Skip to main content
NIHPA Author Manuscripts logoLink to NIHPA Author Manuscripts
. Author manuscript; available in PMC: 2014 Feb 17.
Published in final edited form as: Annu Rev Entomol. 2011;56:21–40. doi: 10.1146/annurev-ento-120709-144823

Insect Seminal Fluid Proteins: Identification and Function

Frank W Avila 1, Laura K Sirot 1, Brooke A LaFlamme 1,*, C Dustin Rubinstein 1,*, Mariana F Wolfner 1,#
PMCID: PMC3925971  NIHMSID: NIHMS433385  PMID: 20868282

Abstract

Seminal fluid proteins (SFPs) produced in reproductive tract tissues of male insects and transferred to females during mating induce numerous physiological and behavioral post-mating changes in females. These changes include decreasing receptivity to re-mating, affecting sperm storage parameters, increasing egg production, modulating sperm competition, feeding behaviors, and mating plug formation. In addition, SFPs also have anti-microbial functions and induce expression of anti-microbial peptides in at least some insects. Here, we review recent identification of insect SFPs and discuss the multiple roles these proteins play in the post-mating processes of female insects.

Keywords: Egg production, mating receptivity, sperm storage, mating plug, sperm competition, feeding, reproduction, Acp

Introduction

In many insect species, mating initiates a behavioral and physiological ‘switch’ in females, triggering responses in several processes related to fertility. Receipt of seminal fluid—a mixture of proteins and other molecules—by the female is a major component of this switch. Insect seminal fluid proteins (SFPs) are the products of male reproductive tract (RT) secretory tissues—accessory glands (AGs), seminal vesicles, ejaculatory duct, ejaculatory bulb and testes. SFPs are transferred to females with sperm during mating. They are major effectors of a wide range of female post-mating responses, including changing female likelihood of re-mating, increasing ovulation and egg-laying rate, changing female flight and feeding behavior, inducing antimicrobial activities, and modulating sperm storage parameters. Absence of SFPs from the ejaculate adversely affects the reproductive success of both sexes. SFPs identified to date represent numerous protein classes, including proteases/protease inhibitors, lectins, prohormones, peptides and protective proteins such as anti-oxidants; these protein classes are present in the ejaculate of organisms from arthropods to mammals (1). While non-protein molecules are also present in seminal fluid (e.g. prostaglandins in crickets: 2; steroid hormones in mosquitoes: 3), research on the effects of seminal fluid receipt has largely focused on the action of SFPs. While the focus of this review is on insect SFPs, progress in the identification and function of SFPs in tick species is also included.

The past few years have witnessed an explosion in the identification and functional analyses of SFPs in insects due to new proteomic and RNA interference technologies. Since earlier results in this field were reviewed by Gillott (4), Chen (5), and Leopold (6), we focus here primarily on recent developments, referring readers to those comprehensive reviews for details on earlier studies.

Dissection of the nature and function of insect SFPs has relevance beyond understanding insect reproductive molecules and their action. SFPs provide intriguing targets for the control of disease vectors and agricultural pests. As we discuss below, SFPs alter reproductive and/or feeding behaviors in a number of arthropods, including insects that cause economic damage or spread disease. In many of these species, there are no approved and effective methods to control the damage they cause. For example, vaccines for tick-borne pathogens have been developed for a limited number of tick antigens (7) and the principal vectors of malaria and dengue fever—the mosquitoes Aedes aegypti and Anopheles gambiae, respectively—have been notoriously difficult to control. Thus, the best current method of limiting these diseases is to control the spread of their insect vectors. In another area, increased resistance to pesticides has made population control by conventional means difficult for pests such as the bollworm/corn earworm, Helicoverpa armigera (8), the bed bug, Cimex lectularius, (9, 10), and ticks (11). As more is learned about the reproductive biology of specific arthropods, their SFPs may provide tools or targets for the control of disease vectors and agricultural pests.

The study of SFPs also provides insight into the evolutionary patterns of reproductive traits. Although the functional classes of SFPs are conserved, a significant fraction of individual SFPs show signs of unusual, often rapid, evolution at the primary sequence level. The forces driving this pattern are not understood, and the study of SFPs may allow for their identification and dissection. Comparative studies of SFPs, individually and in aggregate, are important because 1) lineage-specific SFPs may be involved in the reproductive isolation between species; 2) highly conserved SFPs or SFP classes may be essential for reproduction; and 3) SFP divergence between closely-related species may illuminate selective pressures underlying SFP evolution. Recent reviews have focused on the evolutionary dynamics of SFPs (1215); therefore we will refer the reader to those and focus here on the nature and function of SFPs.

Identification of SFPs

The identification of proteins produced in secretory tissues of the male RT and demonstrated to be, or likely to be, transferred to females during mating is the primary step in SFP identification. Transcriptomic (EST, microarray; 16–27) and proteomic methods (2839) have given a global view of proteins produced in arthtropod male RT glands and, in some cases, of proteins transferred to females during mating (Table 1). For the purposes of this review, proteins within the seminal fluid and transferred to females will be referred to as SFPs. Proteins synthesized in male AGs will be referred to as Acps.

Table 1.

Recently identified SFPs and RT-expressed genes in insects

Order Family Genus sp. Method No. identified
In male RT Transferred
Diptera Drosophilidae Drosophila
simulans
EST screen26, proteomic36,40 5726a 12240, 336
D. melanogaster cumulative review27, EST
screen57, peptide
purification129,150, EST
screen/bioinf. candidatesb23
proteomic33,35,36,40
microarray19,27
11227a, 4619
27, 44033
1335a
13840, 836
1427, 357,129
150
D. yakuba EST screen23, proteomic30,36 11923 10730, 836
D. erecta EST screen23 11423
D. mojavensis EST screen21, proteomic34 5721, 78634
Tephritidae Ceratitis
capitata
EST screen24 1324
Culcidae Aedes aegypti bioinf. candidates/RT-PCR30
proteomic30,32
6330a 5632
Anopheles
gambiae
bioinf. candidates/RT-PCR25
proteomic37
4625, 2037 1537
Lepidoptera Nymphalidae Heliconius erato EST screen/bioinf. filterc/RT-
PCR22, proteomic38
37122 2538
H. melpomene EST screen/bioinf. filter/RT-
PCR22, proteomic38
34022 1038
Hymenoptera Apidae Apis mellifera proteomic29,163 6929 3329, 57163
Coleoptera Tenebrionida
e
Tribolium
castaneum
microarray20 11220
Orthoptera Gryllidae Allonebrius EST screen17 18317
Gryllus firmus EST screen17,18 24717 18
G
pennsylvannicus
EST screen18, proteomic39 27718 2239
Arachnida
Ixodida
Ixodidae Amblyomma
hebraeum
EST screen16 3516
a

Candidate SFPs were based on criteria of accessory gland-specific or enriched expression and/or presence of a predicted secretion signal sequence

b

SFPs or Acps were identified by finding orthologs of identified SFPs or Acps in other organisms

c

Putative SFP or Acp databases were filtered by applying bioinformatic criteria

Most of the Acp/SFP identification studies in Table 1 examined RNA or proteins found in tissues of the male RT, but did not demonstrate SFP transfer during mating. A novel proteomic method directly identified 146 D. melanogaster SFPs (38 previously unannotated), 125 D. simulans SFPs, and 115 D. yakuba SFPs that are transferred to females during mating (30, 36). Findlay et al. (30) fed females a diet enriched in 15N so that the females produced isotopically ‘‘heavy’’ proteins. After these females were mated to unlabeled males, only proteins transferred from males were detected when the female RTs were analyzed by mass spectrometry. This method was subsequently adapted to identify transferred SFPs in Aedes aegypti (Sirot et al. in prep.).

SFPs identified in these studies (Table 1) include peptides and prohormones and protein classes predicted to play roles in numerous functions including sperm binding (lectins and cysteine rich secretory proteins (CRISPs), proteolysis, lipases, and immunity-related functions. These proteins classes are seen in the ejaculates of several insects species, providing evidence that although the primary sequence of some SFPs evolve rapidly, the protein classes represented in seminal fluid are constrained (40). Further, examining the seminal fluid of the extensively studied Drosophila species reveals rapid gain/loss of Acp genes to be a common feature of Drosophila seminal fluid evolution (30, 40) suggesting that Acp genes evolve de novo, perhaps from non-coding DNA (41, 42). However, even as much of the knowledge of insect SFPs has been obtained via studies in Drosophila species, a recent proteomic study showed that Apis mellifera (honeybee) SFPs shared more sequence similarities with human SFPs than with D. melanogaster SFPs (28). Therefore, future studies of SFPs across representative taxonomic groups should shed light on the fascinating evolutionary history of these proteins.

Function of identified SFPs

Historically, insect SFP function has been analyzed by several approaches, including the injection of purified SFP(s) or protein fractions into virgin females, biochemical analysis, removal of putative SFPs by RNAi or mutation in Drosophila, or ectopic expression of SFPs in unmated Drosophila females. Moreover, the increasing availability of genomic and predicted protein annotations has made functional prediction of SFPs by sequence comparison much easier. For example, comparative structural modeling suggested the structure/function of 28 predicted Acps of D. melanogaster males (43). Flybase annotations to D. melanogaster genes were used to predict the functional classes of 240 candidate SFPs in D. mojavensis (34). Cross-species comparisons to D. simulans and D. yakuba led to the identification of 19 D. melanogaster proteins previously unreported as SFPs (30). The newly identified putative SFPs of these species fall into the same categories previously identified in D. melanogaster (43).

Aside from putative function based on sequence analysis, direct assessment of specific SFPs tissue targets within the mated female may hint at those SFPs function (e.g. localization to the sperm storage organs may suggest a role in sperm storage or maintenance, as seen for a network of D. melanogaster Acps; 44). Thirteen D. melanogaster Acps were shown to target to multiple tissues within the mated female RT, each having a unique targeting pattern (45). Additionally, a subset of Acps leave the female RT and enter the hemolympth (4548), potentially reaching nervous and/or endocrine system targets.

More direct methods have also identified the roles of specific SFPs in processes such as the regulation of genes, behaviors and physiological processes such as sperm storage. These results are discussed below.

Transcriptome changes

Changes in female gene expression post-mating have been examined in D. melanogaster and, to a lesser extent, in An. gambiae and Apis mellifera. In D. melanogaster, the role of SFPs and sperm on transcriptional change in mated females has been dissected by microarray analyses. Levels of over 1700 transcripts are altered at 1–3 hrs post-mating in females (49, 50). The mating-dependent genes have predicted functions in a multitude of biological processes including metabolism, immune defense, and protein modification. However, only a handful of the mating-responsive changes in RNA level are greater than 2-fold, consistent with the hypothesis that sexually mature females are “poised” to respond to mating (50, 51). By 6 hrs post-mating, larger-magnitude changes in RNA levels are observed in a smaller number of genes (52). After a second mating, the expression of immunity related genes is more pronounced (53), suggesting that previously mated females have sufficiently up-regulated metabolic and/or structural genes required for post-mating processes (e.g. ovulation and egg laying) to continue.

