of acrosome-intact spermatozoa and the percentages of spermatozoa capable of binding this ligand increased in association with capacitation (Tesarik et al., 1993b). As immunoreactive acrosin is observed on the sperm plasma membrane early in the induction of the acrosome reaction (Tesarik et al., 1988), it was initially suggested that this fucose receptor could be acrosin (Tesarik et al., 1993b). However, the involvement of fucose receptors unrelated to acrosin could not be excluded. Further studies have established that fucose-containing ligands, such as fucoidin, block both human sperm–zona binding and the zona-induced acrosome reaction (Oehninger et al., 1990, 1991; Mahoney et al., 1991). Fucoidin is a structural analogue of both the sulphated carbohydrate ligands that bind to selectins (Patankar et al., 1993) and blood grouprelated antigens found on the surface of the human zona pellucida (Lucas et al., 1994). It is therefore not surprising that glycodelin (5 human placental protein 14), which contains oligosacchride chains that can function as selectin ligands (Dell et al., 1993; Varki, 1994), and purified blood group antigens terminating in fucose, reduce the capacity for tight binding of human spermatozoa to the human zona (Oehninger et al., 1995; Clark et al., 1995, 1996a,b). Consistent with this data, monoclonal anti-L-selectin (‘leukocyte selectin’) antibodies have identified a 90 kDa human sperm protein which is localized to the sperm head (Lucas et al., 1995). Pretreatment of spermatozoa with these antibodies blocks tight binding of human spermatozoa to the zona pellucida under hemizona conditions (Lucas et al., 1995). Whether this Lselectin-related protein regulates species specific gamete interactions may still be questioned as L-fucose is found on the zona pellucida from a variety of species. Likewise, fucose receptors have been described on spermatozoa from other mammalian species. L-selectins represent one of three groups of related receptors that play a role in the regulation of inflammation and the immune response (McEver, 1994). In all cases, ligand binding is calcium-dependent and the result of an amino-terminal motif related to the carbohydrate recognition domain (CRD) of mannose-binding proteins (Erbe et al., 1992; McEver, 1994). Both selectins and mannose lectins recognize ‘clustered saccharide patches’ (Varki, 1994; Taylor et al., 1992; S.Benoff et al., manuscript submitted). Mannose lectins are also capable of binding fucose (Summerfield and Taylor, 1986; Taylor and Summerfield, 1987; Childs et al., 1990; Taylor et al., 1992; Iobst et al., 1994b). Thus, the mannose lectin may have been misidentified as a selectin. The role of sperm surface L-selectin-related protein in human male infertility has never been directly studied. Fertilization antigen-1 Fertilization antigen-1 (FA-1) is an antigen that is implicated in the generation of involuntary human immunoinfertility (Naz, 1987a). FA-1 has been localized on the human sperm surface using monoclonal antibodies (Naz et al., 1992). All spermato625 S.Benoff zoa exhibit antibody binding on their midpiece and tail regions. Additionally, 20–40% of spermatozoa exhibit antibody binding on the post-acrosome region of the head. This same antibody blocks human sperm–zona pellucida binding but has no effect on sperm movement characteristics or on the acrosome reaction. FA-1 exists in the form of a monomer of 23 kDa and/or a dimer of 51 kDa (Naz, 1988). The 51 kDa form exhibits autophosphorylating activity in an in-vitro kinase assay. As observed for the 94 6 3 kDa human sperm antigen described above, the phosphotyrosine residues of FA-1 appear to regulate binding of human zona glycoproteins (Naz and Ahmad, 1994). The sulphhydryl groups of FA-1 may also be important in sperm–zona binding (Naz and Ahmad, 1994). Whether or not FA-1 regulates the tight binding of human gametes is in doubt for several reasons. Firstly, it has not been demonstrated whether FA-1 binds to carbohydrate or protein ligands of the zona. Secondly, human FA-1 readily binds porcine ZP3 and bioneutralizes the boar sperm binding activity of the porcine zona pellucida (Naz et al., 1991b). Thirdly, antibodies to human FA-1 inhibit the binding of acrosomeintact mouse and rabbit spermatozoa to their homologous zonae (Naz et al., 1984; Naz, 1987b, 1990). Fourthly, antibodies to FA-1 markedly reduce the number of human spermatozoa fusing with zona-free hamster ova (Naz et al., 1992). Thus, the exact role of FA-1 in human (mammalian?) fertilization remains to be defined. Common features of sperm surface zona binding proteins Although it is clear that the sugar binding affinities of the sperm plasma membrane differ among species (Table III; Figure 3), the results obtained from the studies described above indicate that sperm surface zona binding proteins have five features in common. Firstly, irrespective of whether a candidate zona receptor is an integral sperm plasma membrane protein or is only peripherally attached, it is distributed on the sperm head over the acrosome cap. Secondly, the number of candidate zona receptor molecules found in this region is increased in association with sperm capacitation. Thirdly, such molecules exhibit lectin-like binding of carbohydrates: they are not immunoglobulins, they bind ligands in a non-enzymatic fashion, and they exhibit a preference for polyvalent ligands but may or may not agglutinate erythrocytes. Fourthly, candidate zona receptors which are integral membrane proteins exhibit an association with the sperm cytoskeleton. Fifthly, ligand or antibody binding leads to aggregation and clearance from the acrosome cap region. This movement is associated with signal transduction, potentially via tyrosine phosphorylation, and initiation of the acrosome reaction. Conclusions Cell–cell recognition by mammalian gametes clearly involves oligosaccharide modification of conserved proteins. The binding of spermatozoa to egg is a biphasic process of non-specific, loose attachment followed by species-specific tight binding. Species specificity of gamete adhesion may involve derivatization of zonae (both in terms of glycan composition and glycan 626 three-dimensional structure), expression of lectins or enzymes on sperm plasma membranes, or possibly both. Whichever occurs, initial recognition of an egg by spermatozoa appears to initiate a train of changes in the sperm head: aggregation of apical sperm head membrane proteins, translocation of some of these (and possibly other) proteins to the equatorial region of the sperm head, and induction of a calcium influx leading to an acrosome reaction. The chemical effectors of these events are not well characterized. Limited proteolysis of these proteins may be involved in the generation of functional zona binding sites on the sperm surface. Interactions with cytoskeletal actin may facilitate surface protein movement. Tyrosine phosphorylation occurs following zona binding in all species studied thus far. There is some evidence that G-proteins and cAMP are involved in early secondary signalling events initiated by ZP3 binding. Is sperm cytoskeleton involved in the events between the initial adhesion of spermatozoa to egg and secondary binding of spermatozoa to ZP2? Speculation on this question is based on results from immunocytochemistry and a cell-free system in which polymerization and depolymerization of actin was observed (Welch and O’Rand, 1985; Saxena et al., 1986b; Peterson et al., 1990; Spungin et al., 1995; Benoff et al., 1996b; Visonti et al., 1996). Sperm capacitation requires changes in intracellular pH, calcium content and plasma membrane protein expression (including zona binding proteins) and efflux of cholesterol from the plasma membrane. A time-dependent increase in polymerized actin associated with plasma membrane proteins is also observed. The initial zona ligand binding event activates protein clearance from the periacrosomal plasma membrane, G-protein signalling, with subsequent phosphorylation of sperm proteins including phospholipase C (PLC). In a manner similar to that observed in somatic cells, phosphorylation may induce a conformational change in PLC so that it may now be transferred from the cytosol to the actin cytoskeleton. Calcium influx would simultaneously occur. Under conditions of increased calcium, PLC activity would prime membranes for fusion by breakdown of phospholipids to produce the fusogenic compound diacylglycerol and additionally liberate actin-severing proteins. The actin severing-proteins would then induce actin depolymerization, thereby removing the barrier to membrane fusion. Specific fusion of the plasma and outer acrosomal membranes would occur in the presence of local elevations in calcium. Following acrosome exocytosis, sperm proteins such as acrosin could effect secondary, tight binding to the zona pellucida. This provides an explanation of how protein clearance and membrane fusion occur in response to an initial recognition event. This hypothesis is eminently testable. A relatively large number of sperm surface proteins have been implicated in these initial egg recognition events. Over 20 have been reported for the six mammalian species reviewed here. Their characterizations are still very incomplete, and some may be shown to be functionally equivalent by future experiments, or to have multiple functionalities not presently recognized. However, the data available to date does not support the supposition that a single sperm protein, however modified, can be responsible for the various species-specific recognition events of all mammals. No single protein itself Carbohydrates and fertilization: an overview possesses all the functionalities required to induce the acrosome reaction. Up to five proteins per species, distinguishable by molecular size on Western gels, have been shown to block fertilization or zona binding, or to induce acrosome reactions. More likely than a single protein is a hetero-molecular complex of proteins which recognizes the zona in a species-specific manner and induces a co-operative signal for an acrosome reaction. Multiple proteins recognizing different zona carbohydrate structures fixed in a defined spatial relationship to each other by the zona protein framework must provide a wider alphabet for species specificity than can a single lectin or enzyme. Blocking of zona binding by an antibody would disrupt the tertiary structure of such a complex, and would indicate only that the protein possessing the epitope targeted forms part of this complex. Induction of an acrosome reaction by cross-linking of proteins would occur irrespective of whether specific cross-linked proteins were involved in the specieslimited recognition event itself, or in some early internal signalling step within the complex. Both types of experiments must be interpreted cautiously. Lectin activities on the sperm head have been demonstrated in several species and are commonly implicated in somatic cell–cell recognition. These appear likely candidates for the regulation of the initial sperm– egg recognition events, if only because no other function within such a complex could easily be attributed to them. Lectins found in some species are clearly extracellular proteins, and therefore cannot induce intracellular signals without the intervention of transmembrane proteins with intracytoplasmic activities. Such transmembrane proteins have been reported in these same species. Lectins of other species have membranespanning regions and may be able to signal directly with their cytoplasmic tails, but may not be able to create all secondary messages necessary to induce acrosome reactions. Cloning and site-specific mutagenesis experiments in the next few years will identify the functional domains of the sperm surface proteins currently believed to be involved in early egg recognition events. It will be interesting to see if a common group of functionalities can be characterized which all mammalian species possess, and if these functionalities are associated with analogous or distinctly different polypeptide chains in different species. However, as fertilization in vivo occurs in the specialized milieu created in the ampulla of the Fallopian tube, we must be aware of the signals the natural environment can induce in spermatozoa, and the sperm surface remodelling it can cause as we draw conclusions as to the mechanisms of early stages in fertilization. Acknowledgments Supported in part by National Institutes of Health Grant No. ES 06100. Appreciation is expressed to Ian R.Hurley, George W.Cooper, and J.Michael Bedford for stimulating discussions and critical reading of the manuscript, to David L.Rosenfeld, Gerald M.Scholl, and Avner Hershlag, for contribution of clinical IVF results, to Asha Jacob and Leslie O.Goodwin for contribution of unpublished findings on the role of tyrosine phosphorylation in the activation of human sperm Ltype voltage-dependent calcium channels, to Francine S.Mandel for statistical consultations, to Steven M.Schrader and Michael J.Breitenstein for contribution of data concerning the sugar binding affinities of rabbit spermatozoa, to Mary C.Mahony for the gift of macaque spermatozoa, to David L.Keefe and Rudolf Oldenbourg for contribution of polarized light micrograph of the hamster zona pellucida, and to Michele Barcia and Stephanie Canaras for technical assistance. References Aarons, D., Boettger-Tong, H., Holt, G. and Poirier, G.R. (1991) Acrosome reaction induced by immunoaggregation of a proteinase inhibitor bound to the murine sperm head. Mol. Reprod. Dev., 30, 258–264. Aarons, D., Battle, T., Boettger-Tong, H. and Poirier, G.R. (1992) Role