In the lower RT (the lower common oviduct, seminal receptacle, female accessory glands, spermathecae, and anterior uterus), the levels of over 500 transcripts are changed post-mating (54). A distinct shift—from gene silence to activation—is observed soon after the onset of mating (54). A dramatic peak in differential gene expression is seen at 6 hrs post-mating (54), consistent with the whole-body transcriptome results described above (52). In the oviduct, mating induces an up-regulation of immune-related transcripts and increases levels of RNA for cytoskeleton-related proteins (55). Some oviduct mating-responsive genes respond only to the first mating, while others to both the first and second mating (56). The female RT transcriptome suggests that the structural changes occurring after the first mating (presumably due to mating-responsive gene expression) are sufficient for continued post-mating processes.

Since some of the mating-dependent gene expression change is due to SFP receipt (50), transcriptome change in mates of males lacking specific SFPs was investigated. The ovulation inducing SFP ovulin and the sperm storage protein Acp36DE do not contribute extensively to female transcriptome change at 1–3 hrs post-mating (52). However, two other SFPs, Acp29AB and Acp62F, substantially affect the female transcriptome (52). Surprisingly, Acp29AB and Acp62F contribute to the up-regulation of genes involved in egg production and muscle development, even though analyses of mates to Acp29AB or Acp62F null males do not detect ovulation or egg-laying defects (57, 58).

The sex peptide (SP), a 36 amino acid Acp with roles in egg production, receptivity, feeding, receptivity and sleep behaviors in mated females (5961) see sections below), affects expression of 52 genes in the head and abdomen of mated females. The majority of these RNAs changed only 2–3 fold (62). In the head, SP regulated RNA levels of genes for proteins involved in metabolism, proteolysis, signal transduction and transcription. In abdomens, the SP up-regulated antimicrobial peptide genes via the Toll and IMD pathways (63); a C-terminal motif of SP is responsible for mediating this effect (62). Despite the induction of antimicrobial peptide genes by mating, hemolymph challenge did not detect an immune response in mated females (64, 65).

RNA levels for 141 genes in An. gambiae females experience changes at 2, 6, and 24 hrs post-mating (66), with the number of genes with at least 2-fold change in expression levels increasing with time. Interestingly, changes in transcript levels of many of these genes persist for at least 4 days after mating. Mating responsive expression changes were examined specifically in the head, the gut, the ovaries, the lower RT (tissues below the ovaries), and the two major organs of the lower RT (the atrium, where the ejaculate is received and the spermatheca). In both the gut and the lower RT, several RNAs expressed tissue-specifically change in levels post-mating. Many of these are predicted proteolysis regulators. In the spermatheca, a predicted vitellogenin was also highly up-regulated post-mating. As with D. melanogaster, Rogers et al. (66) conclude that the female atrium is poised to respond to mating. However, they propose that the spermatheca may rely on signals received during mating to regulate genes involved in sperm storage and maintenance.

Large-scale transcriptional changes occur in the ovaries and brains of Apis mellifera queens post-mating (67, 68). In the ovaries, 366 transcripts are differentially expressed post-mating, with the regulated genes largely involved in cell division, gametogenesis, reproduction and oogenesis (67). The RNA levels of 971 genes are differentially expressed in female A. mellifera brains post-mating, with an over-representation of genes involved in protein folding, protein catabolism, and the stress response (67). The types of genes regulated by mating in A.mellifera overlap with those seen in the previously mentioned D. melanogaster studies (e.g. genes involved in the immune response), suggesting that the post-mating transcriptional response may be conserved across species (68). In addition, insemination quantity affects gene expression in the brain (69), suggesting that ejaculate volume and, possibly, quantity of specific SFPs received, may act as a cue for this processes in mated A. mellifera females.

Antimicrobial functions of SFPs

Aside from roles in mediating the up-regulation of antimicrobial genes in mated females (50, 63), some SFPs have intrinsic anti-microbial function. Three D. melanogaster SFPs (from the AG and ejaculatory duct) have antimicrobial activity on E. coli growth in vitro (70). An additional three D. melanogaster SFPs have antimicrobial activity in vivo—able, upon ectopic expression, to reduce bacterial loads in females with S. marcescens (71); the relationship of these three genes to the three identified biochemically is unknown. Although analogous antimicrobial activities have not been detected in the seminal fluid of the bed bug C. lectularius, its seminal fluid does contain bacteriolytic activity, specifically, a lysozyme-like immune activity capable of degrading bacteria (72). These findings suggest that SFPs might play a protective role within the RTs of mated females, possibly aiding females’ ability to clear microbes introduced during mating.

Structural and conformational changes of the female RT

The receipt of seminal fluid induces physiological and structural changes of female RTs. In the oviduct of D. melanogaster females, tissue-wide post-mating changes include the differentiation of cellular junctions, remodeling of the extracellular matrix, increased myofibril formation, and increased innervation of this tissue (55). Post-mating increases in neural activity to the oviduct occurs in the form of vesicle release from RT nerve termini and is modulated distinctly by mating, Acps receipt, and sperm receipt. Mating and/or the receipt of Acps or sperm have differing effects on vesicle release in different regions of the RT at different times post-mating, inferred from the intensity of labeled vesicles (51). Immediate post-mating change in neural vesicle release occurs in the lower common oviduct, seminal receptacle, and uterus. By 3 hrs post-mating—by which time females are ovulating at high rates and egg production has reached maximal levels (73)—vesicle release is inhibited in the common oviduct and lateral oviducts, with Acps modulating changes in nerve termini innervating the seminal receptacle (51).

SFPs receipt also affects the lower RT, inducing a series of conformational changes in the uteri of mated D. melanogaster females, initiating in the first moments of copulation and continuing after mating has ended (74). At least part of this process aids in the storage of sperm, allowing them to access the storage organs. Acps, and not sperm, are the ejaculatory components required to trigger these changes (74). The Acp(s) that initiate this process is unknown. However, Acp36DE is essential for their progression. Incomplete progression of these changes in the absence of Acp36DE leaves sperm lagging in the mid-uterus instead of forming a dense mass adjacent to the sperm storage organ entrances (75). This finding, coupled to the abnormally low numbers of sperm stored in Acp36DE null mates (76, 77), suggests that the post-mating uterine conformation changes aid sperm movement, en masse, toward storage.

An. gambiae female RTs also undergo structural changes upon mating (66). In virgin females, the apical cytoplasm of atrium cells have extensive smooth endoplasmic reticulum surrounded by high numbers of mitochondria. The basal poles of the cells have a high density of rough endoplasmic reticulum. In mated females, both the smooth and the rough endoplasmic reticula mostly disappear, and the mitochondria become distributed throughout the cells. Rogers et al. (66) propose that these structural changes may result in a barrier to re-mating in this species.

Sperm maintenance in, and release from, storage

In addition to roles of SFPs in sperm storage, SFPs are involved in the maintenance of sperm viability in, and their release from, storage (e.g. 78). Seminal fluid secretions from male AGs of both the honeybee A. mellifera and the leafcutter ant Atta colombica promote sperm viability (79, 80). However, AG secretions from one male do not positively affect the viability of another male’s sperm (81). In ants and bees, the effects of AG secretions on sperm survival differ between monandrous and polyandrous species (81): AG secretions from monandrous species promote sperm survival, even when the seminal fluid and sperm are from different males. AG secretions from polyandrous species, however, are detrimental to sperm survival—even to sperm of related males—suggesting a sensitive recognition system exists during sperm competition (81). The negative effects of AG fluid on sperm survival are mitigated by spermathecal secretions in the Atta leafcutter ant, suggesting that females of this species are able to control ejaculate competition once sperm are stored (81). Similarly, D. melanogaster seminal fluid has a protective function, improving the survival of even rival sperm (82).

In. D. melanogaster, Acps are necessary for the efficient utilization of stored sperm, with the few sperm stored in the absence of Acps not used to fertilize eggs (78). Utilization of sperm involves their retention in, and release from, storage (83). The removal of 5 Acps, individually, from the male ejaculate lead to sperm retention in both storage organs after mating (83, Avila et al. in prep). Four of these proteins (CG9997, CG1652, CG1656, CG17575)—a serine protease, 2 C-type lectins, and a CRISP, respectively—are required for the localization of the fifth, SP, to sperm (44), acting in a functional pathway that targets SP to the storage organs in mated females (44). SP, responsible for eliciting numerous post-mating responses is unique in exerting its effects in mated females for several days. SP’s effects persist long-term due to its physically binding sperm, maintaining its presence in the female RT as long as sperm remain in storage (84). Sperm binding is a function of the N-terminus of the peptide; SP’s C-terminus—which contains the receptivity modulating activity—is gradually cleaved from sperm tails (84). These findings suggest that the phenotypes (in sperm storage but also in receptivity and egg laying—see below) elicited by the absence of CG9997, CG1652, CG1656, and CG17575 from the ejaculate may be attributable to the inability of SP to localize to sperm.

The effects of SFPs are not only in terms of sperm release. Acp29AB, a predicted lectin, is needed for sperm to be retained in the sperm storage organs. Sperm from males homozygous for a Acp29AB loss-of-function mutation enter into but are not well maintained within storage, consequently faring poorly in a sperm competitive environment (57). The latter result is likely due to the reduced numbers of stored sperm—a phenotype analogous to that seen in mates of Acp36DE null males (85).

Receptivity to re-mating

Decreased sexual receptivity of mated females occurs in a wide range of insects, and it has been suggested that inducing this change in females is of benefit to males by decreasing the likelihood of sperm competition. In D. melanogaster, the receipt of SFPs change female behavior—mated females actively reject courting males. The SP plays a central role in inducing this change in female receptivity (44, 61, 8689). The four Acps required for SP’s sperm localization also influence receptivity (83). How SP accomplishes its regulation of female receptivity is not known, but its action requires a G-coupled-protein-receptor (87) and specific neurons (88, 89) in females.

C. capitata females are less receptive to male courtship and are less likely to mate for several days after a single mating than are virgin females (90, 91). Sperm storage may play a role in female receptivity in this species as females who store less sperm post-mating are more likely to re-mate sooner (90, 92). Additionally, C. capitata females switch from a male-pheromone odor preference to a host plant odor preference post-mating (91), a switch possibly mediated by factors in the male seminal fluid.

Queensland fruit fly females, Bactrocera tryoni, when mated to irradiated sterile males (and thus subsequently storing little or no sperm) show no difference in sexual receptivity when compared to mates of non-irradiated males (93), suggesting that products of the seminal fluid, and not sperm, are responsible for the reduced post-mating receptivity observed. In support of this hypothesis, virgin B. tryoni females injected with male RT extracts experience diminished sexual receptivity and a shorter copulation duration when subsequently mated, similar to behaviors seen in previously mated females (94). That male AG size decreases after mating in B. tryoni suggests that this tissue is a major site of SFP synthesis (95).