of monoclonal antibody J-23 in inducing acrosome reactions in capacitated mouse spermatozoa. J. Reprod. Fertil., 96, 49–59. Abdullah, M. and Kierszenbaum, A.L. (1989) Identification of rat testis galactosyl receptor using antibodies to liver asialoglycoprotein receptor: purification and localization on surfaces of spermatogenic cells and sperm. J. Cell Biol., 108, 367–375. Abdullah, M., Widgren, E.E. and O’Rand, M.G. (1991) A mammalian sperm lectin related to rat hepatocyte lectin-2/3. Mol. Cell. Biochem., 103, 155–161. Aguas, A.P. and Pinto da Silva, P. (1989) Bimodal redistribution of surface transmembrane glycoproteins during Ca21-dependent secretion (acrosome reaction) in boar spermatozoa. J. Cell Sci., 93, 467–479. Ahluwalia, B., Farshori, P., Jamuar, M. et al. (1990) Specific localization of lectins in boar and bull spermatozoa. J. Submicrosc. Cytol. Pathol., 22, 53–62. Ahuja, K.K. (1985) Carbohydrate determinants involved in mammalian fertilization. Am. J. Anat., 174, 207–223. Ahuja, K.K. and Bolwell, G.P. (1983) Probable asymmetry in the organization of components of the hamster zona pellucida. J. Reprod. Fertil., 69, 49–55. Anderson, D.J. and Alexander, N.J. (1983) A new look at antifertility vaccines. Fertil. Steril., 40, 557–571. Appeddu, P.A. and Shur, B.D. (1994) Molecular analysis of cell surface beta1,4-galactosyltransferase function during cell migration. Proc. Natl. Acad. Sci. USA, 91, 2095–2099. Arts, E.G.J.M., Kuiken, J., Jager, S. and Hoekstra, D. (1993) Fusion of artificial membranes with mammalian spermatozoa. Specific involvement of the equatorial segment after the acrosome reaction. Eur. J. Biochem., 217, 1001–1009. Arts, E.G.J.M., Jager, S. and Hoekstra, D. (1994) Evidence for the existence of lipid-diffusion barriers in the equatorial segment of human sperm. Biochem. J., 304, 211–218. Austin, C.R. and Bradin, A.W.H. (1956) Early reactions of the rodent egg to spermatozoa penetration. J. Exp. Biol., 33, 358–365. Ayvaliotis, B., Bronson, R., Rosenfeld, D. and Cooper, G. (1985) Conception rates in couples where autoimmunity to sperm is detected. Fertil. Steril., 43, 739–742. Bagavant, H., Fusi, F.M., Baisch, J. et al. (1997) Immunogenicity and contraceptive potential of a human zona pellucida 3 peptide vaccine. Biol. Reprod., 56, 754–770. Bains, H.K., Sehgal, S. and Bawa, S.R. (1992) Human sperm surface mapping with lectins. Acta Anat., 145, 207–211. Barondes, S.H., Castronovo, V., Cooper, D.N.W. et al. (1994) Galectins: a family of animal β-galactose-binding lectins. Cell, 76, 597–598. Barratt, C.L.R., Whitmarsh, A., Hornby, D.P. et al. (1994) Glycosylation of human recombinant ZP3 is necessary to induce the human acrosome reaction. [Abstr. no. 033] Hum. Reprod., 9 (Suppl.). Bearer, E.L. and Friend, D.S. (1982) Modifications of anionic-lipid domains preceding membrane fusion in guinea pig. J. Cell Biol., 92, 604–615. Bearer, E.L. and Friend, D.S. (1990) Morphology of mammalian sperm membranes during differentiation, maturation, and capacitation. J. Electr. Microsc. Tech., 16, 281–297. Bedford, J.M. (1963) Changes in the electrophoretic properties of rabbit spermatozoa during passage through the epididymis. Nature, 200, 1178– 1180. Bedford, J.M. (1968) Ultrastructural changes in the sperm head during fertilization in the rabbit. Am. J. Anat., 123, 329–357. Bedford, J.M. (1971) Techniques and criteria used in the study of fertilization. In Daniel, J.C. Jr. (ed.), Methods in Mammalian Embryology. W.H.Freeman and Co, San Francisco, pp. 39–63. Bedford, J.M. (1972) An electron microscopic study of sperm penetration into the rabbit egg after natural mating. Am. J. Anat., 133, 213–253. Bedford, J.M. (1977) Sperm/egg interaction: the specificity of human spermatozoa. Anat. Rec., 188, 477–487. 627 S.Benoff Bedford, J.M. (1991) The coevolution of mammalian gametes. In Dunbar, B.S. and O’Rand, M.G. (eds.), A Comparative Overview of Mammalian Fertilization. Plenum Press, New York, pp. 3–35. Bedford, J.M., Moore, H.D.M. and Franklin, L.E. (1979) Signficance of the equatorial segment of the acrosome of the spermatozoon in eutherian mammals. Exp. Cell Res., 119, 119–126. Beebe, S.J., Leyton, L., Burks, D. et al. (1992) Recombinant mouse ZP3 inhibits sperm binding and induces the acrosome reaction. Dev. Biol., 151, 48–54. Benau, D.A. and Storey, B.T. (1987) Characterization of the mouse sperm plasma membrane zona-binding site sensitive to trypsin inhibitors. Biol. Reprod., 36, 282–292. Benau, D.A. and Storey, B.T. (1988) Relationship between two types of mouse sperm surface sites that mediate binding of sperm to the zona pellucida. Biol. Reprod., 39, 235–244. Benau, D.A., McGuire, E.J. and Storey, B.T. (1990) Further characterization of the mouse sperm surface zona-binding site with galactosyltransferase activity. Mol. Reprod. Dev., 25, 393–399. Benoff, S. (1993) Preliminaries to fertilization. The role of cholesterol during capacitation of human spermatozoa. Hum. Reprod., 8, 2001–2008. Benoff, S., Cooper, G., Arnold, L. et al. (1992a) Sperm surface mannosespecific lectin: a molecular marker for capacitation events. [Abstr. no. 435] In Scientific Program and Abstracts of the 39th Annual Meeting of the Society for Gynecologic Investigation, p.326. Benoff, S., Rushbrook, J.I., Hurley, I. et al. (1992b) Demonstration of homology between testis-specific mannose-acceptor – a putative zona recognition lectin and the human macrophage mannose-receptor lectin. [Abstr. no. O26] In Scientific Program and Abstracts of the 48th Annual Meeting of the American Fertility Society, p. S12. Benoff, S., Cooper, G.W., Hurley, I. et al. (1993a) Antisperm antibody binding to human sperm inhibits capacitation induced changes in the levels of plasma membrane sterols. Am. J. Reprod. Immunol., 30, 113–130. Benoff, S., Cooper, G.W., Hurley, I. et al. (1993b) Human sperm fertilizing potential in vitro is correlated with differential expression of a head-specific mannose-ligand receptor. Fertil. Steril., 59, 854–862. Benoff, S., Hurley, I., Cooper, G.W. et al. (1993c) Fertilization potential in vitro is correlated with head-specific mannose-ligand receptor expression, acrosome status and membrane cholesterol content. Hum. Reprod., 8, 2155–2166. Benoff, S., Hurley, I., Cooper, G.W. et al. (1993d) Head-specific mannoseligand receptor expression in human spermatozoa is dependent on capacitation-associated membrane cholesterol loss. Hum. Reprod., 8, 2141–2154. Benoff, S., Rushbrook, J.I., Hurley, I. et al. (1993e) Investigations into the structure of a putative zona recognition molecule on the surface of capacitated human sperm -α mannose receptor lectin. [Abstr. no. S220] In Scientific Program and Abstracts of the 40th Annual Meeting of the Society for Gynecologic Investigation, p. 178. Benoff, S., Rushbrook, J.I., Hurley, I. et al. (1993f) Human sperm surface putative zona recognition molecules are mannose-receptor lectins. [Abstr. no. 7] In Program and Abstracts of the 18th Annual Meeting of the American Society of Andrology, p. P-24. Benoff, S., Barcia, M., Rushbrook, J.I. et al. (1994a) Evidence for a myosin motor regulating movement of receptor proteins in the human sperm plasma membrane. [Abstr. no. O-001] In Scientific Program and Abstracts of the 50th Annual Meeting of the American Fertility Society, p. S1. Benoff, S., Cooper, G.W., Hurley, I. et al. (1994b) The effect of calcium ion channel blockers on sperm fertilization potential. Fertil. Steril., 62, 606–617. Benoff, S., Cooper, G.W., Hurley, I.R. and Hershlag, A. (1995a) Calcium-ion channel blockers and sperm fertilization. Assist. Reprod. Rev., 5, 2–13. Benoff, S., Rushbrook, J.I., Hurley, I.R. et al. (1995b) Co-expression of mannose-ligand and non-nuclear progesterone receptors on motile human sperm identifies an acrosome-reaction inducible subpopulation. Am. J. Reprod. Immunol., 34, 100–115. Benoff, S., Hurley, I.R., Mandel, F.S. et al. (1995c) Use of mannose ligands in IVF screens to mimic zona pellucida-induced acrosome reactions and phosphotyrosine kinase activation. [Abstr. no. 0-078] In Scientific Program and Abstracts of the 51st Annual Meeting of the American Society for Reproductive Medicine, p. S38–S39. Benoff, S., Hurley, I.R., Mandel, F.S. et al. (1995d) Reproductive hazards assay: the effect of acute exposure to metal ions on human sperm mannose receptor expression. [Abstr. no. 0-001] In Scientific Program and Abstracts of the 51st Annual Meeting of the American Society for Reproductive Medicine, p. S1. Benoff, S., Barcia, M, Hurley, I.R. et al. (1996a) Classifications of male factor 628 infertility relevant to IVF insemination strategies using mannose ligands, acrosome status and anti-cytoskeletal antibodies. Hum. Reprod., 11, 1905–1911. Benoff, S., Hurley, I.R., Barcia, M. et al. (1996b) Seminal plasma (SP) zinc (Zn) levels can predict IVF fertilization outcome: regulation of acrosome loss through an actin cytoskeleton. [Abstr. no. 0-083] In Scientific Program and Abstracts of the 52nd Annual Meeting of the American Society for Reproductive Medicine, p.S42–S43. Benoff, S., Hurley, I.R., Barcia, M. et al. (1997a) A potential role for cadmium in the etiology of varicocele-associated infertility. Fertil. Steril., 67, 331–342. Bercegeay, S., Jean, M., Lucas, H. and Barriere, P. (1995) Composition of human zona pellucida as revealed by SDS–PAGE after silver staining. Mol. Reprod. Dev., 41, 355–359. Berger, T., Davis, A., Wardrip, N.J. and Hedrick, J.L. (1989a) Sperm binding to the pig zona pellucida and inhibition by solubilized components of the zona pellucida. J. Reprod. Fertil., 86, 559–565. Berger, T., Turner, K.O., Meizel, S. and Hedrick, J.L. (1989b) Zona pellucidainduced acrosome reaction in boar sperm. Biol. Reprod., 40, 414–442. Berrios, M. and Bedford, J.M. (1979) Oocyte maturation: aberrant post-fusion responses of the rabbit primary oocyte to penetrating spermatozoa. J. Cell Sci., 39, 1–12. Berruti, G. (1988) Calpactin-like proteins in human spermatozoa. Exp. Cell Res., 179, 374–384. Bielfeld, P., Faridi, A., Zaneveld, L.J. and DeJonge, C.J. (1994) The zona pellucida-induced acrosome reaction of human spermatozoa is mediated by protein kinases. Fertil. Steril., 61, 536–541. Bleil, J.D. and Wassarman, P.M. (1980a) Mammalian sperm–egg interaction: identification of a glycoprotein in mouse egg zonae pellucidae possessing receptor activity for sperm. Cell, 20, 873–882. Bleil, J.D. and Wassarman, P.M. (1980b) Synthesis of zona pellucida proteins by denuded and follicle-enclosed mouse oocytes during culture in vitro. Proc. Natl. Acad. Sci. USA, 77, 1029–1033. Bleil, J.D. and Wassarman, P.M. (1980c) Structure and function of the zona pellucida: identification and characterization of the proteins of the mouse oocyte’s zona pellucida. Dev. Biol., 76, 185–202. Bleil, J.D. and Wassarman, P.M. (1983) Sperm–egg interactions in the mouse: sequence of events and induction of the acrosome by a zona pellucida glycoprotein. Dev. Biol., 95, 317–324. Bleil, J.D., Beall, C. and Wassarman, P.M. (1981) Mammalian sperm–egg interaction: fertilization of mouse eggs triggers modification of the major zona pellucida glycoprotein, ZP2. Dev. Biol., 86, 189–197. Bleil, J.D. and Wassarman, P.M. (1986) Autoradiographic visualization of the mouse egg’s sperm receptor bound to sperm. J. Cell Biol., 102, 1363–1371. Bleil, J.D. and Wassarman, P.M. (1988) Galactose at the nonreducing terminus of O-linked oligosaccharides of mouse egg zona pellucida glycoprotein ZP3 is essential for the glycoprotein’s sperm receptor activity. Proc. Natl. Acad. Sci. USA, 85, 6778–6782. Bleil, J.D., Greve, J.M. and Wassarman, P.M. (1988) Identification of a secondary sperm receptor in the mouse egg zona pellucida: role in maintenance of binding of acrosome-reacted sperm to eggs. Dev. Biol., 128, 376–385. Bleil, J.D. and Wassarman, P.M. (1990) Identification of a ZP3-binding protein on acrosome-intact mouse sperm by photoaffinity crosslinking. Proc. Natl. Acad. Sci. USA, 87, 5563–5567. Blithe, D.L. (1990) N-Linked oligosaccharides on free alpha interfere with its ability to combine with human chorionic gonadotropin-beta subunit. J. Biol. Chem., 265, 21951–21956. Blobel, C.P., Wofsberg, T.G., Turck, C.W. et al. (1992) A potential fusion peptide and an integrin ligand binding domain in a protein active in sperm– egg fusion. Nature, 356, 248–252. Boettger, H., Richardson, R., Free, D. et al. (1989) Effect of in vitro incubation on a zona binding site found on murine spermatozoa. J. Exp. Zool., 249, 90–98. Boettger-Tong, H., Aarons, D., Biegler, B. et al. (1992) Competition between zonae pellucidae and a proteinase inhibitor for sperm binding. Biol. Reprod., 47, 716–722. Boettger-Tong, H.L., Aarons, D.J., Biegler, B.E. et al. (1993) Binding of a murine protease inhibitor to the acrosome region of the human sperm head. Mol. Reprod. Dev., 36, 346–353. Bookbinder, L.H., Cheng, A. and Bleil, J.D. (1995) Tissue- and speciesspecific expression of sp56, a mouse sperm fertilization protein. Science, 269, 86–89. Bork, P. and Sander, C. (1992) A large domain common to sperm receptors (Zp2 and Zp3) and TGF-β type III receptor. FEBS Letts., 300, 237–240 Carbohydrates and fertilization: an overview Brandelli, A., Miranda, P.V. and Tezón, J.G. (1994) Participation of glycosylated residues in the human sperm acrosome reaction: possible role of N-acetylglucosaminidase. Biochim. Biophys. Acta, 1220, 299–304. Breitenstein, M.J. and Schrader, S.M. (1996) Mannose binding assay in Dutch belted rabbits. [Abstr. no. 111] In Program and abstracts of the 21st Annual Meeting of the American Society of Andrology, p. P-50. Bronson, R.A. and McLaren, A. (1970) Transfer to the mouse oviduct of eggs with and without the zona pellucida. J. Reprod. Fertil., 22, 129–137. Bronson, R.A., Cooper, G.W. and Rosenfeld, D.L. (1984) Sperm antibodies: their role in infertility. Fertil. Steril., 42, 171–183. Bronson, R.A., Cooper, G.W. and Rosenfeld, D.L. (1982a) Correlation between regional specificity of antisperm antibodies to the spermatozoan surface and complement-mediated sperm immobilization. Am. J. Reprod. Immunol., 2, 222–224. Bronson, R.A., Cooper, G.W. and Rosenfeld, D.L. (1982b) Sperm-specific isoantibodies and autoantibodies inhibit binding of human sperm to the human zona pellucida. Fertil. Steril., 38, 724–729. Bronson, R., Cooper, G., Rosenfeld, D. and Witkin, S. (1984) Detection of spontaneously occurring sperm-directed antibodies in infertile couples by immunobead binding and enzyme-linked immunosorbent assay. Ann. N.Y. Acad. Sci., 438, 504–507. Bronson, R.A. and Fusi, F. (1990) Sperm–oolemmal interaction: role of the Arg-Gly-Asp (RGD) adhesion peptide. Fertil. Steril., 54, 527–529. Brown, J., Cebra-Thomas, J.A., Bleil, J.D. et al. (1989) A premature acrosome reaction is programmed by mouse t haplotypes during sperm differentiation and could play a role in transmission ratio distortion. Development, 106, 769–773. Burks, D.J., Carballada, R., Moore, H.D.M. and Saling, P.M. (1995) Interaction of a tyrosine kinase from human sperm with the zona pellucida at fertilization. Science, 269, 83–86. Busacca, M., Fusi, F., Brigante, C. et al. (1989) Evaluation of antisperm antibodies in infertile couples with immunobead test: prevalence and prognostic value. Acta Eur. Fertil., 20, 77–82. Cahova, M. and Draber, P. (1992) Inhibition of fertilization by a monoclonal antibody recognizing the oligosaccharide sequence GalNAcB1-4GalB1-4 on the mouse zona pellucida. J. Reprod. Immunol., 21, 214–256. Callow, J.A., Callow, M.E. and Evans, L.V. (1985). Fertilization in Fucus. In Metz, B. and Monroy, A. (eds), Biology of Fertilization. Vol. 2. Academic Press, New York, pp. 389–408. Calvete, J.J., Sanz, L., Dostalova, Z. and Topfer-Petersen, E. (1995) Spermadhesins: sperm-coating proteins involved in capacitation and zona pellucida binding. Fertilität, 11, 35–40. Canipari, R. and Strickland, S. (1986) Studies on the hormonal regulation of plasminogen activator production in the rat ovary. Endocrinology, 118, 1652–1659. Cardullo, R.A. and Wolf, D.E. (1995) Distribution and dynamics of mouse sperm surface galactosyltransferase: implications for mammalian fertilization. Biochemistry, 34, 10027–10035. Carraway, K.L. and Hull, S.R. (1989) O-Glycosylation pathway for mucintype glycoproteins. BioEssays, 10, 117–121. Cerezo, J.M.S., Beuno, M.I., Skowronski, B. and Cerezo, A.S. (1982) Immunohistochemical localization of concanavalin A and wheat germ lectin receptors in the normal human spermatozoa. Am. J. Reprod. Immunol., 2, 246–249. Chamberlin, M.E. and Dean, J. (1990) Human homolog of the mouse sperm receptor. Proc. Natl. Acad. Sci. USA, 87, 6014–6018. Cheng, A., Le, T., Palacios, M. et al. (1994) Spermatozoa–egg recognition in the mouse: characterization of sp56, a sperm protein having specific affinity for ZP3. J. Cell Biol., 125, 867–878. Cherr, G.N., Meyers, S.A., Yudin, A.I. et al. (1996) The PH-20 protein in cynomologus macaque spermatozoa: identification of two different forms of hyaluronidase activity. Dev. Biol., 175, 142–153. Childs, R.A., Feizi, T. and Yuen, C.-T. (1990) Differential recognition of core and terminal portions of oligosaccharide ligands by carbohydraterecognition domains of two mannose-binding proteins. J. Biol. Chem., 265, 20770–20777. Clark, G.F., Patankar, M.S., Hinsch, K.D. and Oehninger, S. (1995) New concepts in human sperm–zona pellucida interaction. Hum. Reprod., 10 (Suppl. 1), 31–37. Clark, G.F., Oehninger, S. and Seppälä, M. (1996a) Role for glycoconjugates in cellular communication in the human reproductive system. Mol. Hum. Reprod., 2, 513–517. Clark, G.F., Oehninger, S., Patankar, M.S. et al. (1996b) A role for glycoconjugates in human development: the human feto–embryonic defense system hypothesis. Hum. Reprod., 11, 467–473. Clark, J.M. and Koehler, J.K. (1990) Observations of hamster sperm–egg fusion in freeze-fracture replicas including the use of filipin as a sterol marker. Mol. Reprod. Dev., 27, 351–365. Cohen, J., Inge, K.L., Suzman, M. et al. (1989) Videocinematography of fresh and cryopreserved embryos: a retrospective analysis of embryonic morphology and implantation. Fertil. Steril., 52, 820–827. Cohen, J., Alikani, M., Towbridge, J. and Rosenwaks, Z. (1992) Implantation enhancement by selective assisted hatching using zona drilling of human embryos with poor prognosis. Hum. Reprod., 7, 685–691. Cooper, G.W. and Bedford, J.M. (1971) Acquisition of surface charge by the plasma membrane of mammalian spermatozoa during epididymal maturation. Anat. Rec., 169, 300–301. Cooper, G.W. and Bronson, R.A. (1990) Characteriziation of humoral antibodies reactive with spermatozoa, N-acetyl galactosamine, and a putative blood group antigen in seminal plasma. Fertil. Steril., 53, 888–891. Cooper, G.W., Zhu, J.Z. and Rosenfeld, D.L. (1991) Relationship between antisperm IgG’s in female sera and galactosyl and mannosyl terminated antigens in the human sperm surface. [Abstr. No. O-143] In Scientific Program and Abstracts of the 47th Annual Meeting of the American Fertility Society, p. S61. Cornwall, G.A., Tulsiani, D.R.P. and Orbegin-Crist, M.-C. (1991) Inhibition of the mouse sperm surface α-D-mannosidase inhibits sperm–egg binding in vitro. Biol. Reprod., 44, 913–921. Cowan, A.E., Primakoff, P. and Myles, D.G. (1986) Sperm exocytosis increases the amount of PH-20 antigen on the surface of guinea pig sperm. J. Cell Biol., 103, 1289–1297. Cowan, A.E., Myles, D.G. and Koppel, D.E. (1987) Lateral diffusion of PH20 protein on guinea pig sperm: evidence that barriers to diffusion maintain plasma membrane domains in mammalian spermatozoa. J. Cell Biol., 104, 917–923. Cowan, A.E., Myles, D.G. and Koppel, D.E. (1991) Migration of the guinea pig sperm membrane protein PH-20 from one localized surface domain to another does not occur by a simple diffusion-trapping mechanism. Dev. Biol., 144, 189–198. Cross, N.L. (1996) Human seminal plasma prevents sperm from becoming acrosomally responsive to the agonist, progesterone: cholesterol is the major inhibitor. Biol. Reprod., 54, 138–145. Cross, N.L. and Overstreet, J.W. (1987) Glycoconjugates of the human sperm surface: distribution and alterations that accompany capacitation. Gamete Res., 16, 23–35. Cross, N.L., Morales, P., Overstreet, J.W. and Hanson, W.F. (1986) Two simple methods for detecting acrosome-reacted human sperm. Gamete Res., 15, 213–226. Cruetz, C.E. (1992) The annexins and exocytosis. Science, 258, 924–931. Cullis, P.R. and DeKruijff, B. (1979) Lipid polymorphism and the functional roles of lipids in biological membranes. Biochim. Biophys. Acta, 559, 399–420. Cunningham, D.S., Fulgham, D.L., Rayl, D.L. et al. (1991) Antisperm antibodies to sperm surface antigens in women with genital tract infection. Am. J. Obstet. Gynecol., 164, 791–796. Dahms, N.M., Lobel, P. and Kornfeld, S. (1989) Mannose-6-phosphate receptors and lysosomal enzyme targeting. J. Biol. Chem., 264, 12115– 12118. D’Cruz, O.J., Haas, G.G. Jr. and Lambert, H. (1993) Heterogeneity of human sperm surface antigens identified by indirect immunoprecipitation of antisperm antibody bound to biotinylated sperm. J. Immunol., 151, 1062–1074. Dean, J. (1992) Biology of mammalian fertilization: role of the zona pellucida. J. Clin. Invest., 89, 1055–1059. DeAngelis, P.L. and Glabe, C.G. (1987) Polysaccharide structural features that are critical for the binding of sulfated fucans to bindin, the adhesive protein from sea urchin sperm. J. Biol. Chem., 262, 13946–13942. Dell, A., Morris, H.R., Easton, R.L. et al. (1995) Structural analysis of the oligosaccharides derived from glycodelin, a human glycoprotein with potent immunosuppressive and contraceptive activities. J. Biol. Chem., 270, 24116–24126. Doherty, P., Fruns, M. Seaton, P. et al. (1990) A threshold effect of the major isoforms of NCAM on neurite outgrowth. Nature, 343, 464–466. Dostalova, Z., Calvete, J.J., Sanz, L. and Topfer-Petersen, E. (1994) Quantitation of boar sperm adhesins in accessory sex gland fluids and on the surface of epididymal, ejaculated and capaciated spermatozoa. Biochim. Biophys. Acta, 1200, 48–54. Dostalova, Z., Calvete, J.J. and Topfer-Petersen, E. (1995a) Interaction of non-aggregated boar AWN-1 and AQN-3 with phospholipid matrices. A 629 S.Benoff model for coating of spermadhesions to the sperm surface. Biol. Chem. Hoppe-Seyler, 376, 237–242. Dostalova, Z., Calvete, J.J., Sanz, L. and Topfer-Petersen, E. (1995b) Boar sperm adhesion AWN-1. Oligosaccharide and zona pellucida binding characteristics. Eur. J. Biochem., 230, 329–336. Dravland, J.E., Llanos, M.N., Munn, R.J. and Meizel, S. (1984) Evidence for the involvement of a sperm trypsinlike enzyme in the membrane events of the hamster sperm acrosome reaction. J. Exp. Zool., 232, 117–128. Drickamer, K. (1989) Multiple subfamilies of carbohydrate recognition domains in animal lectins. CIBA Foundation Symp., 145, 45–58. Drickamer, K. (1992) Engineering galactose-binding activity into a C-type mannose-binding protein. Nature, 360, 183–186. Drisdel, R.C., Mack, S.R., Anderson, R.A. and Zaneveld, L.J.D. (1995) Purification and partial characterization of acrosome reaction inhibiting glycoprotein from human seminal plasma. Biol. Reprod., 53, 201–208. Drobnis, E.Z. and Katz, D.F. (1991) Videomicroscopy of mammalian fertilization. In Wassarman, P.M. (ed.), Elements of Mammalian Fertilization. Vol. 1.CRC Press Inc, Boca Raton, USA, pp. 269–300. Drobnis, E.Z. (1993) Capacitation and the acrosome reaction. In Scialli, A.R. and Zinaman, M.J. (eds), Reproductive Toxicology and Infertility. McGrawHill Inc, New York, pp. 77–132. Drobnis, E.Z., Yudin, A.I., Cherr, G.N. and Katz, D.F. (1988) Hamster sperm penetration of the zona pellucida: kinematic analysis and mechanical implications. Dev. Biol., 130, 311–323. Ducibella, T., Dubey, A., Gross, V. et al. (1995) A zona biochemical change and spontaneous cortical granule loss in eggs that fail to fertilize in in vitro fertilization. Fertil. Steril., 64, 1154–1161. Dunbar, B.S., Prasad, S.V. and Timmons, T.M. (1991) Comparative structure and function of mammalian zonae pellucidae. In Dunbar, B.S. and O’Rand, M.G. (eds), A Comparative Overview of Mammalian Fertilization. Plenum Press, New York, pp. 97–114. Duncan, A.E. and Fraser, L.R. (1993) Cyclic AMP-dependent phosphorylation of epididymal mouse sperm proteins during capacitation in vitro: identification of an Mr 95000 phosphotyrosine-containing protein. J. Reprod. Fertil., 97, 287–299. East, I.J., Gulyas, B.J. and Dean, J. (1985) Monoclonal antibodies to the murine zona pellucida protein with sperm receptor activity: effects on fertilization and early development. Dev. Biol., 109, 268–273. Edelman, G.M. and Millette, C.F. (1971) Molecular probes of spermatozoon structures. Proc. Natl. Acad. Sci. USA, 68, 2436–2440. Elbein, A.D. (1991) The role of N-linked oligosaccharides in glycoprotein function. Trends Biotech., 9, 346–352. Endo, Y., Mattei, P., Kopf, G.S. and Schultz, R.M. (1987) Effects of a phorbol ester on mouse eggs: dissociation of sperm receptor activity from acrosome reaction-inducing activity of the mouse zona pellucida protein, ZP3. Dev. Biol., 123, 574–577. Emiliozzi, C. and Fenichel, P. (1997) Protein phosphorylation is associated with capacitation of human sperm in vitro but is not sufficient for its completion. Biol. Reprod., 56, 674–679. Eng, L.A. and Oliphant, G. (1978) Rabbit sperm reversible decapacitation by membrane stabilization with a highly purified glycoprotein from seminal plasma. Biol. Reprod., 19, 1083–1094. Epifano, O. and Dean, J. (1994) Biology and structure of the zona pellucida: a target for immunocontraception. Reprod. Fertil. Dev., 6, 319–330. Erbe, D.V., Wolitzky, B.A., Presta, L.G. et al. (1992) Identification of an Eselectin region critical for carbohydrate recognition and cell adhesion. J. Cell Biol., 119, 215–227. Esaguy, N., Welch, J.E. and O’Rand, M.G. (1988) Ultrastructural mapping of sperm plasma membrane autoantigen before and after the acrosome reaction. Gamete Res., 19, 387–399. Esponda, P. and Bedford, J.M. (1985) Epididymal fluid macromolecules do not act as auto- or alloantigens. J. Androl., 6, 359–364. Familiari, G., Nottola, S.A., Micara, G. et al. (1988) Is the sperm–binding capability of the zona pellucida linked to its surface structure? A scanning electron microscopic study of human in vitro fertilization. J. In Vitro Fertil. Embryo Transfer, 5, 134–143. Fayrer-Hosken, R.A., Caudle, A.B. and Shur, B.D. (1991) Galactosyltransferase activity is restricted to the plasma membranes of equine and bovine sperm. Mol. Reprod. Dev., 28, 74–84. Fazeli, A.R., Stenweg, W., Bevers, M.M. et al. (1993) Development of a sperm zona pellucida binding assay for bull semen. Vet. Rec., 132, 14–16. Fazeli, A., Hage, W.J., Cheng, F.-P. et al. (1997) Acrosome-intact boar spermatozoa initiate binding to the homologous zona pellucida in vitro. Biol. Reprod., 56, 430–438. 