In An. gambiae, male reproductive gland proteins also mediate female likelihood of re-mating (96, 97). This conclusion was initially suggested by studies involving females mated to hybrid males with reduced AGs (96, 97) and subsequently verified by injections of male AG homogenates into virgin females (98).

In several moth species, sexually receptive females produce sex pheromone to attract mates. Sex pheromone production declines substantially after mating: calling behavior cease and oviposition behaviors initiate (99). In several of these species, pheromone production is under neuroendocrine control, resulting from the release of pheromone biosynthesis activating neuropeptide (PBAN) into the female hemolymph (100). Reduction of female pheromone levels post-mating is a consequence of PBAN reduction in the female hemolymph (101). Synthetic D. melanogaster SP and the pheromone suppression peptide HezPSP—from H. zea AGs (102, 103)—suppress pheromone production after injection into unmated H. armigera females. This effect occurs in a dose dependent fashion (104, 105). In addition, antibodies raised against the D. melanogaster SP detect signal from H. armigera male RTs (106).

There is growing evidence that SFPs have important effects on female post-mating behavior in lady bird beetles (107, 108), seed beetles (109111), and ground beetles (112). Injection of testis extracts reduces the probability of mating at 3 hrs and 2 days after injection, whereas injections of AG extracts reduces the probability of mating only at 2 days after injection. Further, injection of a small molecular weight (<3kD) fraction of male RTs result in a short term decrease in the probability of mating (at 1 and 3 hrs after injection); whereas injection of a higher molecular weight (>14kD) fraction results in longer-term inhibition of mating (2 and 4 days after injection; 113). In the ground beetle Leptocarabus procerulus, injections of testes or AG homogenates into virgin females each independently decrease the probability of mating (112).

Egg Production

A frequent effect of seminal fluid receipt is an increase in egg production, ovulation and/or egg laying rates in female insects. Transferring SFPs that up-regulate these processes can benefits males, ensuring their sperm fertilize the maximum number of eggs before the female re-mates, and can also benefit females, allowing them to have increased egg production only when sperm are present to fertilize those eggs. D. melanogaster SP stimulates egg laying in mated females(4, 27), and the long-term persistence of this activity requires the four Acps that localize SP to sperm (44, 83, 84).

The prohormone-like SFP ovulin (Acp26Aa) stimulates ovulation (114). Its mechanism is unknown but could involve two non-mutually exclusive mechanisms: direct ovulin interaction with neuromuscular targets along the lateral oviducts of the female RT, or indirectly by affecting the activity of the neuroendocrine system (51, 114). Ovulin is proteolytically cleaved in the female RT in a step-wise manner (reviewed in 12), a process dependent on at least one other Acp, CG11864 (115), a predicted astacin-family metalloprotease that is itself cleaved in the male RT during transfer to females (115). Ectopic expression of full-length ovulin, or either of ovulin’s two C-terminal cleavage products, is sufficient to stimulate ovulation in unmated females (116), suggesting that ovulin cleavage may increase its activity by generating more bioactive components of the protein.

Little is known about SFP-mediated effects on ovulation in other insect species. In Apis mellifera, mating stimulates vitellogenesis and oocyte maturation in females (68, 117). In Ae. aegypti, male reproductive gland proteins modulate an increase in oviposition (reviewed in 4, 32, 118, 119). In Anopheles sp., there is indirect evidence that SFPs regulate female fecundity (120, 121): males have angiotensin converting enzyme (ACE) activity in their reproductive glands and females mated to males fed ACE inhibitors lay fewer eggs than females mated to control males. In H. armigera, crude extracts of the male AGs stimulate egg maturation and oviposition when injected into virgin females, similar to effects seen after mating (122). The receipt of the male ejaculate increases bed bug C. lectularius female reproductive rates, in terms of lifetime egg production, and females receiving more ejaculate enter reproductive senescence later than females who receive less ejaculate (123), suggesting that ejaculate components may compensate for the costs of elevated reproductive rates by delaying reproductive senescence in this species. Ejaculate volume affects seed beetle fecundity, as females receiving smaller ejaculates have lower fecundity than females receiving larger ejaculates, though this effect is not seen in all beetle species (109, 110). In the ladybird beetle, Adalia bipunctata, females ingest SFPs in the male spermatophore. Females prevented from consuming spermatophores have a longer latency to oviposition as well as a lower duration of resistance to re-mating than control females (108).

Mating plug formation

In several insect species, a mating plug is formed within the female RT during and/or after mating. Mating plugs often contain SFPs, and their formation is dependent on receipt of SFPs. Mating plugs have a wide-range of functions, some involved in sperm competition (124126), the formation of a physical barrier to re-mating, as in butterflies (127), or in switching off female receptivity entirely, as in the bumble bee Bombus terrestris (128).

In D. melanogaster, a mating plug is formed shortly after mating begins. This structure has two major regions: a posterior region comprised of ejaculatory bulb proteins (PEB-me, PEBII and PEBIII; (129, 130), and an anterior region comprised of Acps (129). Evidence for a role of the mating plug, and the SFPs within it, in reducing female receptivity has been shown in D. melanogaster: mates to PEBII knockdown males (who form smaller mating plugs) are more receptive to re-mating than controls in the short-term (4 hrs; 130). These results suggest that the mating plug mediates a short-term decline in receptivity before the long-term effects of other SFPs set in. A similar effect is seen in Drosophila hibisci, where the mating plug inhibits courtship by subsequent males and reduces female receptivity (131). In D. hibisci, the mating plug is also suggested to facilitate sperm storage by preventing the back flow of sperm away from the storage organs (132).

In An. gambiae, the mating plug is necessary for proper sperm storage, but does not prevent re-mating by the female (37). Further, a male AG-specific transglutaminase is necessary to form the mating plug (37). Interestingly, although transglutaminases are made in other mosquito species, male AG-specific transglutaminases are only found in mosquitoes that form mating plugs (37).

In D. mojavensis and related species females experience an “insemination reaction mass” (133). While not a mating plug per se, it fills the entire uterus and persists for hours, absorbing nutrients from the male ejaculate that are incorporated into female somatic tissue (134). Proteins with sequence similarity to larval clotting factors in D. melanogaster (34) found in the AGs of D. mojavensis along with proteins with fibrinogen domains found in D. mayaguana (135) and D. mojavensis (34) AGs, are good candidates for proteins involved in forming the clot-like insemination reaction mass.

Longevity

The longevity of mated females is decreased in some (e.g. D. melanogaster: 136) but not all (e.g. cricket: 137) insects. In D. melanogaster, Acps mediate at least part of this longevity reduction (138), for reasons that are as yet unknown, and recently one SFP, SP, has been shown to play a major role in Acp-mediated decrease in longevity (136). In addition, SP and three other Acps (the protease inhibitors Acp62F and CG8137, and the peptide CG10433) are toxic to D. melanogaster upon ectopic expression (71, 139), possibly reflecting of the negative effect of their action under normal mating conditions. However, the mechanism(s) by which these Acps decrease longevity is unknown and as ectopic expression produces protein levels higher than normally encountered during mating, the toxicity observed may not reflect the true effects of these Acps, suggesting that the longevity effects associated with mating may be an indirect effect of SFP receipt (58).

Feeding

SFPs affect the feeding behavior of some female arthropods. D. melanogaster SP increases female feeding post-mating (59). This behavioral change is substantially reduced in egg-less females and increased in virgin females with experimentally elevated rates of egg production, suggesting that increased feeding is tied to the post-mating increases in ovulation and/or oviposition (140). However, egg-less D. melanogaster females continue to show mating-dependent decreases in life span similar to that of fertile, wild-type females, suggesting that the decreased longevity observed in mated D. melanogaster females is not attributable to over-feeding or to the energetic costs of egg production (140).

In female ticks, the feeding cycle consists of a preparatory phase, a slow feeding phase, and a rapid feeding phase (141). After completion of this cycle, females will have increased in weight almost 100-fold—an engorgement process that lasts ∼6–10 days and completes before females lay an egg batch (141). The transition weight between the slow and rapid feeding phases is termed the ‘critical weight’. Most virgin females do not feed past the critical weight (142, 143). Initiation of the rapid feeding phase is dependent on the receipt of a testis/vas deferens derived engorgement factor called voraxin (144146). Voraxin consists of two components (voraxin α and β) and was shown to be sufficient to stimulate engorgement of feeding when injected into virgin females (146). Additionally, female feeding to engorgement was reduced by 74% when reared on rabbits immunized with recombinant voraxin (146). Paradoxically, RNAi knockdown of voraxin had no effect on female engorgement after mating with knockdown males (147) and experiments in the American dog tick Dermacentor variabilis found that silencing engorgement factor α and β homologs via RNAi failed to reduce engorgement (148). Thus, the feeding role of these proteins has yet to be fully ascertained.

Activity levels

The increase in D. melanogaster female feeding observed post-mating coincides with a decrease in female ‘siesta’ sleep (a quiescent sleep-like state) post-mating (60). This effect is mediated by receipt of SP, which decreases siesta sleep by 70% (60), consequently increasing foraging and egg-laying activity of mated females. In conjunction with the negative impact of SP on female life-span (136), it has been suggested that SP’s effect on female longevity may be the result of increases in stress due to sleep deprivation and to increased locomotor activity (60).

Flight behavior is altered post-mating in honeybee A. mellifera queens. At ∼1–2 weeks of age, queens mate multiply during “mating flights”—inseminated by an average of 12 males (149). Mating makes queens less likely to attempt flight again (68). Additionally, insemination by single vs. multiple drones affects several behaviors, including flight behavior (68), suggesting that queens might use ejaculate volume (and possibly contents) as a cue for flight attempts.

Female effectors of SFPs

Little is known about the female molecules that interact with SFPs and are subsequently responsible for inducing the myriad post-mating changes observed in insects and other arthropods. A notable exception is the receptor for the D. melanogaster SP, the sex peptide receptor (SPR). SPR, identified in an extensive RNAi screen, is a G-protein-coupled-receptor that acts through a cAMP-dependent pathway (87). It should be noted that the ejaculatory duct peptide DUP99B, having a C-terminus similar to that of SP, also interacts with SPR (87, 150). SPR’s expression in neurons that express sex-specific fruitless transcript is necessary and sufficient to re-establish SPR’s receptivity and egg-laying effects (87). Further, SPR expression is necessary in sensory neurons innervating the female RT that express the pickpocket marker, possibly reducing the output of these neurons to the central nervous system (88, 89). The ability of the D. melanogaster SP to interact in vitro with Aedes aegypti and Bombyx mori SPR orthologs (87) suggests that SFPs analogous to the SP are present in the seminal fluid of these, and potentially other, insects. This interpretation is consistent with the ability of D. melanogaster SP to induce post-mating responses when injected into in unmated H. armigera females (104, 106, 151).