630 Fei, W. and Hao, X. (1990) Wheat germ agglutinin (WGA) receptors on human sperm membrane and male infertility. Arch. Androl., 24, 97. Feinberg, J.M., Rainteau, D.P., Kaetzel, M.A. et al. (1991) Differential localization of annexins in ram germ cells: a biochemical and immunocytochemical study. J. Histochem. Cytochem., 39, 955–963. Feng, W., Matzuk, M.M., Mountjoy, K. et al. (1995) The asparagine-linked oligosaccharides of the human chorionic gonadotropin B subunit facilitate correct disulfide bond pairing. J. Biol. Chem., 270, 11851–11859. Feuchter, F.A., Tabet, A., Green, M. and Gresham, D. (1986) Exposure and translocation of cytoskeletal-linked surface glycoproteins during capacitation and the acrosome reaction in human sperm. Anat. Rec., 214, 37a. Fierro, R., Foliguet, B., Grignon, G. et al. (1996) Lectin-binding sites on human sperm during acrosome reaction: modifications judged by electron microscopy/flow cytometry. Arch. Androl., 36, 187–196. Fjallbrant, B. and Nilsson, S. (1977) Decrease of sperm antibody titre in males, and conception after treatment of chronic prostatitis. Int. J. Fertil., 22, 255–261. Florman, H.M. (1994) Sequential focal and global elevations of sperm intracellular Ca21 are initiated by the zona pellucida during acrosomal exocytosis. Dev. Biol., 165, 152–164. Florman, H.M. and Storey, B.T. (1982) Mouse gamete interactions: the zona pellucida is the site of the acrosome reaction leading to fertilization in vitro. Dev. Biol., 91, 121–130. Florman, H.M. and Wassarman, P.M. (1985) O-Linked oligosaccharides of mouse egg ZP3 account for its sperm receptor activity. Cell, 41, 313–324. Florman, H.M. and First, N.L. (1988) Regulation of acrosomal exocytosis: the zona pellucida-induced acrosome reaction of bovine spermatozoa is controlled by extrinsic positive regulatory elements. Dev. Biol., 128, 464–473. Florman, H.M., Bechtol, K.B. and Wassarman, P.M. (1984) Enzymatic dissection of the functions of the mouse egg’s receptor for sperm. Dev. Biol., 106, 243–255. Franken, D.R., Kruger, T.F., Menkveld, R. et al. (1990) Hemizona assay and teratozoospermia: increasing sperm insemination concentrations to enhance zona pellucida binding. Fertil. Steril., 54, 497–503. Fraser, L.R. (1984) Mouse sperm capacitation in vitro involves loss of a surface-associated inhibitory component. J. Reprod. Fertil., 72, 373–384. Friend, D.S., Orci, L., Perrelet, A. and Yanagimachi, R. (1977) Membrane particle changes attending the acrosome reaction in guinea pig spermatozoa. J. Cell Biol., 74, 561–577. Fukada, M.N., Dell, A., Oates, J.E. and Fukada, M. (1985) Embryonal lactosaminoglycan. J. Biol. Chem., 260, 6623–6631. Fusi, F. and Bronson, R.A. (1990) Effects of incubation time in serum and capacitation on spermatozoal reactivity with antisperm antibodies. Fertil. Steril., 54, 887–893. Fusi, F.M. and Bronson, R.A. (1992) Sperm surface fibronectin expression following capacitation. J. Androl., 13, 28–35. Fusi, F.M., Vignali, M., Busacca, M. and Bronson, R.A. (1992) Evidence for an integrin cell adhesion receptor on the oolemma of unfertilized human oocytes. Mol. Reprod. Dev., 31, 215–222. Fusi, F.M., Vignali, M., Gailit, J. and Bronson, R.A. (1993) Mammalian oocytes exhibit specific recognition of the RGD (Asp-Gly-Arg) tripeptide and express oolemmal integrins. Mol. Reprod. Dev., 36, 212–219. Gabriel, L.K., Franken, D.R., van der Horst, G. et al. (1994) Wheat germ agglutinin receptors on human sperm membranes and sperm morphology. Andrologia, 26, 5–8. Gadella, B.M., Colenbrander, B., Van Golde, L.M. and Lopes-Cardozo, M. (1993) Boar seminal vesicles secrete arylsulfatases into seminal plasma: evidence that desulfation of seminolipid occurs only after ejaculation. Biol. Reprod., 48, 483–489. Gahmberg, C.G., Kotovuori, P. and Tontti, E. (1992) Cell surface carbohydrate in cell adhesion. Sperm cells and leukocytes bind to their target cells through specific oligosaccharide ligands. Acta Pathol. Microbiol. Immunol. Scand., 100 (Suppl. 27), 39–52. Gallagher, J.T. (1984) Carbohydrate-binding properties of lectins: a possible approach to lectin nomenclature and classification. Biosci. Rep., 4, 621–632. Gambaryan, A.S., Piskarev, V.E., Yamskov, I.A. et al. (1995) Human influenza virus recognition of sialyloligosaccharides. FEBS Lett., 366, 57–60. Garbers, D.L. (1989) Molecular basis of fertilization. Ann. Rev. Biochem., 58, 719–742. Gerton, G.L. (1985) Biochemical studies of the envelope transformations in Xenopus laevis eggs. In Hedrick, J.L. (ed.), The Molecular and Cellular Biology of Fertilization. Plenum Press, New York, pp. 133–150. Gilbert, B.R., Cooper, G.W. and Rosenfeld, D.L. (1991) Removal of sperm Carbohydrates and fertilization: an overview coating antigens by Percoll gradient centrifugation. [Abstr. no. O-126] In Scientific Program and Abstracts of the 47th Annual Meeting of the American Fertility Society, p. S54. Gillis, G., Peterson, R., Russell, L. et al. (1978) Isolation and characterization of membrane vesicles from human and boar spermatozoa: methods using nitrogen cavitation and ionophore induced vesiculation. Prep. Biochem., 8, 363–378. Glander, H.J., Herrman, K. and Haustein, U.F. (1987) The equatorial fibronectin band (EFB) in human spermatozoa–a diagnostic help for male fertility? Andrologia, 19, 456–459. Glander, H.J., Schaller, J. and Dethloff, J. (1994) Influence of the epididymis on integrins and matrix proteins of human spermatozoa. Acta Chir. Hung., 34, 257–261. Gmachl, M. and Kreil, G. (1993) Bee venom hyaluronidase is homologous to a membrane protein of mammalian sperm. Proc. Natl. Acad. Sci. USA, 90, 3569–3573. Go, K.J. and Wolf, D.P. (1983) The role of sterols in sperm capacitation. Adv. Lipid Res., 20, 317–330. Goluboff, E.T., Mertz, J.R., Tres, L.L. and Kierszenbaum, A.L. (1995) Galactosyl receptor in human testis and sperm is antigenically related to the minor C-type (Ca(21)-dependent) lectin variant of human and rat liver. Mol. Reprod. Dev., 40, 460–466. Gong, X., Dubois, D.H., Miller, D.J. and Shur, B.D. (1995) Activation of a G protein complex by aggregation of beta-1,4,galactosyltransferase on the surface of sperm. Science, 269, 1718–1721. Goochee, C.F., Gramer, M.J., Andersen, D.C. et al. (1991) The oligosaccharides of glycoproteins: bioprocess factors affecting oligosaccharide structure and their effect on glycoprotein properties. Biotechnology, 9, 1347–1355. Goodwin, L.O., Leeds, N.B., Hurley, I. et al. (1997) Isolation and characterization of the primary structure of testis-specific L-type calcium channel: implications for contraception. Mol. Hum. Reprod., 3, 255–268. Greve, J.M. and Wassarman, P.M. (1985) Mouse extracellular coat is a matrix of interconnected filaments possessing a structural repeat. J. Mol. Biol., 181, 253–264. Greve, J.M., Salzmann, G.S., Roller, R.J. and Wassarman, P.M. (1982) Biosynthesis of the major zona pellucida glycoprotein secreted by oocytes during mammalian oogenesis. Cell, 31, 749–759. Gulyas, B.J. (1980) Cortical granules of mammalian eggs. Int. Rev. Cytol., 63, 357–392. Gwatkin, R.B.L., Williams, D.T., Hartmann, J.F. and Kniazuk, M. (1973) The zona reaction of hamster and mouse eggs: production in vitro by a trypsinlike protease from cortical granules. J. Reprod. Fertil., 32, 259–265. Haas, G.G.Jr., DeBault, L.E., D’Cruz, O. and Shuey, R. (1988) The effect of fixatives and/or air-drying on the plasma and acrosomal membranes of human sperm. Fertil. Steril., 50, 487–492. Halberg, D.L., Wager, R.E., Farrell, D.C. et al. (1987) Major and minor forms of the rat liver asialoglycoprotein receptor are independent galactosebinding proteins. J. Biol. Chem., 262, 9828–9838. Han, H.L., Mack, S.R., DeJonge, C. and Zaneveld, L.J.D. (1990) Inhibition of the human sperm acrosome reaction by a high molecular weight factor from human seminal plasma. Fertil. Steril., 54, 1177–1179. Han, K.K. and Martinage, A. (1993) Post-translational chemical modifications of proteins. III. Current developments in analytical procedures for identification and quantitation of post-translational chemically modified amino acid(s) and its derivatives. Int. J. Biochem., 25, 957–970. Hanqing, M., Tai-Ying, Y. and Zhao-Wen, S. (1991) Isolation, characterization, and localization of the zona pellucida binding proteins of boar sperm. Mol. Reprod. Dev., 28, 124–130. Hardy, D.M. and Garbers, D.L. (1994) Species-specific binding of sperm proteins to the extracellular matrix (zona pellucida) of the egg. J. Biol. Chem., 269, 19000–19004. Hardy, D.M. and Garbers, D.L. (1995) A sperm membrane protein that binds in a species-specific manner to the egg extracellular matrix is homologous to von Willebrand factor. J. Biol. Chem., 270, 26025–26028. Harris, J.D., Hibler, D.W., Fonteno, G.K. et al. (1994) Cloning and characteriziation of zona pellucida genes and cDNAs from a variety of mammalian species: the ZPA, ZPB and ZPC gene families. DNA Sequence, 4, 361–393. Hedrick, J.L. and Wardrip, N.J. (1986) Isolation of the zona pellucida and purification of its glycoprotein families from pig oocytes. Anal. Biochem., 157, 63–70. Henkel, R., Cooper, S., Kaskar, K. et al. (1995) Influence of elevated pH levels on structural and functional characteristics of the human zona pellucida: functional morphological aspects. J. Assist. Reprod. Genet., 12, 644–649. Hershlag, A., Cooper, G.W. and Benoff, S. (1995) Pregnancy following discontinuation of a calcium channel blocker in the male partner. Hum. Reprod., 10, 599–606. Hershlag, A., Scholl, G., Rosenfeld, D.L. et al. (1996) Mannose ligand receptor assay as a test to predict fertilization in vitro: a prospective study. [Abstr. no. O-079] In Scientific Program and Abstracts of the 52nd Annual Meeting of the American Society for Reproductive Medicine, pp. S40–S41. Hindsgaul, O. and Ballou, C.E. (1984) Affinity purification of mycobacterial polymethyl polysaccharides and a study of polysaccharide-lipid interactions by 1H NMR. Biochemistry, 23, 577–584. Hirano, T., Takasaki, S., Hedrick, J.L. et al. (1993) O-linked neutral sugar chains of porcine zona pellucida glycoproteins. Eur. J. Biochem., 214, 763–769. Hokke, C.H., Damm, J.B.L., Kamerling, J.P. and Vliegenthart, J.F.G. (1993) Structure of three acidic O-linked carbohydrate chains of porcine zona pellucida glycoproteins. FEBS Lett., 329, 29–34. Hokke, C.H., Damm, J.B.L., Penninkhof, B. et al. (1994) Structure of the Olinked carbohydrate chains of porcine zona pellucida glycoproteins. Eur. J. Biochem., 221, 491–512. Holden, C.A., Hyne, R.V., Sathananthan, A.H. and Trounson, A.O. (1990) Assessment of the human sperm acrosome reaction using Concanavalin A lectin. Mol. Reprod. Dev., 25, 247–257. Hoshi, K., Aita, T., Yanagida, K. et al. (1990) Variation in the cholesterol/ phospholipid ratio in human spermatozoa and its relationship with capacitation. Hum. Reprod., 5, 71–75. Howard, S.C., Wittwer, A.J. and Welply, J.K. (1991) Oligosaccharides at each glycosylation site make structure-dependent contributions to biological properties of human tissue plasminogen activator. Glycobiology, 1, 4175– 4180 Huang, T.T.F., Ohzu, E. and Yanagimachi, R. (1982) Evidence suggesting that L-fucose is part of the recognition signal for sperm–zona pellucida attachment in mammals. Gamete Res., 5, 355–361. Huang, T.T.F., Fleming, A.D. and Yanagimachi, R. (1984) Only acrosomereacted spermatozoa can bind to and penetrate zona pellucida: a study of the guinea pig. J. Exp. Zool., 217, 287–290. Humphreys-Beher, M.G. and Blackwell, R.E. (1989) Identification of a deoxyribonucleic acid allelic variant for β1,4 galactosyltransferase expression associated with male sperm binding/penetration infertility. Am. J. Obstet. Gynecol., 160, 1160–1165. Humphreys-Beher, M.G., Garrison, P.W. and Blackwell, R.E. (1990) Detection of antigalactosyltransferase antibodies in plasma from patients with antisperm antibodies. Fertil. Steril., 54, 133–137. Hunnicutt, G.R., Primakoff, P. and Myles, D.G. (1996) Sperm surface protein PH-20 is bifunctional: one activity is a hyaluronidase and a second, distinct activity is required for secondary sperm–zona binding. Biol. Reprod., 55, 80–86. Huszar, G., Sbracia, M., Vigue, L. et al. (1997) Sperm plasma membrane remodeling during spermiogenetic maturation in men: relationship among plasma membrane β1,4-galactosyltransferase, cytoplasmic creatine phosphokinase, and creatine phosphokinase isoform ratios. Biol. Reprod., 56, 1020–1024. Hynes, R. (1985) Molecular biology of fibronectin. Ann. Rev. Cell Biol., 1, 67–90. Iobst, S.T., Wormald, MR., Weis, W.I. et al. (1994a) Binding of sugar ligands to Ca21-dependent animal lectins. I. Analysis of mannose binding by sitedirected mutagenesis and NMR. J. Biol. Chem., 269, 15505–15511. Iobst, S.T. and Drickamer, K. (1994b) Binding of sugar ligands to Ca(21)dependent animal lectins. II. generation of high-affinity galactose binding by site-directed mutagenesis. J. Biol. Chem., 269, 15512–15519. Isenberg, G. (1991) Actin binding proteins – lipid interactions. J. Muscle Res. Cell Motil., 12, 136–144. Isojima,S. (1989) Characterization of epitopes of seminal plasma antigens stimulating human monoclonal antibody sperm–immobilizing antibodies: a personal review. Reprod. Fertil. Dev., 1, 193–204. Ji,I. and Ji,T.H. (1990) Differential interactions of human choriogonadotropin and its antagonistic aglycosylated analog with their receptor. Proc. Natl. Acad. Sci. USA, 87, 4396–4400. Jonakova, V., Sanz, L., Calvete, J.J. et al. (1991) Isolation and biochemical characterization of a zona pellucida-binding glycoprotein of boar sperm. FEBS Lett., 280, 183–186. Jones, R. (1989) Identification of carbohydrate binding proteins in mammalian spermatozoa (human, bull, boar, ram, stallion and hamster) using [125I]fucoidin and [125I]neoglycoprotein probes. Hum. Reprod., 4, 550–557. Jones, R. (1991) Interaction of zona pellucida glycoproteins, sulphated 631 S.Benoff carbohydrates and synthetic polymers with proacrosin, the putative eggbinding protein from mammalian spermatozoa. Development, 111, 1155– 1163. Jones, R., Brown, C.R. and Lancaster, R.T. (1988) Carbohydrate-binding properties of boar sperm proacrosin and assessment of its role in sperm– egg recognition and adhesion during fertilization. Development, 102, 781–792. Jones, R., Shalgi, R., Hoyland, J. and Phillips, D.M. (1990) Topographical rearrangement of a plasma membrane antigen during capacitation of rat spermatozoa in vitro. Dev. Biol., 139, 349–362. Johnson, L.R., Pilder, S.H., Bailey, J. and Olds-Clarke, P. (1995) Sperm from mice carrying one or two t haplotypes are deficient in investment and oocyte penetration. Dev. Biol., 168, 138–149. Kalab, P., Visconti, P., Leclerc, P. and Kopf, G.S. (1994) p95, the major phosphotyrosine-containing protein in mouse spermatozoa, is a hexokinase with unique properties. J. Biol. Chem., 269, 3810–3817. Kallojoki, M., Malmi, R., Virtanen, I. and Suominen, J. (1985) Glycoconjugates of human sperm surface. A study with fluorescent lectins and Lens culinaris agglutination affinity chromatography. Cell Biol. Int. Rep., 9, 151–164. Kaufman, M.H., Fowler, R.E., Barratt, E. and McDougall, R.D. (1989) Ultrastructural and histochemical changes in the murine zona pellucida during the final stages of oocyte maturation prior to ovulation. Gamete Res., 24, 35–48. Kawai, Y., Shimaji, H., Hama, T. and Mayumi, T. (1991) Studies on mouse sperm lectin: the existence and the participation in sperm–egg binding. J. Pharmacobiol-Dyn., 14, 244–249. Keefe, D., Tran, P. and Oldenberg, R. (1996) Polarization microscopy and digital image processing improves non-invasive imaging of oocytes and identifies bilaminar structure of the zona pellucida. J. Soc. Gynecol. Invest., 3 (Suppl.), 182A. Keenan, J.A., Sacco, A.G., Subramanian, M.G. et al. (1991) Endocrine response in rabbits immunized with native versus deglycosylated procine zona pellucida. Biol. Reprod., 44, 150–156. Kery, V., Krepinsky, J.J.F., Warren, C.D. et al. (1992) Ligand recognition by purified human mannose receptor. Arch. Biochem. Biophys., 298, 49–55. Kinloch, R.A., Mortillo, S., Stewart, C.L. and Wassarman, P.M. (1991) Embryonal carcinoma cells transfected with ZP3 genes differentially glycosylate similar polypeptides and secrete active mouse sperm receptor. J. Cell Biol., 115, 655–664. Kinloch, R.A., Sakai, Y. and Wassarman, P.M. (1995) Mapping the mouse ZP3 combining site for sperm by exon mapping and site-directed mutagenesis. Proc. Natl. Acad. Sci. USA, 92, 263–267. Klentzeris, L.D., Fishel, S., McDermott, H. et al. (1995) Positive correlation between expression of beta-1 integrin cell adhesion molecules and fertilizing ability of human spermatozoa in vitro. Mol. Hum. Reprod., 1, see Hum. Reprod. 10, 728–33. Kobata, A. (1992) Structures and functions of the sugar chains of glycoproteins. Eur. J. Biochem., 209, 483–501. Kochibe, N. and Furukawa, K. (1980) Purification and properties of a novel fucose-specific hemagglutinin of Aleuria aurantia. Biochemistry, 19, 2841–2846. Koehler, J.K. (1981) Lectins as probes of the spermatozoon surface. Arch. Androl., 6, 197–217. Kong, M., Richardson, R.T., Widgren, E.E. and O’Rand, MG. (1995) Sequence and localization of the mouse sperm autoantigenic protein, Sp17. Biol. Reprod., 53, 579–590. Kornfeld, R. and Kornfeld, S. (1985) Assembly of asparagine-linked oligosaccharides. Ann. Rev. Biochem., 54, 631–664. Kumar, G.P., Laloraya, M., Agrawal, P. and Laloraya, M.M. (1990) The involvement of surface sugars of mammalian spermatozoa in epididymal maturation and in vitro sperm–zona recognition. Andrologia, 22, 184–194. Kurpisz, M. and Alexander, N.J. (1995) Carbohydrate moieties on the sperm surface: physiological relevance. Fertil. Steril., 63, 158–165. Kurpisz, M., Clark, G.F., Mahony, M. et al. (1989) Mouse monoclonal antibodies against human sperm: evidence for immunodominant glycosylated antigenic sites. Clin. Exp. Immunol., 78, 250–255. Kusan, F.B., Fleming, A.D. and Seidel, G.E. (1984) Successful fertilization in vitro of fresh intact oocytes by perivitelline (acrosome reacted) spermatozoa of the rabbit. Fertil. Steril., 41, 766–770. Lansford, B., Haas, G.G. Jr., Debault, L.E. and Wolf, D.P. (1990) Effect of sperm-associated antibodies on the acrosomal status of human sperm. J. Androl., 11, 532–538. Lassalle, B. and Testart, J. (1994) Human zona pellucida recognition associated with removal of sialic acid from human sperm surface. J. Reprod. Fertil., 101, 703–711. 632 Lathrop, W.F., Carmichael, E.P., Myles, D.G. and Primakoff, P. (1990) cDNA cloning reveals the molecular structure of a sperm surface protein, PH-20, involved in sperm–egg adhesion and the wide distribution of its gene among mammals. J. Cell Biol., 111, 2939–2949. Lee, A.G. (1977) Lipid phase transitions and phase diagrams. I. Lipid phase transitions. Biochim. Biophys. Acta, 472, 237–281. Lee, M.C. and Damjanov, I. (1985) Lectin binding sites on human sperm and spermatogenic cells. Anat. Rec., 212, 282–297. Lee, R.T., Lin, P. and Lee, Y.C. (1984) New synthetic cluster ligands for galactose/N-acetylgalactosamine-specific lectin of mammalian liver. Biochemistry, 23, 4255–4261. Lee, V.H. and Dunbar, B.S. (1993) Developmental expression of the rabbit 55-kDa zona pellucida protein and messenger RNA in ovarian follicles. Dev. Biol., 155, 371–382. Lee, Y.C. (1989) Binding modes of mammalian hepatic Gal/GalNAc receptors. CIBA Foundation Symp., 145, 80–93. Lee, Y.C. (1992) Biochemistry of carbohydrate–protein interaction. FASEB J., 6, 3193–3200. Lee, Y.C., Townsend, R.R., Hardy, M.R. et al. (1983) Binding of synthetic oligosaccharides to the hepatic Gal/GalNAc lectin. Dependence on fine structural features. J. Biol. Chem., 258, 199–202. Leung, A.T., Imagawa, T. and Campbell, K.P. (1987) Structural characterization of the 1,4 dihydropyridine receptor of the voltage-dependent Ca21 channel from rabbit skeletal muscle. Evidence for two distinct high molecular weight subunits. J. Biol. Chem., 262, 7943–7946. Leyton, L. and Saling, P. (1989a) Evidence that aggregation of mouse sperm receptors by ZP3 triggers the acrosome reaction. J. Cell Biol., 108, 2163–2168. Leyton, L. and Saling, P. (1989b) 95 kDa sperm proteins bind ZP3 and serve as tyrosine kinase substrates in response to zona binding. Cell, 57, 1123–1130. Leyton, L., LeGuen, P., Bunch, D. and Saling, P. (1992) Regulation of mouse gamete interaction by a sperm tyrosine kinase. Proc. Natl. Acad. Sci. USA, 89, 11692–11695. Leyton, L., Tomes, C. and Saling, P. (1995) LL95 monoclonal antibody mimics functional effects of ZP3 on mouse sperm: evidence that the antigen recognized is not hexokinase. Mol. Reprod. Dev., 42, 347–358. Liang, L.-F., Chamow, S.M. and Dean, J. (1990) Oocyte-specific expression of mouse Zp-2: developmental regulation of the zona pellucida genes. Mol. Cell. Biol., 10, 1507–1515. Liang, L.-F. and Dean, J. (1993a) Conservation of mammalian secondary sperm receptor genes enables the promoter of the human gene to function in mouse oocytes. Dev. Biol., 156, 399–408. Liang, L.-F. and Dean, J. (1993b) Oocyte development: molecular biology of the zona pellucida. Vitam. Horm., 47, 115–159. Liang, Z.G., O’Hern, P.A., Yavetz, B. et al. (1994) Human cDNAs identified by sera from infertile patients: a molecular biological approach to immunocontraceptive development. Reprod. Fertil. Dev., 6, 297–305. Lin, Y., Kimmel, L.H., Myles, D.G. and Primakoff, P. (1993) Molecular cloning of the human and monkey sperm surface protein PH-20. Proc. Natl. Acad. Sci. USA, 90, 10071–10075. Lin, Y., Mahan, K., Lathrop, W.F. et al. (1994) A hyaluronidase activity of the sperm plasma membrane protein PH-20 enables sperm to penetrate the cumulus cell layer surrounding the egg. J. Cell Biol., 125, 1157–1163. Lingwood, C.A. (1985) Protein–glycolipid interactions during spermatogenesis: binding of specific germ cell proteins to sulfatoxygalactosylacylalkylglycerol, the major glycolipid of mammalian male germ cells. Can. J. Biochem. Cell. Biol., 63, 1077–1085. Lis, H. and Sharon, N. (1993) Protein glycosylation. Structural and functional aspects. Eur. J. Biochem., 218, 1–27. Litscher, E.S., Juntunen, K., Seppo, A. et al. (1995) Oligosaccharide constructs with defined structures that inhibit binding of mouse sperm to unfertilized egg in vitro. Biochemistry, 34, 4662–4669. Litscher, E.S. and Wassarman, P.M. (1996) Characterization of a mouse ZP3derived glycopeptide, gp55, that exhibits sperm receptor and acrosome reaction-inducing activity in vitro. Biochemistry, 35, 3980–3985. Liu, C., Litscher, E.S. and Wassarman, P.M. (1995) Transgenic mice with reduced numbers of functional sperm receptors on their eggs reproduce normally. Mol. Biol. Cell, 6, 577–585. Liu, C., Litscher, E.S., Mortillo, S. et al. (1996) Targeted disruption of the mZP3 gene results in production of eggs lacking a zona pellucida and infertility in female mice. Proc. Natl. Acad. Sci. USA, 93, 5431–5436. Liu, D.Y. and Baker, H.W.G. (1994) Acrosome status and morphology of human spermatozoa bound to the zona pellucida and oolemma determined using oocytes that failed to fertilize in vitro. Hum. Reprod., 9, 673–679. Carbohydrates and fertilization: an overview Liu, M.S., Abersold, R., Fann, C.-H. and Lee, C.-Y.G. (1992) Molecular and developmental studies of sperm acrosome antigen recognized by HS-63 monoclonal antibody. Biol. Reprod., 46, 937–948. Llanos, M., Vigil, P., Salgado, A.M. and Morales, P. (1993) Inhibition of the acrosome reaction by trypsin inhibitors and prevention of penetration of spermatozoa through the human zona pellucida. J. Reprod. Fertil., 97, 173–178. Lopez, L.C. and Shur, B.D. (1987) Redistribution of mouse sperm surface galactosyltransferase after the acrosome reaction. J. Cell Biol., 105, 1663–1670. Lopez, L.C., Youakim, A., Evans, S.C. and Shur, B.D. (1991) Evidence for molecular distinction between Golgi and cell surface forms of B1,4galactosyltransferse. J. Biol. Chem., 266, 15984–15991. Lopez-Casillas, F., Payne, H.M., Andres, J.L. and Massague, J. (1994) Betaglycan can act as a dual modulator of TGF-β access to signaling receptors: mapping of ligand binding and GAG attachment sites. J. Cell Biol., 124, 557–568. Loret de Mola, J.R., Garside, W.T., Bucci, J. et al. (1996) Age and hormonal environment affect the thickness of the human zona pellucida. [Abstr. no. O-036] In Scientific Program and Abstracts of the 52nd Annual Meeting of the Americian Society for Reproductive Medicine, p. S19. Lou, Y., Ang, J., Thai, H. et al. (1995) A ZP3 peptide vaccine induces antibody and reversible infertility without ovarian pathology. J. Immunol., 155, 2515–2720. Lucas, H., Bercegeay, S., LePendu, J. et al. (1994) A fucose-containing epitope involved in gamete interaction on the human zona pellucida. Hum. Reprod., 9, 1532–1538. Lucas, H., Le Pendu, J., Harb, J. et al. (1995) Identification of a spermatozoa L-selectin and of 2 potential zona pellucida ligands. C.R. Acad. Sci. Paris, Life Sci., 318, 795–801. Luconi, M., Krausz, C., Forti, G. and Baldi, E. (1996) Extracellular calcium negatively modulates tyrosine phosphorylation and tyrosine kinase activity during capacitation of human spermatozoa. Biol. Reprod., 55, 207–216. Macek, M.B., Lopez, L.C. and Shur, B.D. (1991) Aggregation of β-1,4galactosyltransferase on mouse sperm induces the acrosome reaction. Dev. Biol., 147, 440–444. Mack, S.R., Everingham, J. and Zaneveld, L.J.D. (1986) Isolation and partial characterization of the plasma membrane from human spermatozoa. J. Exp. Zool., 240, 127–136. Mack, S.R., Zaneveld, L.J.D., Peterson, R.N. et al. (1987) Characterization of human sperm plasma membrane: glycolipids and polypeptides. J. Exp. Zool., 243, 339–346. Mahi-Brown, C.A., Yanagimachi, R., Hoffman, J.C. and Huang, T.T.F. (1985) Fertility control in the bitch by active immunization with porcine zona pellucida: use of different adjuvants and patterns of estradiol and progesterone levels in estrous cycles. Biol. Reprod., 32, 761–772. Mahony, M.C., Oehninger, S., Clark, G.F. et al. (1991) Fucoidin inhibits the zona pellucida-induced acrosome reaction in human spermatozoa. Contraception, 44, 657–665. Malmi, R., Kallojoki, M. and Suominen, J. (1987) Distribution of glycoconjugates in human testis. A histochemical study using fluoresceinand rhodamine-conjugated lectins. Andrologia, 19, 322–332. Margalioth, E.J., Cooper, G.W., Taney, F.H. et al. (1992) Capacitated sperm cells react with different types of antisperm antibodies than fresh ejaculated sperm. Fertil. Steril., 57, 393–398. Marshall, R.D. (1972) Glycoproteins. Ann. Rev. Biochem., 41, 673–702. Matzuk, M.M., Keene, J.L. and Boime, I. (1989) Site specificity of the chorionic gonadotropin N-linked oligosaccharides in signal transduction. J. Biol. Chem., 264, 2409–2414. Maymon, B.B.