Ultimately, SFPs must interact in the context of the female RT. Thus, progress in understanding the signaling mechanisms involved in insect reproductive processes may illuminate mechanisms of SFP modulation in female physiology and behavior. Recent reviews have addressed neuropeptide control of insect hormones and sexual receptivity (152) in the reproductive physiology of the locust Locusta migratoria (e.g. 153). SFPs may act up-stream of traditional neural signaling systems. For example, ovulation and subsequent egg laying are presumably mediated by contraction of the female RT. The biogenic amine octopamine (OA) is an important regulator of ovulation-related contractions of the female RT in the Orthopteran Locusta (154), the muscid fly Stomoxys, (155) and Drosophila (156158). Further, RT extracts from male Stomoxys induce changes in muscle contraction in female RTs (155). The Drosophila receptor, OAMB, critical for the ovulation effect, is selectively expressed in oviductal epithelium (159) . OA mediates ovary muscle contraction and oviduct muscle relaxation. It has been proposed that these opposing effects may serve to expel the egg from the ovary while facilitating entry into the common oviduct (153, 157, 160). Perhaps signaling systems, such as OA, or their proximate downstream targets may serve as substrates that are modulated by SFPs.

Social behavior effects

The amount of SFPs transferred may depend, in part, on the mating status of males and females and their social environment before or during mating. Mating status of both sexes can affect the magnitude of female post-mating responses (161163). In a number of insects, females mated to recently-mated males show less pronounced post-mating changes in receptivity and egg production than do females mated to virgin males (e.g., D. melanogaster: 162; Anastrepha obliqua: 163). In some insects, mates of nutritionally-stressed males have less pronounced post-mating changes in receptivity to re-mating than mates of control males (164). These studies suggest that males are limited in the amount of SFPs they can produce and/or store at a given time and that SFP production may be resource-limited.

Given this potential limitation and the importance of SFPs in determining male reproductive success (e.g., via effects on sperm storage, egg production, and re-mating), selective pressures should exist for males to allocate the ejaculate in a manner that maximizes their reproductive success. One way this could be accomplished would be to allocate more SFPs to females mated under conditions of higher sperm competition risk (elevated either because the female has previously mated or because other males are in the vicinity of the mating pair). There is support for such “strategic allocation” of sperm in a number of insect species, including beetles, crickets, and medflies (reviewed in 165). Recent evidence has demonstrated strategic allocation of SFPs as well (reviewed 12). Briefly, male D. melanogaster transfer more sex peptide when they are exposed to another male before and during mating than when they are alone with the female before and during mating (166) . Other evidence is consistent with the hypothesis that strategic SFP allocation increases male reproductive success. For example, the mates of males exposed to other males before mating have longer latencies to re-mating and higher fecundity than mates of males not exposed to other males before mating (167). Thus, D. melanogaster males are able to adjust their ejaculate composition in response to risk of sperm competition, an adjustment that appears to increase male reproductive success. Future research in this area should test for strategic SFP allocation in other insect species.

Conclusions

SFPs have roles in modulating many female behavioral and physiological processes across a wide range of insect species. The recent rapid pace of technological advances in transcript and protein identification has resulted in greatly increased knowledge of suites of SFPs in a number of insect species, and roles of individual SFPs in female post-mating responses are being elucidated. However, several questions still need to be addressed.

First, how do male SFPs interact with each other and with female molecules to effect the changes observed in mated females? Downstream female effectors with, or through/which SFPs exert their functions remain unknown, with the exception of D. melanogaster SP and its receptor SPR. Proteins secreted from female RTs, including the sperm storage organs, offer an exciting list of candidates to test for roles in mediating SFP responses (168170). That hundreds of SFPs are transferred to females suggest that, potentially, many molecular pathways are involved in female post-mating responses. Interactions have been shown to affect protein localization (e.g. SP to sperm; 44) and proteolytic cascades (115) in D. melanogaster, but much needs to be done to characterize these and other pathways.

Second, what extrinsic factors affect the production and transfer of SFPs and the magnitude of their effects on mated females? Few studies have investigated this question, but those that have suggest that effects of SFPs on female post-mating response are influenced by a number of factors. For example, adult female nutrition alters the magnitude of the effects of SP on different phenotypic traits, which show that the responses to mating in general, and SFPs in particular, can vary under different environmental conditions (171, 172). Furthermore, males transfer different amounts of SFPs in different contexts, such as a competitive environment (166). These effects observed in the lab suggest that modulation of SFP action and allocation in the natural setting will be important to consider for fundamental reasons and also in insect pest control.

A third question relates to the unusual evolutionary characteristics of SFPs. Conservation of protein classes indicates fundamentally conserved roles for SFPs, yet individual SFPs tend to evolve rapidly. How do different proteins come to play such roles in different species, and what forces lead to the rapid sequence evolution? These are only a small sampling of the fascinating questions that await answers.

Summary Points.

  1. SFPs have roles in modulating female behavioral and physiological processes in numerous insect species. SFPs are being identified in an increasing number of insects.

  2. Seminal fluid proteins and protein of Diptera, Lepidoptera, Hymenoptera, Coleoptera, Orthoptera, Hemiptera, and Ixodida species are described.

  3. Mating and SFP mediate female post-mating responses in processes such as transcriptional and RT structural changes, up-regulation of anti-microbial peptide genes, altered receptivity to re-mating, sperm storage, mating plug formation, post-mating feeding and female activity levels.

Future Directions.

  1. Determining how male SFPs interact with each other and with female molecules to effect the changes observed in mated females.

  2. Determining how extrinsic factors (e.g. nutrition, differing social conditions) affect the production and transfer of SFPs and the magnitude of their effects in mated females.

  3. Determining how SFPs regulate similar reproductive processes across numerous species in the face of selective pressures and rapid evolution, and examining the forces that drive these changes.

Acknowledgements

We would like to thank our colleagues in the reproductive biology and SFP fields for useful discussion, Jessica Sitnik and Erin Kelleher for helpful comments on the manuscript; and NIH grant RO1-HD038921 (to MFW) for support. We apologize to colleagues whose work could not be included or receive detailed description, due to limitations on the manuscript, length and reference list.

Terms/Definitions

Accessory gland protein

proteins made by, and expected to be secreted from, the accessory gland of male insect reproductive tracts.

Seminal fluid proteins

proteins expressed from tissues of the male reproductive tract and likely transferred to females during mating

Acronyms list

AG

accessory gland

SFP

seminal fluid protein

RT

reproductive tract

Acp

accessory gland protein

SP

sex peptide

SPR

sex peptide receptor

Contributor Information

Frank W. Avila, Email: fwa5@cornell.edu.

Laura K. Sirot, Email: ls286@cornell.edu.

Brooke A. LaFlamme, Email: bal44@cornell.edu.

C. Dustin Rubinstein, Email: cdr25@cornell.edu.

References

  • 1.Poiani A. Complexity of seminal fluid: a review. Behavioral Ecology and Sociobiology. 2006;60:289–310. [Google Scholar]
  • 2.Loher W, Ganjian I, Kubo I, Stanleysamuelson D, Tobe SS. Prostaglandins - Their Role in Egg-Laying of the Cricket Teleogryllus-Commodus. Proceedings of the National Academy of Sciences of the United States of America-Biological Sciences. 1981;78:7835–7838. doi: 10.1073/pnas.78.12.7835. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 3.Pondeville E, Maria A, Jacques JC, Bourgouin C, Dauphin-Villemant C. Anopheles gambiae males produce and transfer the vitellogenic steroid hormone 20-hydroxyecdysone to females during mating. Proceedings of the National Academy of Sciences of the United States of America. 2008;105:19631–19636. doi: 10.1073/pnas.0809264105. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 4.Gillott C. Male accessory gland secretions: modulators of female reproductive physiology and behavior. Annu Rev Entomol. 2003;48:163–184. doi: 10.1146/annurev.ento.48.091801.112657. [DOI] [PubMed] [Google Scholar]
  • 5.Chen PS. The Functional-Morphology and Biochemistry of Insect Male Accessory-Glands and Their Secretions. Annual Review of Entomology. 1984;29:233–255. [Google Scholar]
  • 6.Leopold RA. Role of Male Accessory Glands in Insect Reproduction. Annual Review of Entomology. 1976;21:199–221. [Google Scholar]
  • 7.de la Fuente J, Kocan KM, Blouin EF. Tick Vaccines and the transmission of tick-borne pathogens. Veterinary Research Communications. 2007;31:85–90. doi: 10.1007/s11259-007-0069-5. [DOI] [PubMed] [Google Scholar]
  • 8.Heckel DG, Gahan LJ, Baxter SW, Zhao JZ, Shelton AM, Gould F, Tabashnik BE. The diversity of Bt resistance genes in species of Lepidoptera. Journal of Invertebrate Pathology. 2007;95:192–197. doi: 10.1016/j.jip.2007.03.008. [DOI] [PubMed] [Google Scholar]
  • 9.Kolb A, Needham GR, Neyman KM, High WA. Bedbugs. Dermatol Ther. 2009;22:347–352. doi: 10.1111/j.1529-8019.2009.01246.x. [DOI] [PubMed] [Google Scholar]
  • 10.Romero A, Potter MF, Potter DA, Haynes KF. Insecticide resistance in the bed bug: a factor in the pest’s sudden resurgence? J Med Entomol. 2007;44:175–178. doi: 10.1603/0022-2585(2007)44[175:IRITBB]2.0.CO;2. [DOI] [PubMed] [Google Scholar]
  • 11.Graf JF, Gogolewski R, Leach-Bing N, Sabatini GA, Molento MB, Bordin EL, Arantes GJ. Tick control: an industry point of view. Parasitology. 2004;129(Suppl):S427–S442. doi: 10.1017/s0031182004006079. [DOI] [PubMed] [Google Scholar]
  • 12.Sirot LK, LaFlamme BA, Sitnik JL, Rubinstein CD, Avila FW, Chow CY, Wolfner MF. Molecular social interactions: Drosophila melanogaster seminal fluid proteins as a case study. Adv Genet. 2009;68:23–56. doi: 10.1016/S0065-2660(09)68002-0. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 13.Fricke C, Perry J, Chapman T, Rowe L. The conditional economics of sexual conflict. Biol Lett. 2009;5:671–674. doi: 10.1098/rsbl.2009.0433. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 14.Findlay GD, Swanson WJ. Proteomics enhances evolutionary and functional analysis of reproductive proteins. Bioessays. 2010;32:26–36. doi: 10.1002/bies.200900127. [DOI] [PubMed] [Google Scholar]
  • 15.Pizzari T, Snook RR. Perspective: sexual conflict and sexual selection: chasing away paradigm shifts. Evolution. 2003;57:1223–1236. doi: 10.1111/j.0014-3820.2003.tb00331.x. [DOI] [PubMed] [Google Scholar]
  • 16.Weiss BL, Stepczynski JM, Wong P, Kaufman WR. Identification and characterization of genes differentially expressed in the testis/vas deferens of the fed male tick, Amblyomma hebraeum. Insect Biochem Mol Biol. 2002;32:785–793. doi: 10.1016/s0965-1748(01)00161-8. [DOI] [PubMed] [Google Scholar]
  • 17.Braswell WE, Andres JA, Maroja LS, Harrison RG, Howard DJ, Swanson WJ. Identification and comparative analysis of accessory gland proteins in Orthoptera. Genome. 2006;49:1069–1080. doi: 10.1139/g06-061. [DOI] [PubMed] [Google Scholar]
  • 18.Andres JA, Maroja LS, Bogdanowicz SM, Swanson WJ, Harrison RG. Molecular evolution of seminal proteins in field crickets. Mol Biol Evol. 2006;23:1574–1584. doi: 10.1093/molbev/msl020. [DOI] [PubMed] [Google Scholar]
  • 19.Chintapalli VR, Wang J, Dow JAT. Using FlyAtlas to identify better Drosophila melanogaster models of human disease. Nature Genetics. 2007;39:715–720. doi: 10.1038/ng2049. [DOI] [PubMed] [Google Scholar]
  • 20.Parthasarathy R, Tan A, Sun Z, Chen Z, Rankin M, Palli SR. Juvenile hormone regulation of male accessory gland activity in the red flour beetle, Tribolium castaneum. Mech Dev. 2009;126:563–579. doi: 10.1016/j.mod.2009.03.005. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 21.Wagstaff BJ, Begun DJ. Molecular population genetics of accessory gland protein genes and testis-expressed genes in Drosophila mojavensis and D. arizonae. Genetics. 2005;171:1083–1101. doi: 10.1534/genetics.105.043372. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 22.Walters JR, Harrison RG. EST analysis of male accessory glands from Heliconius butterflies with divergent mating systems. BMC Genomics. 2008;9:592. doi: 10.1186/1471-2164-9-592. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 23.Begun DJ, Lindfors HA, Thompson ME, Holloway AK. Recently evolved genes identified from Drosophila yakuba and D. erecta accessory gland expressed sequence tags. Genetics. 2006;172:1675–1681. doi: 10.1534/genetics.105.050336. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 24.Davies SJ, Chapman T. Identification of genes expressed in the accessory glands of male Mediterranean Fruit Flies (Ceratitis capitata) Insect Biochem Mol Biol. 2006;36:846–856. doi: 10.1016/j.ibmb.2006.08.009. [DOI] [PubMed] [Google Scholar]
  • 25.Dottorini T, Nicolaides L, Ranson H, Rogers DW, Crisanti A, Catteruccia F. A genome-wide analysis in Anopheles gambiae mosquitoes reveals 46 male accessory gland genes, possible modulators of female behavior. Proc Natl Acad Sci U S A. 2007;104:16215–16220. doi: 10.1073/pnas.0703904104. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 26.Swanson WJ, Clark AG, Waldrip-Dail HM, Wolfner MF, Aquadro CF. Evolutionary EST analysis identifies rapidly evolving male reproductive proteins in Drosophila. Proc Natl Acad Sci U S A. 2001;98:7375–7379. doi: 10.1073/pnas.131568198. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 27.Ram KR, Wolfner MF. Seminal influences: Drosophila Acps and the molecular interplay between males and females during reproduction. Integrative and Comparative Biology. 2007;47:427–445. doi: 10.1093/icb/icm046. [DOI] [PubMed] [Google Scholar]
  • 28.Baer B, Heazlewood JL, Taylor NL, Eubel H, Millar AH. The seminal fluid proteome of the honeybee Apis mellifera. Proteomics. 2009;9:2085–2097. doi: 10.1002/pmic.200800708. [DOI] [PubMed] [Google Scholar]
  • 29.Collins AM, Caperna TJ, Williams V, Garrett WM, Evans JD. Proteomic analyses of male contributions to honey bee sperm storage and mating. Insect Mol Biol. 2006;15:541–549. doi: 10.1111/j.1365-2583.2006.00674.x. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 30.Findlay GD, Yi X, Maccoss MJ, Swanson WJ. Proteomics reveals novel Drosophila seminal fluid proteins transferred at mating. PLoS Biol. 2008;6:e178. doi: 10.1371/journal.pbio.0060178. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 31.Reinhardt K, Wong CH, Georgiou AS. Detection of seminal fluid proteins in the bed bug, Cimex lectularius, using two-dimensional gel electrophoresis and mass spectrometry. Parasitology. 2009;136:283–292. doi: 10.1017/S0031182008005362. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 32.Sirot LK, Poulson RL, McKenna MC, Girnary H, Wolfner MF, Harrington LC. Identity and transfer of male reproductive gland proteins of the dengue vector mosquito, Aedes aegypti: potential tools for control of female feeding and reproduction. Insect Biochem Mol Biol. 2008;38:176–189. doi: 10.1016/j.ibmb.2007.10.007. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 33.Takemori N, Yamamoto MT. Proteome mapping of the Drosophila melanogaster male reproductive system. Proteomics. 2009;9:2484–2493. doi: 10.1002/pmic.200800795. [DOI] [PubMed] [Google Scholar]
  • 34.Kelleher ES, Watts TD, LaFlamme BA, Haynes PA, Markow TA. Proteomic analysis of Drosophila mojavensis male accessory glands suggests novel classes of seminal fluid proteins. Insect Biochem Mol Biol. 2009;39:366–371. doi: 10.1016/j.ibmb.2009.03.003. [DOI] [PubMed] [Google Scholar]
  • 35.Walker MJ, Rylett CM, Keen JN, Audsley N, Sajid M, Shirras AD, Isaac RE. Proteomic identification of Drosophila melanogaster male accessory gland proteins, including a pro-cathepsin and a soluble gamma-glutamyl transpeptidase. Proteome Sci. 2006;4:9. doi: 10.1186/1477-5956-4-9. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 36.Findlay GD, MacCoss MJ, Swanson WJ. Proteomic discovery of previously unannotated, rapidly evolving seminal fluid genes in Drosophila. Genome Res. 2009;19:886–896. doi: 10.1101/gr.089391.108. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 37.Rogers DW, Baldini F, Battaglia F, Panico M, Dell A, Morris HR, Catteruccia F. Transglutaminase-mediated semen coagulation controls sperm storage in the malaria mosquito. PLoS Biol. 2009;7:e1000272. doi: 10.1371/journal.pbio.1000272. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 38.Walters JR, Harrison RG. Combined EST and proteomic analysis identifies rapidly evolving seminal fluid proteins in Heliconius butterflies. Mol Biol Evol. 2010 doi: 10.1093/molbev/msq092. [DOI] [PubMed] [Google Scholar]
  • 39.Andres JA, Maroja LS, Harrison RG. Searching for candidate speciation genes using a proteomic approach: seminal proteins in field crickets. Proceedings of the Royal Society B-Biological Sciences. 2008;275:1975–1983. doi: 10.1098/rspb.2008.0423. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 40.Mueller JL, Ravi Ram K, McGraw LA, Bloch Qazi MC, Siggia ED, Clark AG, Aquadro CF, Wolfner MF. Cross-species comparison of Drosophila male accessory gland protein genes. Genetics. 2005;171:131–143. doi: 10.1534/genetics.105.043844. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 41.Begun DJ, Lindfors HA. Rapid evolution of genomic Acp complement in the melanogaster subgroup of Drosophila. Mol Biol Evol. 2005;22:2010–2021. doi: 10.1093/molbev/msi201. [DOI] [PubMed] [Google Scholar]