-S., Maymon, R., Ben-Nun, I. et al. (1994) Distribution of carbohydrates in the zona pellucida of human oocytes. J. Reprod. Fertil., 102, 81–86. McEver, R.P. (1994) Selectins. Curr. Opin. Immunol., 6, 75–84. Mertz, J.R., Banda, P.W. annd Kierszenbaum, A.L. (1995) Rat sperm galactosyl receptor: purification and identification by polyclonal antibodies raised against multiple antigen peptides. Mol. Reprod. Dev., 41, 374–383. Millar, S.E., Chamow, S.M., Baur, A.W. et al. (1989) Vaccination with a synthetic zona pellucida peptide produces long-term contraception in female mice. Science, 246, 935–938. Miller, D.J. and Ax, R.L. (1990) Carbohydrates and fertilization. Mol. Reprod. Dev., 26, 184–198. Miller, D.J., Macek, M.B. and Shur, B.D. (1992) Complementarity between sperm surface β-1,4-galactosyltransferase and egg-coat ZP3 mediates sperm–egg binding. Nature, 357, 589–593. Miller, D.J., Gong, X., Decker, G. and Shur, B.D. (1993) Egg cortical granule N-acetylglucosaminidase is required for the mouse zona block to polyspermy. J. Cell Biol., 123, 1431–1440. Mladenovic, I., Hajdukovic, L., Genbacev, O. et al. (1993) Lectin binding as a biological test in vitro for the prediction of functional activity of human spermatozoa. Hum. Reprod., 8, 258–265. Modlinski, J.A. (1970) The role of the zona pellucida in the development of mouse eggs in vivo. J. Embryol. Exp. Morphol., 23, 539–547. Moller, C.C. and Wassarman, P.M. (1989) Characterization of a proteinase that cleaves zona pellucida glycoprotein ZP2 following activation of mouse eggs. Dev. Biol., 132, 103–112. Monroe, J.R., Altenbern, D.C. and Mathur, S. (1990) Changes in sperm antibody test results when spermatozoa are subjected to capacitating conditions. Fertil. Steril., 54, 1114–1120. Moore, H.D.M. and Bedford, J.M. (1978) Ultrastructure of the equatorial segment of hamster spermatozoa during penetration of oocytes. J. Ultrastruct. Res., 62, 110–117. Moore, H.D.M., Hartman, T.D., Bye, A.P. et al. (1987) Monoclonal antibody against a sperm antigen Mr 95000 inhibits attachment of human spermatozoa to the zona pellucida. J. Reprod. Immunol., 11, 157–166. Moos, J., Kalab, P., Kopf, G.S. and Schiltz, R.M. (1994) Rapid, nonradioactive, and quantitative method to analyze zona pellucida modifications in single mouse eggs. Mol. Reprod. Dev., 38, 91–93. Moos, J., Faundes, D., Kopf, G.S. and Schultz, R.M. (1995) Composition of the human zona pellucida and modifications following fertilization. Mol. Hum. Reprod., 1, see Hum. Reprod., 10, 2467–2471. Morales, P., Cross, N.L. Overstreet, J.W. and Hanson, F.W. (1989) Acrosome intact and acrosome-reacted human sperm can initiate binding to the zona pellucida. Dev. Biol., 133, 385–392. Morales, P., Vigil, P., Franke, D.R. et al. (1994) Sperm–zona interaction: studies on the kinetics of zona pellucida binding and acrosome reaction of human spermatozoa. Andrologia, 26, 131–137. Morgentaler, A., Schopperle, W.M., Crocker, R.H. and DeWolf, W.C. (1990) Protein differences between normal and oligospermic human sperm demonstrated by two-dimensional gel electrophoresis. Fertil. Steril., 54, 902–905. Mori, E., Takasaki, S., Hedrick, J.L. et al. (1991) Neutral oligosaccharide structures linked to asparagines of porcine zona pellucida glycoproteins. Biochemistry, 30, 2078–2087. Mori, K., Daitoh, T., Irahara, M. et al. (1989) Significance of D-mannose as a sperm receptor site on the zona pellucida in human fertilization. Am J. Obstet. Gynecol., 161, 207–211. Mori, K., Daitoh, T., Kamada, M. et al. (1993) Blocking of fertilization by carbohydrates. Hum. Reprod., 8, 1729–1732. Mortillo, S. and Wassarman, P.M. (1991) Differential binding of gold-labeled zona pellucida glycoproteins mZP2 and mZP3 to mouse sperm membrane compartments. Development, 113, 141–149. Mortimer, D., Curtis, E.F. and Camenzind, A.R. (1990) Combined use of fluorescent peanut agglutinin and Hoechst 33258 to monitor the acrosomal status and vitality of human spermatozoa. Hum. Reprod., 5, 99–103. Muramatsu, T., Gachelin, G., Nicolas, J.F. et al. (1978) Carbohydrate structure and cell differentiation: unique properties of fucosyl-glycopeptides from embryonal carcinoma cells. Proc. Natl. Acad. Sci. USA, 75, 2315–2319. Myles, D.G. Hyatt, H. and Primakoff, P. (1987) Binding of both acrosomeintact and acrosome-reacted guinea pig sperm to the zona pellucida during in vitro fertilization. Dev. Biol., 121, 559–567. Myles, D.G., Kimmel, L.H., Blobel, C.P. et al. (1994) Identification of a binding site in the disintegrin domain of fertilin required for sperm–egg fusion. Proc. Natl. Acad. Sci. USA, 91, 4195–4198. Myles, D.G. and Primakoff, P. (1984) Localized surface antigens of guinea pig sperm migrate to new regions prior to fertilization. J. Cell Biol., 99, 1634–1641. Myles, D.G. and Primakoff, P. (1997) Why did the sperm cross the cumulus? To get to the oocyte. Functions of the sperm surface proteins PH-20 and fertilin in arriving at, and fusing with, the egg. Biol. Reprod., 56, 320–327. Naaby-Hansen, S. (1990) Electrophoretic map of acidic and neutral human spermatozoal proteins. J. Reprod. Immunol., 17, 167–185. Nagdas, S.K., Araki, Y., Chayko, C.A. et al. (1994) O-Linked trisaccharide and N-linked polyacetyllactosaminyl glycans are present on mouse ZP2 and ZP3. Biol. Reprod., 51, 262–272. Naito, K. Toyoda, Y. and Yanagimachi, R. (1992) Production of normal mice from oocytes fertilized and developed without zonae pellucidae. Hum. Reprod., 7, 281–285. Naz, R.K. (1987a) Involvement of the fertilization antigen (FA-1) in involuntary immunoinfertility in humans. J. Clin. Invest., 80, 1375–1383. 633 S.Benoff Naz, R.K. (1987b) The fertilization antigen (FA-1) causes a reduction of fertility in actively immunized female rabbits. J. Reprod. Immunol., 11, 117–133. Naz, R.K. (1988) The fertilization antigen (FA-1): applications in immunocontraception and infertility in humans. Am. J. Reprod. Immunol., 16, 21–27. Naz, R.K. (1990) Sperm surface antigens involved in mammalian fertilization. Curr. Opin. Immunol., 2, 48–51. Naz, R.K. and Ahmad, K. (1994) Molecular identities of human sperm proteins that bind human zona pellucida: nature of sperm–zona interaction, tyrosine kinase activity, and involvement of FA-1. Mol. Reprod. Dev., 39, 397–408. Naz, R.K., Ahmad, K. and Kumar, R. (1991a) Role of membrane phosphotyrosine proteins in human spermatozoal function. J. Cell Sci., 99, 157–165. Naz, R.K., Sacco, A.G. and Yurewicz, E.C. (1991b) Human spermatozoal FA1 binds with ZP3 of porcine zona pellucida. J. Reprod. Immunol., 20, 43–58. Naz, R.K., Brazil, C. and Overstreet, J.W. (1992) Effects of antibodies to sperm surface fertilization antigen-1 on human sperm–zona pellcuida interaction. Fertil. Steril., 57, 1304–1310. Naz, R.K., Alexander, N.J., Isahakia, M. and Hamilton, M.S. (1984) Monoclonal antibody against human germ cell glycoprotein that inhibits fertilization. Science, 225, 342–344. Naz, R.K., Morte, C., Garcia-Framis, V. et al. (1993) Characterization of sperm-specific monoclonal antibody and isolation of 95-kilodalton fertilization antigen-2 from human sperm. Biol. Reprod., 49, 1236–1244. Nevo, A.C., Michaeli, I. and Schindler, H. (1961) Electrophoretic properties of bull and of rabbit spermatozoa. Exp. Cell Res., 23, 69–83. Nicolson, G.L. (1974) The interactions of lectins with animal cell surfaces. Int. Rev. Cytol., 39, 89–190. Nicolson, G.L. and Yanagimachi, R. (1972) Terminal saccharides on sperm plasma membranes: identification by specific agglutinins. Science, 177, 276–279. Nicolson, G.L. and Yanagimachi, R. (1974) Mobility and the restriction of mobility of plasma membrane lectin-binding components. Science, 184, 1294–1296. Nicolson, G., Lacorbiere, M. and Yanagimachi, R. (1972) Quantitative determination of plant agglutinin membrane sites on mammalian spermatozoa. Proc. Soc. Exp. Biol. Med., 141, 661–663. Nikas, G., Paraschos, T., Psychoyos, A. and Handyside, A.H. (1994) The zona reaction in human oocytes seen with scanning electron microscopy. Hum. Reprod., 9, 2135–2138. Noguchi, S., Hatanaka, Y., Tobita, T. and Nakano, M. (1992) Structural analysis of the N-linked carbohydrate chains of the 55-kDa glycoprotein family (PZP3) from porcine zona pellucida. Eur. J. Biochem., 204, 1089– 1100. Noguchi, S. and Nakano, M. (1993) Structural characterization of the Nlinked carbohydrate chains from mouse zona pellucida glycoproteins ZP2 and ZP3. Biochim. Biophys. Acta, 1158, 217–226. Ochs, D., Wolf, D.P. and Ochs, R.L. (1986) Intermediate filament proteins in human sperm heads. Exp. Cell Res., 167, 495–504. Oehninger, S., Acosta, A. and Hodgen, G.D. (1990) Antagonistic and agonistic properties of saccharide moieties in the hemizona assay. Fertil. Steril., 53, 143–149. Oehninger, S., Clark, G.F., Acosta, A.A. and Hodgen, G.D. (1991) Nature of the inhibitory effect of complex saccharide moieties on the tight binding of human spermatozoa to the human zona pellucida. Fertil. Steril., 55, 165–169. Oehninger, S., Mahony, M.C., Swanson, J.R. and Hodgen, G.D. (1993) The specificity of human spermatozoa/zona pellucida interaction under hemizona assay conditions. Mol. Reprod. Dev., 35, 57–61. Oehninger, S., Coddington, C.C., Hodgen, G.D. and Seppälä, M. (1995) Factors affecting fertilization: endometrial placental protein 14 reduces the capacity of human spermatozoa to bind to the human zona pellucida. Fertil. Steril., 63, 377–383. Oikawa, T., Nicolson, G.L. and Yanagimachi, R. (1974) Inhibition of hamster fertilization by phytoagglutinins. Exp. Cell Res., 83, 239–246. Olds-Clarke, P., Pilder, P.H., Visconit, P.E. et al. (1996) Sperm from mice carrying two t haplotypes do not possess a tyrosine phosphorylated form of hexokinase. Mol. Reprod. Dev., 43, 94–104. O’Rand, M.G. (1982) Modification of the sperm membrane during capacitation. Ann. N.Y. Acad. Sci., 383, 392–402. O’Rand, M.G. and Fisher, S.J. (1987) Localization of zona pellucida binding sites on rabbit spermatozoa and induction of the acrosome reaction by solublized zonae. Dev. Biol., 119, 551–559. O’Rand,M.G. and Widgren, E.E. (1994) Identification of sperm antigen targets 634 for immunocontraception: B-cell epitope analysis of Sp17. Reprod. Fertil. Dev., 6, 289–296. O’Rand, M.G., Matthews, J.E., Welch, J.E. and Fisher, S.J. (1985) Identification of zona binding proteins of rabbit, pig, human and mouse spermatozoa on nitrocellulose blots. J. Exp. Zool., 235, 423–428. O’Rand, M.G., Widgren, E.E. and Fisher, S.J. (1988) Characterization of the rabbit sperm membrane autoantigen, RSA, as a lectin-like zona binding protein. Dev. Biol., 129, 231–240. O’Rand, M.G., Beavers, J., Widgren, E.E. and Tung, K.S. (1993) Inhibition of fertility in female mice by immunization with a B-cell epitope, the synthetic sperm peptide, P10G. J. Reprod. Immunol., 25, 89–102. Overstreet, J.W. and Bedford, J.M. (1974) Comparison of the penetrability of the egg vestements of follicular oocytes, unfertilized and fertilized ova of the rabbit. Dev. Biol., 41, 185–192. Overstreet, J.W., Lin, Y., Yudin, A.I. et al. (1995) Location of the PH-20 protein on acrosome-intact and acrosome-reacted spermatozoa of cynomologus macaques. Biol. Reprod., 52, 105–114. Papahadjopoulos, D. (1978) Calcium-induced phase changes and fusion in natural and model membranes. In Poste, G. and Nicolson, G.L. (eds), Membrane Fusion. Elsevier/North-Holland Biomedical Press, Amsterdam, pp. 765–790. Parrish, J.J., Susko-Parrish, J.L., Handrow, R.R. et al. (1989) Capacitation of bovine spermatozoa by oviductal fluid. Biol. Reprod., 40, 1020–1025. Patankar, M.S., Oehninger, S., Barnett, T. et al. (1993) A revised structure for fucoidin may explain some of its biological activities. J. Biol. Chem., 268, 21770–21776. Paterson, M. and Aitken, R.J. (1990) Development of vaccines targeting the zona pellucida. Curr. Opin. Immunol., 2, 753–747. Paterson, M., Koothan, P.T., Morris, K.D. et al. (1992) Analysis of the contraceptive potential of antibodies against native and deglycosylated porcine ZP3 in vivo and in vitro. Biol. Reprod., 46, 523–534. Patrizio, P., Bronson, R., Silber, S.J. et al. (1992) Testicular origin of immunobead-reacting antigens in human sperm. Fertil. Steril., 57, 183–186. Peterson, R.N. Russell, L.D., Bundman, D. and Freund, M. (1980) Sperm– egg interactions: evidence for boar sperm plasma membrane receptors for porcine zona pellucida. Science, 207, 73–74. Peterson, R.N., Russell, L.D., Bundman, D. et al. (1981) The interaction of living boar sperm and sperm plasma membrane vesicles with the porcine zona pellucida. Dev. Biol., 84, 144–156. Peterson, R.N., Russell, L.D., Hunt, W. et al. (1983) Characterization of boar sperm plasma membranes by two-dimensional PAGE and isolation of specific groups of polypepetides by anion exchange chromatography and lectin affinity chromatography. J. Androl., 4, 71–81. Peterson, R.N., Henry, L., Hunt, W. et al. (1985) Further characterization of boar sperm plasma membrane proteins with affinity for the porcine zona pellucida. Gamete Res., 12, 91–100. Peterson, R.N., Gillott, M., Hunt, W. and Russell, L.D. (1987) Organization of the boar spermatozoan plasma membrane: Evidence for separate domains (subdomains) of integral membrane proteins in the plasma membrane overlying the principal segment. J. Cell Sci., 88, 343–349. Peterson, R.N. and Hunt, W.P. (1989) Identification, isolation, and properties of a plasma membrane protein involved in the adhesion of boar sperm to the porcine zona pellucida. Gamete Res., 23, 103–118. Peterson, R.N., Bozzola, J.J., Hunt, W.P. and Darabi, A. (1990) Characterization of membrane-associated actin in boar spermatozoa. J. Exp. Zool., 253, 202–214. Peterson, R.N., Campbell, P., Hunt, W.P. and Bozzola, J.J. (1991) Twodimensional polyacrylamide gel electrophoresis characterization of APz, a sperm protein involved in zona binding in the pig and evidence for its binding to specific zona glycoproteins. Mol. Reprod. Dev., 28, 260–271. Phelps, B.M., Primakoff, P., Koppel, D.E. et al. (1988) Restricted lateral diffusion of PH-20, a PI-anchored sperm membrane protein. Science, 240, 1780–1782. Phelps, B.M., Koppel, D.E., Primakoff, P. and Myles, D.G. (1990) Evidence that proteolysis of the surface is an initial step in the mechanism of formation of sperm cell surface domains. J. Cell Biol., 111, 1839–1847. Philpott, C.C., Ringuette, M.J. and Dean, J. (1987) Oocyte-specific expression and developmental regulation of ZP3, the sperm receptor of the mouse zona pellucida. Dev. Biol., 121, 568–575. Pillai, M.C. and Meizel, S. (1991) Trypsin inhibitors prevent the progesteroneinitiated increase in intracellular calcium required for the human sperm acrosome reaction. J. Exp. Zool., 258, 384–393. Poirier, G.R., Robinson, R., Richardson, R. et al. (1986) Evidence for a binding site on the sperm plasma membrane which recognizes