  • 42.Wagstaff BJ, Begun DJ. Adaptive evolution of recently duplicated accessory gland protein genes in desert Drosophila. Genetics. 2007;177:1023–1030. doi: 10.1534/genetics.107.077503. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 43.Mueller JL, Ripoll DR, Aquadro CF, Wolfner MF. Comparative structural modeling and inference of conserved protein classes in Drosophila seminal fluid. Proc Natl Acad Sci U S A. 2004;101:13542–13547. doi: 10.1073/pnas.0405579101. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 44.Ram KR, Wolfner MF. A network of interactions among seminal proteins underlies the long-term postmating response in Drosophila. Proc Natl Acad Sci U S A. 2009;106:15384–15389. doi: 10.1073/pnas.0902923106. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 45.Ravi Ram K, Ji S, Wolfner MF. Fates and targets of male accessory gland proteins in mated female Drosophila melanogaster. Insect Biochem Mol Biol. 2005;35:1059–1071. doi: 10.1016/j.ibmb.2005.05.001. [DOI] [PubMed] [Google Scholar]
  • 46.Monsma SA, Harada HA, Wolfner MF. Synthesis of two Drosophila male accessory gland proteins and their fate after transfer to the female during mating. Dev Biol. 1990;142:465–475. doi: 10.1016/0012-1606(90)90368-s. [DOI] [PubMed] [Google Scholar]
  • 47.Pilpel N, Nezer I, Applebaum SW, Heifetz Y. Mating-increases trypsin in female Drosophila hemolymph. Insect Biochem Mol Biol. 2008;38:320–330. doi: 10.1016/j.ibmb.2007.11.010. [DOI] [PubMed] [Google Scholar]
  • 48.Lung O, Wolfner MF. Drosophila seminal fluid proteins enter the circulatory system of the mated female fly by crossing the posterior vaginal wall. Insect Biochem Mol Biol. 1999;29:1043–1052. doi: 10.1016/s0965-1748(99)00078-8. [DOI] [PubMed] [Google Scholar]
  • 49.Lawniczak MK, Begun DJ. A genome-wide analysis of courting and mating responses in Drosophila melanogaster females. Genome. 2004;47:900–910. doi: 10.1139/g04-050. [DOI] [PubMed] [Google Scholar]
  • 50.McGraw LA, Gibson G, Clark AG, Wolfner MF. Genes regulated by mating, sperm, or seminal proteins in mated female Drosophila melanogaster. Curr Biol. 2004;14:1509–1514. doi: 10.1016/j.cub.2004.08.028. [DOI] [PubMed] [Google Scholar]
  • 51.Heifetz Y, Wolfner MF. Mating, seminal fluid components, and sperm cause changes in vesicle release in the Drosophila female reproductive tract. Proc Natl Acad Sci U S A. 2004;101:6261–6266. doi: 10.1073/pnas.0401337101. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 52.McGraw LA, Clark AG, Wolfner MF. Post-mating gene expression profiles of female Drosophila melanogaster in response to time and to four male accessory gland proteins. Genetics. 2008;179:1395–1408. doi: 10.1534/genetics.108.086934. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 53.Innocenti P, Morrow EH. Immunogenic males: a genome-wide analysis of reproduction and the cost of mating in Drosophila melanogaster females. J Evol Biol. 2009;22:964–973. doi: 10.1111/j.1420-9101.2009.01708.x. [DOI] [PubMed] [Google Scholar]
  • 54.Mack PD, Kapelnikov A, Heifetz Y, Bender M. Mating-responsive genes in reproductive tissues of female Drosophila melanogaster. Proc Natl Acad Sci U S A. 2006;103:10358–10363. doi: 10.1073/pnas.0604046103. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 55.Kapelnikov A, Rivlin PK, Hoy RR, Heifetz Y. Tissue remodeling: a mating-induced differentiation program for the Drosophila oviduct. BMC Dev Biol. 2008;8:114. doi: 10.1186/1471-213X-8-114. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 56.Kapelnikov A, Zelinger E, Gottlieb Y, Rhrissorrakrai K, Gunsalus KC, Heifetz Y. Mating induces an immune response and developmental switch in the Drosophila oviduct. Proc Natl Acad Sci U S A. 2008;105:13912–13917. doi: 10.1073/pnas.0710997105. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 57.Wong A, Albright SN, Giebel JD, Ram KR, Ji S, Fiumera AC, Wolfner MF. A role for Acp29AB, a predicted seminal fluid lectin, in female sperm storage in Drosophila melanogaster. Genetics. 2008;180:921–931. doi: 10.1534/genetics.108.092106. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 58.Mueller JL, Linklater JR, Ravi Ram K, Chapman T, Wolfner MF. Targeted gene deletion and phenotypic analysis of the Drosophila melanogaster seminal fluid protease inhibitor Acp62F. Genetics. 2008;178:1605–1614. doi: 10.1534/genetics.107.083766. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 59.Carvalho GB, Kapahi P, Anderson DJ, Benzer S. Allocrine modulation of feeding behavior by the Sex Peptide of Drosophila. Curr Biol. 2006;16:692–696. doi: 10.1016/j.cub.2006.02.064. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 60.Isaac RE, Li C, Leedale AE, Shirras AD. Drosophila male sex peptide inhibits siesta sleep and promotes locomotor activity in the post-mated female. Proc Biol Sci. 2010;277:65–70. doi: 10.1098/rspb.2009.1236. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 61.Liu H, Kubli E. Sex-peptide is the molecular basis of the sperm effect in Drosophila melanogaster. Proc Natl Acad Sci U S A. 2003;100:9929–9933. doi: 10.1073/pnas.1631700100. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 62.Domanitskaya EV, Liu H, Chen S, Kubli E. The hydroxyproline motif of male sex peptide elicits the innate immune response in Drosophila females. FEBS J. 2007;274:5659–5668. doi: 10.1111/j.1742-4658.2007.06088.x. [DOI] [PubMed] [Google Scholar]
  • 63.Peng J, Zipperlen P, Kubli E. Drosophila sex-peptide stimulates female innate immune system after mating via the Toll and Imd pathways. Curr Biol. 2005;15:1690–1694. doi: 10.1016/j.cub.2005.08.048. [DOI] [PubMed] [Google Scholar]
  • 64.Wigby S, Domanitskaya EV, Choffat Y, Kubli E, Chapman T. The effect of mating on immunity can be masked by experimental piercing in female Drosophila melanogaster. J Insect Physiol. 2008;54:414–420. doi: 10.1016/j.jinsphys.2007.10.010. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 65.Fedorka KM, Linder JE, Winterhalter W, Promislow D. Post-mating disparity between potential and realized immune response in Drosophila melanogaster. Proc Biol Sci. 2007;274:1211–1217. doi: 10.1098/rspb.2006.0394. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 66.Rogers DW, Whitten MM, Thailayil J, Soichot J, Levashina EA, Catteruccia F. Molecular and cellular components of the mating machinery in Anopheles gambiae females. Proc Natl Acad Sci U S A. 2008;105:19390–19395. doi: 10.1073/pnas.0809723105. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 67.Kocher SD, Richard FJ, Tarpy DR, Grozinger CM. Genomic analysis of post-mating changes in the honey bee queen (Apis mellifera) BMC Genomics. 2008;9:232. doi: 10.1186/1471-2164-9-232. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 68.Kocher SD, Tarpy DR, Grozinger CM. The effects of mating and instrumental insemination on queen honey bee flight behaviour and gene expression. Insect Mol Biol. 2009 doi: 10.1111/j.1365-2583.2009.00965.x. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 69.Richard FJ, Tarpy DR, Grozinger CM. Effects of insemination quantity on honey bee queen physiology. PLoS One. 2007;2:e980. doi: 10.1371/journal.pone.0000980. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 70.Lung O, Kuo L, Wolfner MF. Drosophila males transfer antibacterial proteins from their accessory gland and ejaculatory duct to their mates. J Insect Physiol. 2001;47:617–622. doi: 10.1016/s0022-1910(00)00151-7. [DOI] [PubMed] [Google Scholar]
  • 71.Mueller JL, Page JL, Wolfner MF. An ectopic expression screen reveals the protective and toxic effects of Drosophila seminal fluid proteins. Genetics. 2007;175:777–783. doi: 10.1534/genetics.106.065318. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 72.Otti O, Naylor RA, Siva-Jothy MT, Reinhardt K. Bacteriolytic activity in the ejaculate of an insect. Am Nat. 2009;174:292–295. doi: 10.1086/600099. [DOI] [PubMed] [Google Scholar]
  • 73.Bloch Qazi MC, Heifetz Y, Wolfner MF. The developments between gametogenesis and fertilization: ovulation and female sperm storage in Drosophila melanogaster. Dev Biol. 2003;256:195–211. doi: 10.1016/s0012-1606(02)00125-2. [DOI] [PubMed] [Google Scholar]
  • 74.Adams EM, Wolfner MF. Seminal proteins but not sperm induce morphological changes in the Drosophila melanogaster female reproductive tract during sperm storage. J Insect Physiol. 2007;53:319–331. doi: 10.1016/j.jinsphys.2006.12.003. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 75.Avila FW, Wolfner MF. Acp36DE is required for uterine conformational changes in mated Drosophila females. Proc Natl Acad Sci U S A. 2009;106:15796–15800. doi: 10.1073/pnas.0904029106. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 76.Neubaum DM, Wolfner MF. Mated Drosophila melanogaster females require a seminal fluid protein, Acp36DE, to store sperm efficiently. Genetics. 1999;153:845–857. doi: 10.1093/genetics/153.2.845. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 77.Bloch Qazi MC, Wolfner MF. An early role for the Drosophila melanogaster male seminal protein Acp36DE in female sperm storage. J Exp Biol. 2003;206:3521–3528. doi: 10.1242/jeb.00585. [DOI] [PubMed] [Google Scholar]
  • 78.Xue L, Noll M. Drosophila female sexual behavior induced by sterile males showing copulation complementation. Proc Natl Acad Sci U S A. 2000;97:3272–3275. doi: 10.1073/pnas.060018897. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 79.den Boer SP, Baer B, Dreier S, Aron S, Nash DR, Boomsma JJ. Prudent sperm use by leaf-cutter ant queens. Proc Biol Sci. 2009;276:3945–3953. doi: 10.1098/rspb.2009.1184. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 80.den Boer SP, Boomsma JJ, Baer B. Honey bee males and queens use glandular secretions to enhance sperm viability before and after storage. J Insect Physiol. 2009;55:538–543. doi: 10.1016/j.jinsphys.2009.01.012. [DOI] [PubMed] [Google Scholar]
  • 81.den Boer SP, Baer B, Boomsma JJ. Seminal fluid mediates ejaculate competition in social insects. Science. 2010;327:1506–1509. doi: 10.1126/science.1184709. [DOI] [PubMed] [Google Scholar]
  • 82.Holman L. Drosophila melanogaster seminal fluid can protect the sperm of other males. Functional Ecology. 2009;23:180–186. [Google Scholar]
  • 83.Ram KR, Wolfner MF. Sustained post-mating response in Drosophila melanogaster requires multiple seminal fluid proteins. PLoS Genet. 2007;3:e238. doi: 10.1371/journal.pgen.0030238. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 84.Peng J, Chen S, Busser S, Liu H, Honegger T, Kubli E. Gradual release of sperm bound sex-peptide controls female postmating behavior in Drosophila. Curr Biol. 2005;15:207–213. doi: 10.1016/j.cub.2005.01.034. [DOI] [PubMed] [Google Scholar]
  • 85.Chapman T, Neubaum DM, Wolfner MF, Partridge L. The role of male accessory gland protein Acp36DE in sperm competition in Drosophila melanogaster. Proc Biol Sci. 2000;267:1097–1105. doi: 10.1098/rspb.2000.1114. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 86.Chapman T, Bangham J, Vinti G, Seifried B, Lung O, Wolfner MF, Smith HK, Partridge L. The sex peptide of Drosophila melanogaster: female post-mating responses analyzed by using RNA interference. Proc Natl Acad Sci U S A. 2003;100:9923–9928. doi: 10.1073/pnas.1631635100. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 87.Yapici N, Kim YJ, Ribeiro C, Dickson BJ. A receptor that mediates the post-mating switch in Drosophila reproductive behaviour. Nature. 2008;451:33–37. doi: 10.1038/nature06483. [DOI] [PubMed] [Google Scholar]
  • 88.Yang CH, Rumpf S, Xiang Y, Gordon MD, Song W, Jan LY, Jan YN. Control of the postmating behavioral switch in Drosophila females by internal sensory neurons. Neuron. 2009;61:519–526. doi: 10.1016/j.neuron.2008.12.021. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 89.Hasemeyer M, Yapici N, Heberlein U, Dickson BJ. Sensory neurons in the Drosophila genital tract regulate female reproductive behavior. Neuron. 2009;61:511–518. doi: 10.1016/j.neuron.2009.01.009. [DOI] [PubMed] [Google Scholar]
  • 90.Miyatake T, Chapman T, Partridge L. Mating-induced inhibition of remating in female Mediterranean fruit flies Ceratitis capitata. J Insect Physiol. 1999;45:1021–1028. doi: 10.1016/s0022-1910(99)00083-9. [DOI] [PubMed] [Google Scholar]
  • 91.Jang EB. Effects of Mating and Accessory-Gland Injections on Olfactory-Mediated Behavior in the Female Mediterranean Fruit-Fly, Ceratitis-Capitata. Journal of Insect Physiology. 1995;41:705–710. [Google Scholar]
  • 92.Mossinson S, Yuval B. Regulation of sexual receptivity of female Mediterranean fruit flies: old hypotheses revisited and a new synthesis proposed. J Insect Physiol. 2003;49:561–567. doi: 10.1016/s0022-1910(03)00027-1. [DOI] [PubMed] [Google Scholar]
  • 93.Harmer AM, Radhakrishnan P, Taylor PW. Remating inhibition in female Queensland fruit flies: effects and correlates of sperm storage. J Insect Physiol. 2006;52:179–186. doi: 10.1016/j.jinsphys.2005.10.003. [DOI] [PubMed] [Google Scholar]
  • 94.Radhakrishnan P, Taylor PW. Seminal fluids mediate sexual inhibition and short copula duration in mated female Queensland fruit flies. J Insect Physiol. 2007;53:741–745. doi: 10.1016/j.jinsphys.2006.10.009. [DOI] [PubMed] [Google Scholar]
  • 95.Radhakrishnan P, Taylor PW. Ability of male Queensland fruit flies to inhibit receptivity in multiple mates, and the associated recovery of accessory glands. J Insect Physiol. 2008;54:421–428. doi: 10.1016/j.jinsphys.2007.10.014. [DOI] [PubMed] [Google Scholar]
  • 96.Bryan JH. Results of consecutive matings of female Anopheles gambiae species B with fertile and sterile males. Nature. 1968;218:489. doi: 10.1038/218489a0. [DOI] [PubMed] [Google Scholar]
  • 97.Bryan JH. Further studies on consecutive mating in Anopheles gambiae complex. Nature. 1972;239:519. doi: 10.1038/239519a0. [DOI] [PubMed] [Google Scholar]
  • 98.Shutt B, Stables L, Aboagye-Antwi F, Moran J, Tripet F. Male accessory gland proteins induce female monogamy in anopheline mosquitoes. Medical and Veterinary Entomology. 2010;24:91–94. doi: 10.1111/j.1365-2915.2009.00849.x. [DOI] [PubMed] [Google Scholar]
  • 99.Bali G, Raina AK, Kingan TG, Lopez JD. Ovipositional behavior of newly colonized corn earworm (Lepidoptera: Noctuidae) females and evidence for an oviposition stimulating factor of male origin. Annals of the Entomological Society of America. 1996;89:475–480. [Google Scholar]
  • 100.Rafaeli A. Neuroendocrine control of pheromone biosynthesis in moths. Int Rev Cytol. 2002;213:49–91. doi: 10.1016/s0074-7696(02)13012-9. [DOI] [PubMed] [Google Scholar]
  • 101.Nagalakshmi VK, Applebaum SW, Azrielli A, Rafaeli A. Female sex pheromone suppression and the fate of sex-peptide-like peptides in mated moths of Helicoverpa armigera. Arch Insect Biochem Physiol. 2007;64:142–155. doi: 10.1002/arch.20167. [DOI] [PubMed] [Google Scholar]
  • 102.Raina AK, Kingan TG, Giebultowicz JM. Mating-Induced Loss of Sex-Pheromone and Sexual Receptivity in Insects with Emphasis on Helicoverpa-Zea and Lymantria-Dispar. Archives of Insect Biochemistry and Physiology. 1994;25:317–327. [Google Scholar]
  • 103.Kingan TG, Thomas-Laemont PA, Raina AK. Male accessory gland factors elicit change from ‘virgin’ to ‘mated’ behaviour in the femael corn earworm moth Helicoverpa zea . Journal of Experimental Biology. 1993;183:61–76. [Google Scholar]
  • 104.Fan Y, Rafaeli A, Gileadi C, Kubli E, Applebaum SW. Drosophila melanogaster sex peptide stimulates juvenile hormone synthesis and depresses sex pheromone production in Helicoverpa armigera. J Insect Physiol. 1999;45:127–133. doi: 10.1016/s0022-1910(98)00106-1. [DOI] [PubMed] [Google Scholar]
  • 105.Eliyahu D, Nagalakshmi V, Applebaum SW, Kubli E, Choffat Y, Rafaeli A. Inhibition of pheromone biosynthesis in Helicoverpa armigera by pheromonostatic peptides. J Insect Physiol. 2003;49:569–74. doi: 10.1016/s0022-1910(03)00028-3. [DOI] [PubMed] [Google Scholar]
  • 106.Nagalakshmi VK, Applebaum SW, Kubli E, Choffat Y, Rafaeli A. The presence of Drosophila melanogaster sex peptide-like immunoreactivity in the accessory glands of male Helicoverpa armigera. J Insect Physiol. 2004;50:241–248. doi: 10.1016/j.jinsphys.2003.12.002. [DOI] [PubMed] [Google Scholar]
  • 107.Perry JC, Rowe L. Neither mating rate nor spermatophore feeding influences longevity in a ladybird beetle. Ethology. 2008;114:504–511. [Google Scholar]
  • 108.Perry JC, Rowe L. Ingested spermatophores accelerate reproduction and increase mating resistance but are not a source of sexual conflict. Animal Behaviour. 2008;76:993–1000. [Google Scholar]
  • 109.Moya-Larano J, Fox CW. Ejaculate size, second male size, and moderate polyandry increase female fecundity in a seed beetle. Behavioral Ecology. 2006;17:940–946. [Google Scholar]
  • 110.Ronn J, Katvala M, Arnqvist G. The costs of mating and egg production in Callosobruchus seed beetles. Animal Behaviour. 2006;72:335–342. [Google Scholar]
  • 111.Yamane T, Miyatake T, Kimura Y. Female mating receptivity after injection of male-derived extracts in Callosobruchus maculatus. Journal of Insect Physiology. 2008;54:1522–1527. doi: 10.1016/j.jinsphys.2008.09.001. [DOI] [PubMed] [Google Scholar]
  • 112.Takami Y, Sasabe M, Nagata N, Sota T. Dual function of seminal substances for mate guarding in a ground beetle. Behavioral Ecology. 2008;19:1173–1178. [Google Scholar]
  • 113.Yamane T, Kimura Y, Katsuhara M, Miyatake T. Female mating receptivity inhibited by injection of male-derived extracts in Callosobruchus chinensis. Journal of Insect Physiology. 2008;54:501–507. doi: 10.1016/j.jinsphys.2007.11.009. [DOI] [PubMed] [Google Scholar]
  • 114.Heifetz Y, Lung O, Frongillo EA, Jr., Wolfner MF. The Drosophila seminal fluid protein Acp26Aa stimulates release of oocytes by the ovary. Curr Biol. 2000;10:99–102. doi: 10.1016/s0960-9822(00)00288-8. [DOI] [PubMed] [Google Scholar]
  • 115.Ravi Ram K, Sirot LK, Wolfner MF. Predicted seminal astacin-like protease is required for processing of reproductive proteins in Drosophila melanogaster. Proc Natl Acad Sci U S A. 2006;103:18674–18679. doi: 10.1073/pnas.0606228103. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 116.Heifetz Y, Vandenberg LN, Cohn HI, Wolfner MF. Two cleavage products of the Drosophila accessory gland protein ovulin can independently induce ovulation. Proc Natl Acad Sci U S A. 2005;102:743–748. doi: 10.1073/pnas.0407692102. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 117.Tanaka ED, Hartfelder K. The initial stages of oogenesis and their relation to differential fertility in the honey bee (Apis mellifera) castes. Arthropod Structure & Development. 2004;33:431–442. doi: 10.1016/j.asd.2004.06.006. [DOI] [PubMed] [Google Scholar]
  • 118.Klowden MJ. The check is in the male: Male mosquitoes affect female physiology and behavior. Journal of the American Mosquito Control Association. 1999;15:213–220. [PubMed] [Google Scholar]
  • 119.Clements AN. The Biology of Mosquitoes. London: Chapman & Hall; 2000. [Google Scholar]
  • 120.Isaac RE, Lamango NS, Ekbote U, Taylor CA, Hurst D, Weaver RJ, Carhan A, Burnham S, Shirras AD. Angiotensin-converting enzyme as a target for the development of novel insect growth regulators. Peptides. 2007;28:153–162. doi: 10.1016/j.peptides.2006.08.029. [DOI] [PubMed] [Google Scholar]
  • 121.Ekbote U, Looker M, Isaac RE. ACE inhibitors reduce fecundity in the mosquito, Anopheles stephensi. Comp Biochem Physiol B Biochem Mol Biol. 2003;134:593–598. doi: 10.1016/s1096-4959(03)00019-8. [DOI] [PubMed] [Google Scholar]
  • 122.Jin ZY, Gong H. Male accessory gland derived factors can stimulate oogenesis and enhance oviposition in Helicoverpa armigera (Lepidoptera : Noctuidae) Archives of Insect Biochemistry and Physiology. 2001;46:175–185. doi: 10.1002/arch.1027. [DOI] [PubMed] [Google Scholar]
  • 123.Reinhardt K, Naylor RA, Siva-Jothy MT. Ejaculate components delay reproductive senescence while elevating female reproductive rate in an insect. Proc Natl Acad Sci U S A. 2009;106:21743–21747. doi: 10.1073/pnas.0905347106. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 124.Eberhard WG. Female Control: Sexual Selection by Cryptic Female Choice. Princeton, NJ: Princeton University Press; 1996. [Google Scholar]
  • 125.Birkhead T, Moller A. Female Control of Paternity. Trends in Ecology & Evolution. 1993;8:100–104. doi: 10.1016/0169-5347(93)90060-3. [DOI] [PubMed] [Google Scholar]
  • 126.Simmons LW. Sperm Competition and its Evolutionary Consequences in the Insects. Princeton: Princeton University Press; 2001. [Google Scholar]
  • 127.Orr AG, Rutowski R. The function of the sphragis in Cressida cressida (Fab.) (Lepidoptera, Papilionidae): a visual deterrent to copulation attempts. Journal of Natural History. 1991;25:703–710. [Google Scholar]
  • 128.Baer B, Morgan ED, Schmid-Hempel P. A nonspecific fatty acid within the bumblebee mating plug prevents females from remating. Proceedings of the National Academy of Sciences of the United States of America. 2001;98:3926–3928. doi: 10.1073/pnas.061027998. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 129.Lung O, Wolfner MF. Identification and characterization of the major Drosophila melanogaster mating plug protein. Insect Biochem Mol Biol. 2001;31:543–551. doi: 10.1016/s0965-1748(00)00154-5. [DOI] [PubMed] [Google Scholar]