the murine zona pellucida. Gamete Res., 14, 235–243. Carbohydrates and fertilization: an overview Pratt, S.A., Scully, N.F. and Shur, B.D. (1993) Cell surface beta 1,4 galactosyltransferase on primary spermatocytes facilitates their initial adhesion to Sertoli cells in vitro. Biol. Reprod., 49, 470–482. Primakoff, P., Hyatt, H. and Myles, D.G. (1985) A role for the migrating sperm surface antigen PH-20 in guinea pig spermatozoa binding to the egg zona pellucida. J. Cell Biol., 101, 2239–2244. Primakoff, P., Cowan, A., Hyatt, H. et al. (1988) Purification of the guinea pig sperm PH-20 antigen and detection of a site-specific endoproteolytic cleavage activity in sperm preparations that cleaves PH-20 into two disulfide-linked fragments. Biol. Reprod., 38, 921–934. Primakoff, P., Lathrop, W. and Bronson, R. (1990) Identification of human sperm surface glycoproteins recognized by autoantisera from immune infertile men, women, and vasectomized men. Biol. Reprod., 42, 929–942. Rankin, T., Familari, M., Lee, E. et al. (1996) Mice homozygous for an insertional mutation in the Zp3 gene lack a zona pellucida and are infertile. Development, 122, 2903–2910. Reddy, J.M., Stark, R.A. and Zaneveld, L.J.D. (1978) A high molecular weight antifertility factor from human seminal plasma. J. Reprod. Fertil., 57, 437–466. Reid, K.B.M., Bentley, D.R., Campbell, R.D. et al. (1986) Complement system proteins which interact with C3b or C4b: a superfamily of structurally related proteins. Immunol. Today, 7, 230–234. Renkonen, O. (1983) Polysaccharides of embryonal carcinoma cells of line PC 13. Biochem. Soc. Tran., 11, 265–267. Rhim, S.H., Millar, S.E., Robey, F. et al. (1992) Autoimmune disease of the ovary induced by a ZP3 peptide from the mouse zona pellucida. J. Clin. Invest., 89, 28–35. Richardson, R., Buckingham, T., Boettger, H. and Poirier, G.R. (1987) A monoclonal antibody to a zona pellucida-proteinase inhibitor binding component on murine spermatozoa. J. Reprod. Immunol., 11, 101–116. Richardson, R.T., Yamasaki, N. and O’Rand, M.G. (1994) Sequence of a rabbit sperm zona pellucida binding protein and localization during the acrosome reaction. Dev. Biol., 165, 688–701. Ringuette, M.J., Sobieski, D.A., Chamow, S.M. and Dean, J. (1986) Oocytespecific gene expression: molecular characterization of a cDNA coding for ZP-3, the sperm receptor of the mouse zona pellucida. Proc. Natl. Acad. Sci. USA, 83, 4341–4345. Ringuette, M.J., Chamberlin, M.E., Baur, A.W. et al. (1988) Molecular analysis of cDNA coding for ZP3, a sperm binding protein of the mouse zona pellucida. Dev. Biol., 127, 287–295. Robertson, R.T., Bhalla, V.K. and Williams, W.L. (1971) Purification and the peptide nature of decapacitation factor. Biochem. Biophys. Res. Commun., 45, 1331–1336. Robinson, R., Richardson, R., Hinds, K. et al. (1987) Features of a seminal plasma proteinase inhibitor-zona pellucida binding component on murine spermatozoa. Gamete Res., 16, 217–228. Rogers, B.J., Bastias, C., Coulson, R.L. and Russell, L.D. (1989) Cytochalasin D inhibits penetration of hamster eggs by guinea pig and human spermatozoa. J. Androl., 10, 275–282. Roller, R.J. and Wassarman, P.M. (1983) Role of asparagine-linked oligosaccharides in secretion of glycoproteins of the mouse egg’s extracellular coat. J. Biol. Chem., 258, 13243–13249. Roller, R.J., Kinloch, R.A., Hiraoka, B.Y. et al. (1989) Gene expression during mammalian oogenesis and early embryogenesis: quantification of three messenger RNAs abundant in fully grown mouse oocytes. Development, 106, 251–261. Rosati, F. (1985) Sperm–egg interactions in Ascidians. In Metz, C.B. and Monroy, A. (eds), Biology of Fertilization. Vol. 2. Academic Press, New York, pp. 361–388. Rosiere, T.K. and Wassarman, P.M. (1992) Identification of a region of mouse zona pellucida glycoprotein mZP3 that possesses sperm receptor activity. Dev. Biol., 154, 309–317. Russell, L.D., Peterson, R.N., Russell, T.A. and Hunt, W.P. (1983) Electrophoretic map of boar sperm plasma membrane polypeptides and localization and fractionation of specific polypeptide subclasses. Biol. Reprod., 28, 393–413. Sacco, A.G. (1977) Antigenic cross reactivity between human and pig zona pellucida. Biol. Reprod., 16, 164–173. Sacco, A.G., Yurewicz, E.C., Subramanian, M.G. and DeMayo, F.J. (1981) Zona pellucida composition: Species cross-reactivity and contraceptive potential of antiserum to a purified pig zona antigen (PPZA). Biol. Reprod., 25, 997–1008. Sacco, A.G., Yurewicz, E.C. and Subramanian, M.G. (1986) Carbohydrate influences the immunogenic and antigenic characteristics of the ZP3 macromolecule (Mr 55000) of the pig zona pellucida. J. Reprod. Fertil., 76, 575–586. Sacco, A.G., Yurewicz, E.C., Subramanian, M.G. and Matzat, P.D. (1989) Porcine zona pellucida: Association of sperm receptor activity with the alpha glycoprotein component of the Mr 5 55,000 family. Biol. Reprod., 41, 523–532. Sadler, J.E. (1984) Biosynthesis of glycoproteins: formation of O-linked oligosaccharides. In Ginsburg, V. and Robbins, P.W. (eds), Biology of Carbohydrates. John Wiley and Sons, New York, pp. 199–288. Sairam, M.R. (1989) Role of carbohydrates in glycoprotein hormone signal transduction. FASEB J., 3, 1915–1926. Sakai, Y., Araki, Y., Yamashita, T. et al. (1988) Inhibition of in vitro fertilization by a monoclonal antibody reacting with the zona pellucida of the oviductal egg but not that of the ovarian egg of the golden hamster. J. Reprod. Immunol., 14, 177–190. Saling, P.M. (1981) Involvement of trypsin-like activity in binding of mouse spermatozoa to zonae pellucidae. Proc. Natl. Acad. Sci. USA, 78, 6231–6235. Saling, P.M. and Storey, B.T. (1979) Mouse gamete interactions during fertilization in vitro. J. Cell Biol., 83, 544–555. Saling, P.M., Sowinski, J. and Storey, B.T. (1979) An ultrastructural study of epididymal mouse spermatozoa binding to zonae pellucidae in vitro: sequential relationship to the acrosome reaction. J. Exp. Zool., 209, 229–238. Salzmann, G.S., Greve, J.M., Roller, R.J. and Wassarman, P.M. (1983) Biosynthesis of the sperm receptor during oogenesis in the mouse. EMBO J., 2, 1451–1456. Sanz, L., Clavete, J.J., Mann, K. et al. (1991) The amino acid sequence of AQN-3, a carbohydrate-binding protein isolated from boar sperm. FEBS Lett., 291, 33–36. Sanz, L., Calvete, J.J., Jonakova, Z. and Topfer-Petersen, E. (1992) Boar spermadhesins AQN-1 and AWN are sperm-associated acrosin inhibitor acceptor proteins. FEBS Lett., 300, 63–66. Sanz, L., Calvete, J.J., Mann, K. et al. (1993) Isolation and biochemical characterization of heparin-binding proteins from boar sperm seminal plasma: a dual role for spermadhesins in fertilization. Mol. Reprod. Dev., 35, 37–43. Sappino, A.P., Huarte, J., Belin, D. and Vassalli, J.D. (1989) Plasminogen activators in tissue remodeling and invasion: mRNA localization in mouse ovaries and implanting embryos. J. Cell Biol., 109, 2471–2479. Sato, M. and Muramatsu, T. (1985) Reactivity of five N-acetylgalactosaminerecognizing lectins with preimplantation embryos, early postimplanatation embryos, and teratocarcinoma cells of the mouse. Differentiation, 29, 29–38. Sathananthan, A.H. and Trounson, A.O. (1982) Ultrastructure of cortical granule release and zona interaction in monospermic and polyspermic human ova fertilized in vitro. Gamete Res., 6, 225–234. Sathananthan, A.H. and Trounson, A.O. (1985) The human pronuclear ovum: fine structure of monospermic and polyspermic fertilization in vitro. Gamete Res., 12, 385–398. Sathananthan, A.H., Trounson, A.O., Wood, C. and Leeton, J.F. (1982) Ultrastructural observations on the penetration of human sperm the zona pellucida of the human egg in vitro. J. Androl., 3, 356–364. Saxena, N., Peteron, R.N., Sharif, S. et al. (1986a) Changes in the organization of surface antigens during in-vitro capacitation of boar spermatozoa as detected by monoclonal antibodies. J. Reprod. Fertil., 78, 601–614. Saxena, N., Peterson, R.N., Saxena, N.K. and Russell, L.D. (1986b) Microfilaments appear in boar spermatozoa during capacitation. J. Exp. Zool., 239, 423–427. Schachner,M. and Martini, R. (1995) Glycans and the modulation of neuralrecognition molecule function. Trends Neurosci., 18, 183–191. Schaller, J., Glander, H.J. and Dethloff, J. (1993) Evidence of β1 integrins and fibronectin on spermatogenic cells in human testis. Hum. Reprod., 8, 1873–1878. Sela, B.-A., Wang, J.L. and Edelman, G.M. (1975) Antibodies reactive with cell surface carbohydrates. Proc. Natl. Acad. Sci. USA, 72, 1127–1131. Sensi, M., Pricci, F., Andreani, D. and Di Mario, U. (1991) Advanced nonenzymatic glycation products (AGE): their relevance to aging and the pathogenesis of late diabetic complications. Diabetes Res., 16, 1–9. Seppo, A., Penttila, L., Niemela, R. et al. (1995) Enzymatic synthesis of octadecameric saccharides of multiply branched blood group I-type, carrying four distal a1,3-galactose or B1,3-GlcNAc residues. Biochemistry, 34, 4655–4661. Shabanowitz, R.B. (1991) Mouse antibodies to human zona pellucida: evidence that human Zp3 is strongly immunogenic and contains two distinct isomer chains. Biol. Reprod., 43, 260–270. Shabanowitz, R.B. and O’Rand, M.G. (1988a) Characterization of the human 635 S.Benoff zona pellucida from fertilized and unfertilized eggs. J. Reprod. Fertil., 82, 151–161. Shabanowitz, R.B. and O’Rand, M.B. (1988b) Molecular changes in the human zona pellucida associated with fertilization and human sperm–zona interactions. Ann. N.Y. Acad. Sci., 541, 621–632. Shalgi, R., Matityahu, A. and Nebel, L. (1986) The role of carbohydrates in sperm–egg interactions in rats. Biol. Reprod., 34, 446–452. Shalgi, R., Maymon, R., Bar-Shira, B. et al. (1991) Distribution of lectin receptor sites in the zona pellucida of follicular and ovulated rat oocytes. Mol. Reprod. Dev., 29, 365–372. Sharon, N. (1993) Lectin–carbohydrate complexes of plants and animals: an atomic view. Trends Biochem. Sci., 18, 221–226. Sharon, N. and Lis, H. (1989) Lectins as cell recognition molecules. Science, 246, 227–246. Shi, Q.-X. and Roldan, E.R.S. (1995) Bicarbonate/CO2 is not required for zona pellucida or progesterone-induced acrosomal exocytosis of mouse spermatozoa but is essential for capacitation. Biol. Reprod., 52, 540–546. Shimizu, S., Tsuji, M. and Dean, J. (1983) In vitro biosynthesis of three sulfated glycoproteins of murine zonae pellucidae by oocytes grown in follicle culture. J. Biol. Chem., 258, 5858–5863. Shur, B.D. (1991) Cell surface β-1,4galactosyltransferase: twenty years later. Glycobiology, 1, 563–575. Shur, B.D. and Bennett, D. (1979) A specific defect in galactosyltransferase regulation on sperm bearing mutant alleles of the T/t locus. Dev. Biol., 71, 243–259. Shur, B.D.and Hall, N.G. (1982a) Sperm surface galactosyltransferase activities during in vitro capacitation. J. Cell Biol., 95, 567–573. Shur, B.D. and Hall, N.G. (1982b) A role for mouse sperm surface galactosyltransferase in sperm binding to the egg zona pellucida. J. Cell Biol., 95, 574–579. Shur, B.D. and Neely, C.A. (1988) Plasma membrane association, purification, and partial characterization of mouse sperm surface β1,4 galactosyltransferase. J. Biol. Chem., 263, 17706–17714. Shur, B.D. and Scully, N.F. (1990) Elevated galactosyltransferase activity on t-bearing sperm segregates with T/t-complex distorter locus. Genet. Res., 55, 177–181. Siegelbaum, S.A. (1994) Ion channel control by tyrosine phosphorylation. Curr. Biol., 4, 242–245. Singer, R., Sagiv, M., Allaouf, D. et al. (1986) Separation of normozoospermic human spermatozoa into a subpopulation by selective agglutination with peanut agglutinin. Andrologia, 18, 17–24. Singer, S.L., Lambert, H., Cross, N.L. and Overstreet, J.W. (1985) Alteration of the human sperm surface during in vitro capacitation as assessed by lectin-induced agglutination. Gamete Res., 12, 291–299. Sim, R.B. and Malhotra, R. (1994) Interactions of carbohydrates and lectins with complement. Biochem. Soc. Transact., 22, 106–111. Sinha, S., Kumar, G.P. and Laloraya, M. (1994) Abnormal physical architecture of the lipophilic domains of human sperm membrane in oligospermia: a logical cause for low fertility profiles. Biochem. Biophys. Res. Commun., 198, 266–273. Skinner, S.M., Mills, T., Kirchick, H.J. and Dunbar, B.S. (1984) Immunization with zona pellucida proteins results in abnormal ovarian follicular differentiation and inhibition of gonadotropin-induced steroid secretion. Endocrinology, 115, 2418–2432. Snell, W.J. and White, J.M. (1996) The molecules of mammalian fertilization. Cell, 85, 629–637. Snow, K. and Ball, G.D. (1992) Characteriziation of human sperm antigens and antisperm antibodies in infertile patients. Fertil. Steril., 58, 1011–1019. Spungin, B., Margalit, I. and Breitbart, H. (1995) Sperm exocytosis reconstructed in a cell-free system: evidence for the involvement of phospholipase C and actin filaments in membrane fusion. J. Cell Sci., 108, 2525–2535. Struck, D.K. and Lennarz, W.J. (1980) The function of saccharide-lipids in synthesis of glycoproteins. In Lennarz, W.J. (ed.), The Biochemistry of Glycoproteins and Proteoglycans. Plenum Press, New York, pp. 35–84. Sullivan,R., Roose,P. and Berube,B. (1989) Immunodetectable galactosyltransferase is associated only with human spermatozoa of high buoyant density. Biochem. Biophys. Res. Commun., 162, 184–188. Summerfield, J.A. and Taylor, M.E. (1986) Mannose-binding proteins in human serum: identification of mannose-specific immunoglobulins and a calcium-dependent lectin, of broader carbohydrate specificity, secreted by hepatocytes. Biochim. Biophys. Acta, 883, 197–203. Suzuki, F. and Yanagimachi, R. (1989) Changes in the distribution of intramembranous