  • 130.Bretman A, Lawniczak MK, Boone J, Chapman T. A mating plug protein reduces early female remating in Drosophila melanogaster. J Insect Physiol. 2010;56:107–113. doi: 10.1016/j.jinsphys.2009.09.010. [DOI] [PubMed] [Google Scholar]
  • 131.Polak M, Wolf LL, Starmer WT, Barker JSF. Function of the mating plug in Drosophila hibisci Bock. Behavioral Ecology and Sociobiology. 2001;49:196–205. [Google Scholar]
  • 132.Polak M, Starmer WT, Barker JSF. A mating plug and male mate choice in Drosophila hibisci Bock. Anim Behav. 1998;56:919–926. doi: 10.1006/anbe.1998.0850. [DOI] [PubMed] [Google Scholar]
  • 133.Patterson JT. A New Type of Isolating Mechanism in Drosophila. Proceedings of the National Academy of Sciences of the United States of America. 1946;32:202–208. doi: 10.1073/pnas.32.7.202. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 134.Markow TA, Ankney PF. Drosophila Males Contribute to Oogenesis in a Multiple Mating Species. Science. 1984;224:302–303. doi: 10.1126/science.224.4646.302. [DOI] [PubMed] [Google Scholar]
  • 135.Almeida FC, Desalle R. Orthology, function and evolution of accessory gland proteins in the Drosophila repleta group. Genetics. 2009;181:235–245. doi: 10.1534/genetics.108.096263. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 136.Wigby S, Chapman T. Sex peptide causes mating costs in female Drosophila melanogaster. Curr Biol. 2005;15:316–321. doi: 10.1016/j.cub.2005.01.051. [DOI] [PubMed] [Google Scholar]
  • 137.Wagner WE, Jr., Kelley RJ, Tucker KR, Harper CJ. Females receive a life-span benefit from male ejaculates in a field cricket. Evolution. 2001;55:994–1001. doi: 10.1554/0014-3820(2001)055[0994:fralsb]2.0.co;2. [DOI] [PubMed] [Google Scholar]
  • 138.Chapman T, Liddle LF, Kalb JM, Wolfner MF, Partridge L. Cost of Mating in Drosophila-Melanogaster Females Is Mediated by Male Accessory-Gland Products. Nature. 1995;373:241–244. doi: 10.1038/373241a0. [DOI] [PubMed] [Google Scholar]
  • 139.Lung O, Tram U, Finnerty C, Eipper-Mains M, Kalb JM, Wolfner MF. The Drosophila melanogaster seminal fluid protein Acp62F is a protease inhibitor that is toxic upon ectopic expression. Genetics. 2002;160:211–224. doi: 10.1093/genetics/160.1.211. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 140.Barnes AI, Wigby S, Boone JM, Partridge L, Chapman T. Feeding, fecundity and lifespan in female Drosophila melanogaster. Proc Biol Sci. 2008;275:1675–1683. doi: 10.1098/rspb.2008.0139. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 141.Balashov YS. Bloodsucking ticks (Ixodoidea)-vectors of disease of man and animals (English translation) Misc. Publ. Entomol. Soc. Amer. 1972;8:163–376. [Google Scholar]
  • 142.Harris RA, Kaufman WR. Neural Involvement in the Control of Salivary-Gland Degeneration in the Ixodid Tick Amblyomma-Hebraeum. Journal of Experimental Biology. 1984;109 281-&. [Google Scholar]
  • 143.Friesen KJ, Kaufman WR. Salivary gland degeneration and vitellogenesis in the ixodid tick Amblyomma hebraeum: Surpassing a critical weight is the prerequisite and detachment from the host is the trigger. Journal of Insect Physiology. 2009;55:936–942. doi: 10.1016/j.jinsphys.2009.06.007. [DOI] [PubMed] [Google Scholar]
  • 144.Lomas LO, Kaufman WR. The Influence of a Factor from the Male Genital-Tract on Salivary-Gland Degeneration in the Female Ixodid Tick, Amblyomma-Hebraeum. Journal of Insect Physiology. 1992;38:595–601. doi: 10.1002/arch.940210302. [DOI] [PubMed] [Google Scholar]
  • 145.Kaufman WR, Lomas LO. ‘‘Male factors’’ in ticks: Their role in feeding and egg development. Invertebrate Reproduction & Development. 1996;30:191–198. [Google Scholar]
  • 146.Weiss BL, Kaufman WR. Two feeding-induced proteins from the male gonad trigger engorgement of the female tick Amblyomma hebraeum. Proc Natl Acad Sci U S A. 2004;101:5874–5879. doi: 10.1073/pnas.0307529101. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 147.Smith A, Guo X, de la Fuente J, Naranjo V, Kocan KM, Kaufman WR. The impact of RNA interference of the subolesin and voraxin genes in male Amblyomma hebraeum (Acari: Ixodidae) on female engorgement and oviposition. Exp Appl Acarol. 2009;47:71–86. doi: 10.1007/s10493-008-9195-1. [DOI] [PubMed] [Google Scholar]
  • 148.Donohue KV, Khalil SMS, Ross E, Mitchell RD, Roe RM, Sonenshine DE. Male engorgement factor: Role in stimulating engorgement to repletion in the ixodid tick, Dermacentor variabilis. Journal of Insect Physiology. 2009;55:909–918. doi: 10.1016/j.jinsphys.2009.05.019. [DOI] [PubMed] [Google Scholar]
  • 149.Tarpy DR, Page RE. The curious promiscuity of queen honey bees (Apis mellifera): evolutionary and behavioral mechanisms. Annales Zoologici Fennici. 2001;38:255–265. [Google Scholar]
  • 150.Rexhepaj A, Liu HF, Peng J, Choffat Y, Kubli E. The sex-peptide DUP99B is expressed in the male ejaculatory duct and in the cardia of both sexes. European Journal of Biochemistry. 2003;270:4306–4314. doi: 10.1046/j.1432-1033.2003.03823.x. [DOI] [PubMed] [Google Scholar]
  • 151.Fan Y, Rafaeli A, Moshitzky P, Kubli E, Choffat Y, Applebaum SW. Common functional elements of Drosophila melanogaster seminal peptides involved in reproduction of Drosophila melanogaster and Helicoverpa armigera females. Insect Biochem Mol Biol. 2000;30:805–812. doi: 10.1016/s0965-1748(00)00052-7. [DOI] [PubMed] [Google Scholar]
  • 152.Gäde G, Hoffmann K. Neuropeptides regulating development and reproduction in insects. Physiol Entomol. 2005;30:103–121. [Google Scholar]
  • 153.Lange AB. Neural mechanisms coordinating the female reproductive system in the locust. Front Biosci. 2009;14:4401–4415. doi: 10.2741/3536. [DOI] [PubMed] [Google Scholar]
  • 154.Orchard I, Lange AB. Evidence for octopaminergic modulation of an insect visceral muscle. J Neurobiol. 1985;16:171–181. doi: 10.1002/neu.480160303. [DOI] [PubMed] [Google Scholar]
  • 155.Cook BJ, Wagner RM. Some pharmacological properties of the oviduct muscularis of the stable fly Stomoxys calcitrans. Comp Biochem Physiol C, Comp Pharmacol Toxicol. 1992;102:273–280. doi: 10.1016/0742-8413(92)90111-j. [DOI] [PubMed] [Google Scholar]
  • 156.Monastirioti M. Distinct octopamine cell population residing in the CNS abdominal ganglion controls ovulation in Drosophila melanogaster. Dev Biol. 2003;264:38–49. doi: 10.1016/j.ydbio.2003.07.019. [DOI] [PubMed] [Google Scholar]
  • 157.Middleton CA, Nongthomba U, Parry K, Sweeney ST, Sparrow JC, Elliott CJH. Neuromuscular organization and aminergic modulation of contractions in the Drosophila ovary. BMC Biol. 2006;4:17. doi: 10.1186/1741-7007-4-17. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 158.Rodríguez-Valentín R, López-González I, Jorquera R, Labarca P, Zurita M, Reynaud E. Oviduct contraction in Drosophila is modulated by a neural network that is both, octopaminergic and glutamatergic. J. Cell. Physiol. 2006;209:183–198. doi: 10.1002/jcp.20722. [DOI] [PubMed] [Google Scholar]
  • 159.Lee H-G, Seong C-S, Kim Y-C, Davis RL, Han K-A. Octopamine receptor OAMB is required for ovulation in Drosophila melanogaster. Dev Biol. 2003;264:179–190. doi: 10.1016/j.ydbio.2003.07.018. [DOI] [PubMed] [Google Scholar]
  • 160.Thomas A. Nervous control of egg progression into the common oviduct and genital chamber of the stick-insect Carausius morosus . J Insect Physiol. 1979;25:811–821. [Google Scholar]
  • 161.Friberg U. Male perception of female mating status: its effect on copulation duration, sperm defence and female fitness. Animal Behaviour. 2006;72:1259–1268. [Google Scholar]
  • 162.Hihara F. Effects of the male accessory gland secretion on oviposition and remating in females of Drosophila melanogaster . Zool. Mag. 1981;90:303–316. [Google Scholar]
  • 163.Perez-Staples D, Aluja M, Macias-Ordonez R, Sivinski J. Reproductive trade-offs from mating with a successful male: the case of the tephritid fly Anastrepha obliqua. Behavioral Ecology and Sociobiology. 2008;62:1333–1340. [Google Scholar]
  • 164.Aluja M, Rull J, Sivinski J, Trujillo G, Perez-Staples D. Male and female condition influence mating performance and sexual receptivity in two tropical fruit flies (Diptera: Tephritidae) with contrasting life histories. Journal of Insect Physiology. 2009;55:1091–1098. doi: 10.1016/j.jinsphys.2009.07.012. [DOI] [PubMed] [Google Scholar]
  • 165.Wedell N, Gage MJG, Parker GA. Sperm competition, male prudence and sperm-limited females. Trends in Ecology & Evolution. 2002;17:313–320. [Google Scholar]
  • 166.Wigby S, Sirot LK, Linklater JR, Buehner N, Calboli FC, Bretman A, Wolfner MF, Chapman T. Seminal fluid protein allocation and male reproductive success. Curr Biol. 2009;19:751–757. doi: 10.1016/j.cub.2009.03.036. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 167.Bretman A, Fricke C, Chapman T. Plastic responses of male Drosophila melanogaster to the level of sperm competition increase male reproductive fitness. Proc Biol Sci. 2009;276:1705–1711. doi: 10.1098/rspb.2008.1878. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 168.Allen AK, Spradling AC. The Sf1-related nuclear hormone receptor Hr39 regulates Drosophila female reproductive tract development and function. Development. 2008;135:311–321. doi: 10.1242/dev.015156. [DOI] [PubMed] [Google Scholar]
  • 169.Prokupek AM, Kachman SD, Ladunga I, Harshman LG. Transcriptional profiling of the sperm storage organs of Drosophila melanogaster. Insect Mol Biol. 2009;18:465–75. doi: 10.1111/j.1365-2583.2009.00887.x. [DOI] [PubMed] [Google Scholar]
  • 170.Baer B, Eubel H, Taylor NL, O'Toole N, Millar AH. Insights into female sperm storage from the spermathecal fluid proteome of the honeybee Apis mellifera. Genome Biology. 2009;10 doi: 10.1186/gb-2009-10-6-r67. - [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 171.Fricke C, Bretman A, Chapman T. Female nutritional status determines the magnitude and sign of responses to a male ejaculate signal in Drosophila melanogaster. J Evol Biol. 2010;23:157–165. doi: 10.1111/j.1420-9101.2009.01882.x. [DOI] [PubMed] [Google Scholar]
  • 172.Rogina B. The effect of sex peptide and calorie intake on fecundity in female Drosophila melanogaster. ScientificWorldJournal. 2009;9:1178–1189. doi: 10.1100/tsw.2009.126. [DOI] [PMC free article] [PubMed] [Google Scholar]

RESOURCES