particles and filipin-reactive membrane sterols during 636 in vitro capacitation of golden hamster spermatozoa. Gamete Res., 23, 335–347. Suzuki-Toyota, F., Maekawa, M., Cheng, A. and Bleil, J.D. (1995) Immunocolloidal gold labeled surface replica, and its application to detect sp56, the egg recognition and binding protein, on the mouse spermatozoon. J. Electron Microsc., 44, 135–139. Talbot, P. and Chacon, R.S. (1982) Ultrastructural observations on binding and membrane fusion between human sperm and zona pellucida-free hamster oocytes. Fertil. Steril., 37, 240–248. Tanphaichitr, N., Tayabali, A., Gradil, C. et al. (1992) Role for a germ cell-specific sulfolipid-immobilizing protein (SLIP1) in mouse in vivo fertilization. Mol. Reprod. Dev., 32, 17–22. Tanphaichitr, N., Smith, J., Mongkolsirikieart, S. et al. (1993) Role of a gamete-specific sulfoglycolipid immobilizing protein on mouse sperm–egg binding. Dev. Biol., 156, 164–175. Taylor, M.E. and Summerfield, J.A. (1987) Carbohydrate-binding proteins of human serum: isolation of two mannose/fucose-specific lectins. Biochim. Biophys. Acta, 915, 60–67. Taylor, M.E., Bezouska, K. and Drickamer, K. (1992) Contribution to ligand binding by multiple carbohydrate recognition domains in the macrophage mannose receptor. J. Biol. Chem., 267, 1719–1726. Tesarik, J., Drahoud, J. and Peknicova, J. (1988) Subcellular immunochemical localization of acrosin in human spermatozoa during the acrosome reaction and zona pellucida penetration. Fertil. Steril., 50, 133–141. Tesarik, J., Mendoza, C. and Carreras, A. (1991) Expression of D-mannose binding sites on human spermatozoa: comparison of fertile donors and infertile patients. Fertil. Steril., 56, 113–118. Tesarik, J., Mendoza, C. and Carreras, A. (1993a) Fast acrosome reaction measure: a highly sensitive method for evaulating stimulus-induced acrosome reaction. Fertil. Steril., 59, 424–430. Tesarik, J., Mendoza, C., Ramirez, J.P. and Moos, J. (1993b) Solubilized human zona pellucida competes with a fucosylated neoglycoprotein for binding sites on the human sperm surface. Fertil. Steril., 60, 344–350. Thaler, C.D. and Cardullo, R.A. (1995) Biochemical characterization of a glycosylphosphatidylinositol-linked hyaluronidase on mouse sperm. Biochemistry, 34, 7788–7795. Thall, A.D., Malt, P. and Lowe, J.B. (1995) Occyte Gal alpha 1,3 Gal epitopes implicated in sperm adhesion to the zona pellucida glycoprotein ZP3 are not required for fertilization in the mouse. J. Biol. Chem., 270, 21437–21440. Topfer-Petersen, E. and Henschen, A. (1987) Acrosin shows zona and fucose binding, novel properties for a serine protease. FEBS Lett., 226, 38–42. Tsang, R.S., Chan, K.H., Chau, P.Y. et al. (1987) A murine monoclonal antibody specific for the outer core oligosaccharide of Salmonella lipopolysaccharide. Infect. Immunol., 55, 211–216. Tsuiji, A., Hoshiai, H., Takahashi, S. et al. (1986) Sperm–egg interaction observed by scanning electron microscopy. Arch. Androl., 16, 35–47. Tsuji, Y., Clausen, H., Nudelman, E. et al. (1988) Human sperm carbohydrate antigens defined by antisperm human monoclonal antibody derived from an infertile women bearing antisperm antibodies in her serum. J. Exp. Med., 168, 343–356. Tulsiani, D.R., Skudlarek, M.D. and Orgebin-Crist, M.C. (1989) Novel alphaD-mannosidase of rat sperm plasma membranes: characterization and potential role in sperm–egg interactions. J. Cell. Biol., 109, 1257–1267. Tulsiani, D.R., Skudlarek, M.D. and Orgebin-Crist, M.C. (1990) Human sperm plasma membranes possess alpha-D-mannosidase activity but no galactosyltransferase activity. Biol. Reprod., 42, 843–858. Tulsiani, D.R.P., Nagdas, S.K., Cornwall, G.A. and Orgebin-Crist, M.-C. (1992) Evidence for the presence of high mannose/hybrid oligosaccharide chain(s) on the mouse ZP2 and ZP3. Biol. Reprod., 46, 93–100. Tulsiani, D.R.P., NagDas, S.K., Skudlarek, M.D. and Orgebin-Crist, M.C. (1995) Rat sperm plasma membrane mannosidase: localization and evidence for proteolytic processing during epididymal maturation. Dev. Biol., 167, 584–595. Tung, K.S.K. (1975) Human sperm antigens and antisperm antibodies. I. Studies on vasectomy patients. Clin. Exp. Immunol., 20, 93–104. VandeVoort, C.A., Tollner, T.L. and Overstreet, J.W. (1992) Sperm–zona pellucida interaction in cynomologus and rhesus macaques. J. Androl., 13, 428–432. Van Duin, M., Polman, J.E.M., De Breet, I.T.M. et al. (1994) Recombinant human zona pellucida protein ZP3 produced by Chinese hamster ovary cells induces the human sperm acrosome reaction and promotes sperm– egg fusion. Biol. Reprod., 51, 607–617. Varki, A. (1993) Biological roles of oligosaccharides: all of the theories are correct. Glycobiology, 3, 97–130. Carbohydrates and fertilization: an overview Varki, A. (1994) Selectin ligands. Proc. Natl. Acad. Sci. USA, 91, 7390–7397. Vasquez, J.M., Margargee, S.F., Kunze, E. and Hammerstedt, R.H. (1990) Lectin and heparin-binding features of human spermatozoa as analyzed by flow cytometry. Am. J. Obstet. Gynecol., 163, 2006–2012. Villarroya, S. and Scholler, R. (1984) Localization of human sperm antigens with monoclonal antibodies. Path. Biol. Paris, 32, 832–834. Villarroya, S. and Scholler,R. (1986) Regional heterogeneity of human spermatozoa detected with monoclonal antibodies. J. Reprod. Fertil., 76, 435–447. Viriyapanich, P. and Bedford, J.M. (1981) Sperm capacitation in the Fallopian tube of the hamster and its suppression by endocrine factors. J. Exp. Zool., 217, 403–407. Visconti, P.E., Bailey, J.L., Moore, G.D. et al. (1995a) Capacitation of mouse spermatozoa. I. Correlation between the capacitation state and protein tyrosine phosphorylation. Development, 121, 1129–1137. Visconti, P.E., Moore, G.D., Bailey, J.L. et al. (1995b) Capacitation of mouse spermatozoa. II. Protein tyrosine phosphorylation and capacitation are regulated by a cAMP-dependent pathway. Development, 121, 1139–1150. Visconti, P.E., Olds-Clarke, P., Moss, S.B. et al. (1996) Properties and localization of a tyrosine kinase phosphorylated form of hexokinase in mouse sperm. Mol. Reprod. Dev., 43, 82–93. Vuento, M., Kuusela, P., Virkki, M. and Koskimies, A. (1984) Characterization of fibronectin on human spermatozoa. Hoppe-Seylers Z. Physiol. Chem., 365, 757–762. Wakayama, T., Ogura, A., Suto, J. et al. (1996) Penetration by field vole spermatozoa of mouse and hamster zonae pellucidae without acrosome reaction. J. Reprod. Fertil., 107, 97–102. Ward, C.R. and Kopf, G.S. (1993) Molecular events mediating sperm activation. Dev. Biol., 158, 9–34. Ward, C.R., Storey, B.T. and Kopf, G.S. (1994) Selective activation of Gi1 and Gi2 in mouse sperm by the zona pellucida, the egg’s extracellular matrix. J. Biol. Chem., 269, 13254–13258. Wassarman, P.M. (1987) The biology and chemistry of fertilization. Science, 235, 553–560. Wassarman, P.M. (1988) Zona pellucida glycoproteins. Ann. Rev. Biochem., 57, 415–442. Wassarman, P.M. (1989) Role of carbohydrates in receptor-mediated fertilization in mammals. CIBA Foundation Symp., 145, 135–149. Wassarman, P.M. (1990) Profile of a mammalian sperm receptor. Development, 108, 1–17. Wassarman, P.M. (1992) Mouse gamete adhesion molecules. Biol. Reprod., 46, 186–191. Wassarman, P.M., Bleil, J., Florman, H. et al. (1985a) The mouse egg’s receptor for sperm: what is it and how does it work? Cold Spring Harbor Symp. Quant. Biol., 50, 11–19. Wassarman, P.M., Florman, H.M. and Greve, J.M. (1985b) Receptor mediated sperm–egg interactions in mammals. In Metz, C.B. and Monroy, A. (eds), Biology of Fertilization. Vol. 2. Academic Press, New York, pp. 341–360. Wassarman, P.M. and Mortillo, S. (1991) Structure of the mouse egg extracellular coat, the zona pellucida. Int. Rev. Cytol., 130, 85–109. Wassarman, P.M. and Litscher, E.S. (1995) Sperm–egg recognition mechanisms in mammals. Curr. Top. Dev. Biol., 30, 1–19. Watkins, W.M. (1987) Biochemical genetics of blood group antigens: retrospect and prospect. Biochem. Soc. Transact., 15, 620–624. Welch, J.E. and O’Rand, M.G. (1985) Identification and distribution of actin in spermatogenic cells and spermatozoa of the rabbit. Dev. Biol., 109, 411–417. White, D.R., Phillips, D.M. and Bedford, J.M. (1990) Factors affecting the acrosome reaction in human spermatozoa. J. Reprod. Fertil., 90, 71–80. Williams, R.M. and Jones, R. (1993) Specificity of binding of zona pellucida glycoproteins to sperm proacrosin and related proteins. J. Exp. Zool., 266, 65–73. Wilson, I.B.H., Gavel, Y. and Von Heijne, G. (1991) Amino acid distributions around O-linked glycosylation sites. Biochem. J., 275, 529–534. Witkin, S.S. and Toth, A. (1983) Relationship between genital tract infections, sperm antibodies in seminal fluid, and infertility. Fertil. Steril., 40, 805–808. Wittwer, A.J. and Howard, S.C. (1990) Glycosylation at Asn-184 inhibits conversion of single-chain to two-chain tissue-type plasminogen activator by plasmin. Biochemistry, 29, 4175–4180. Wittwer, A.J., Howard, S.C., Carr, L.S. et al. (1989) Effects of N-glycosylation on in vitro activity of Bowes melanoma and human colon fibroblast derived tissue plasminogen activator. Biochemistry, 28, 7662–7669. Wolf, D.E., Lipscomb, A.C. and Maynard, V.M. (1988) Causes of nondiffusing lipid in the plasma membrane of mammalian spermatozoa. Biochemistry, 27, 860–865. Wolf, D.P. (1996) Sperm–egg interaction. In Centola, G.M. and Ginsburg, K.A. (eds), Evaluation and Treatment of the Infertile Male. Cambridge University Press, Cambridge, UK, pp. 6–18. Wollina, U., Schrieber, G., Zollman, C. et al. (1989a) Lectin- binding sites in normal human testis. Andrologia, 21, 127–130. Wollina, U., Schreiber, G., Zollman, C. and Hipler, C. (1989b) Testicular lectin histochemistry in the Sertoli-Cell-only-Syndrome. Andrologia, 21, 555–558. Wright, R.M., John, E., Klotz, K. et al. (1990) Cloning and sequencing of cDNAs encoding for the human intra-acrosome antigen SP-10. Biol. Reprod., 42, 693–701. Xu, C., Rigney, D.R. and Anderson, D.J. (1994) Two-dimensional electrophoretic profile of human sperm membrane proteins. J. Androl., 15, 595–602. Yamasaki, N., Richardson, R.T. and O’Rand, M.G. (1995) Expression of the rabbit sperm protein Sp17 in COS cells and interaction of recombinant Sp17 with the rabbit zona pellucida. Mol. Reprod. Dev., 40, 48–55. Yamashita, K., Kochibe, N., Ohkura, T. et al. (1985) Fractionation of Lfucose-containing oligosaccharides on immobilized Aleuria aurantia lectin. J. Biol. Chem., 260, 4688–4693. Yamashita, K., Totani, K., Ohkura, T. et al. (1987) Carbohydrate binding properties of complex-type oligosaccharides on immobilized Datura stramonium lectin. J. Biol. Chem., 262, 1602–1607. Yamashita, K., Umetsu, K., Suzuki, T. et al. (1988) Carbohydrate binding specificity of immobilized Allomyrina dichotoma lectin II. J. Biol. Chem., 263, 17482–17489. Yanagimachi, R. (1977) Specificity of sperm–egg interaction. In Edidin, M and Johnson, M.H. (eds), Immunobiology of Gametes. Cambridge University Press, Cambridge, UK, pp. 187–207. Yanagimachi, R. (1994) Mammalian fertilization. In Knobil, E. and Neill, J.D. (eds), The Physiology of Reproduction. 2nd edn. Raven Press Ltd, New York, pp. 189–317. Yanagimachi, R, Noda, Y.D., Fujimoto, M. and Nicolson, G.L. (1972) The distribution of negative surface charges on mammalian spermatozoa. Am. J. Anat., 135, 497–520. Yanagimachi, R. and Suzuki, F. (1985) A further study of lysolecithinmediated acrosome reaction of guinea pig spermatozoa. Gamete Res., 11, 29–40. Yednock, T.A. and Rosen, S.D. (1989) Lymphocyte homing. Adv. Immunol., 44, 313–378. Yonezawa, N., Hatanaka, Y., Takeyama, H. and Nakano, M. (1995) Binding of pig sperm receptor in the zona pellucida to the boar sperm acrosome. J. Reprod. Fertil., 103, 1–8. Youakim, A., Hathaway, H.J., Miller, D.J. et al. (1994) Overexpressing sperm surface β1,4-galactosyltransferase in transgenic mice affects multiple aspects of sperm–egg interactions. J. Cell Biol., 126, 1573–1583. Yurewicz, E.C., Sacco, A.G. and Subramanian, M.G. (1987) Structural characterization of the Mr 5 55,000 antigen (ZP3) of porcine oocyte zona pellucida. J. Biol. Chem., 262, 564–571. Yurewicz, E.C., Pack, B.A. and Sacco, A.G. (1991) Isolation, composition, and biological activity of sugar chains of porcine oocyte zona pellucida 55K glycoproteins. Mol. Reprod. Dev., 30, 126–134. Yurewicz, E.C., Pack, B.A. and Sacco, A.G. (1992) Porcine oocyte zona pellucida Mr 55,000 glycoproteins: identification of O-glycosylated domains. Mol. Reprod. Dev., 33, 182–188. Yurewicz, E.C., Pack, B.A., Armant, D.R. and Sacco, A.G. (1993) Porcine zona pellucida ZP3α glycoprotein mediates binding of the biotin-labeled Mr 55,000 family (ZP3) to boar sperm membrane vesicles. Mol. Reprod. Dev., 36, 382–389. Zeng, Y., Oberdorf, J.A. and Florman, H.M. (1996) pH regulation in mouse sperm: identification of Na1-,Cl–, and HCO3-dependent and arylaminobenzoate-dependent regulatory mechanisms and characterization of their roles in sperm capacitation. Dev. Biol., 173, 510–520. Zhang, R., Cai, H., Fatima, N. et al. (1995) Functional glycosylation sites of the rat luteinizing hormone receptor required for ligand binding. J. Biol. Chem., 270, 21722–21728. Zhixing, P. and Yifei, W. (1990) Quantification and localization of wheat germ agglutinin receptor on human sperm membrane and male fertility and infertility. Arch. Androl., 24, 103. Zouari, R., DeAlmeida, M., Rodrigues, D. and Jouannet, P. (1993) Localization of antibodies on spermatozoa and sperm movement characteristics are good predictors of in vitro fertilization success in cases of male autoimmune infertility. Fertil. Steril., 59, 606–612. Received on September 12, 1996; accepted on May 15, 1997 637
© Copyright 2025 Paperzz