Next Article in Journal
Hesperidin, A Popular Antioxidant Inhibits Melanogenesis via Erk1/2 Mediated MITF Degradation
Previous Article in Journal
Overview of Nrf2 as Therapeutic Target in Epilepsy
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

The Essential Role of Vitellogenin Receptor in Ovary Development and Vitellogenin Uptake in Bactrocera dorsalis (Hendel)

1
Key Laboratory of Entomology and Pest Control Engineering, College of Plant Protection, Southwest University, Chongqing 400715, China
2
Citrus Research Institute, Southwest University, Chongqing 400712, China
3
College of Biology and Environmental Engineering, Guiyang University, Guiyang 550005, China
*
Author to whom correspondence should be addressed.
Int. J. Mol. Sci. 2015, 16(8), 18368-18383; https://doi.org/10.3390/ijms160818368
Submission received: 29 June 2015 / Revised: 24 July 2015 / Accepted: 30 July 2015 / Published: 7 August 2015
(This article belongs to the Section Biochemistry)

Abstract

:
The vitellogenin receptor (VgR) functions as an essential component in uptaking and transporting vitellogenin (Vg) in female adults, which is involved in ovary development and oviposition. This study aimed to clarify the molecular characteristics and function of VgR in the oriental fruit fly Bactrocera dorsalis (Hendel). Here, we identified the full-length of BdVgR (GenBank Accession No. JX469118), encoding a 1925 residue (aa) protein with a 214.72 kDa molecular mass and several typical motifs of low-density lipoprotein receptor superfamily (LDLR). Phylogenic analysis suggested that BdVgR was evolutionary conserved with other Dipteran VgRs. The expression of BdVgR was exclusively detected in the ovaries rather than head, thorax or other tissues. The developmental expression patterns showed that the signal of BdVgR was detectable in very beginning of adult stage, and positively correlated with the growth rate of ovaries and the expression levels of its ligands. In addition, we also demonstrated that the expression level of BdVgR, and ovary development were significantly suppressed after being injected with BdVgR-targeted dsRNA. Together, all of these results indicated that BdVgR was critical for yolk protein absorption and ovary maturation in B. dorsalis, playing a vital role in female reproduction.

Graphical Abstract

1. Introduction

In insects, including oviparous species, successful reproduction is fundamental in maintaining their population, which relies on two key indispensable steps: vitellogenin (Vg) biosynthesis and deposition. After being synthesized in the fat body and released into the hemolymph, Vg is taken up into the developing ovaries via endocytosis by its receptor, vitellogenin receptor (VgR), which is located on the surface of the oocytes within clathrin-coated pits [1]. Then, this lipoprotein is transported to its target recipient cell by VgR, providing multiple nutritional elements to support the developments of oocytes [2].
It is documented that only a single VgR gene exists in most insect species, which could generate a long transcript, about 7.5 kb, and encode an ovary-specific polypeptide with a molecular weight of 180–214 kDa, approximately twice the VgR in vertebrates [1]. In silico analysis further suggests that insect VgR is a member of the low density lipoprotein receptor (LDLR) superfamily, which is characterized by five highly conserved, but functionally different, amino acids domains: the ligand-binding domain (LBD) containing several Class A cysteine-rich repeats (LDLRA), the epidermal growth factor (EGF)-precursor domain comprising Class B cysteine-rich (LDLRB, EGF-like repeats) and YWXD repeats, the O-linked sugar domain, the transmembrane domain, and the cytoplasmic tail domain with internalization signals [3].
The molecular characteristics of VgRs are not only identified in invertebrates, like insects [3], mites [4], ticks [5,6,7], shrimps [8,9,10], crabs [11], and nematodes [12], but also in vertebrates, such as fish [13,14,15,16], frogs [17] and chickens [18]. However, only five VgRs have been identified in Dipteran insects, including Aedes aegypti [19], Anopheles gambiae [20], Ceratitis capitata (genomic sequence), Drosophila melanogaster [21] and Musca domestica (genomic sequence). Therefore, there is not enough available information to get the full understanding of the reproductive mechanism of this order, considering its economic importance.
The oriental fruit fly, Bactrocera dorsalis (Hendel), is one of the most devastating pests of fruits and vegetables throughout some areas of Asia and the Pacific, causing grave economic losses [22]. Unfortunately, there is still lack of effective strategies to manage this pest, owing to its unique biological properties, such as the rapidly increased insecticide resistance, high fecundity and oviposition traits [22,23,24]. Although, the yolk protein genes of this species, Bdyp1 and Bdyp2, have been well investigated in previous study [25], the information on their receptor is still lack. Thus, illustrating the function of VgR in B. dorsalis will help obtain a more comprehensive understanding of the process of reproduction, and eventually develop new strategies to control this pest.
In this research, we identified the full-length of VgR from B. dorsalis (BdVgR), analyzed and compared the basic molecular and structural characteristics with those from other insects. In addition, we reported the spatial- and temporal-expression pattern of BdVgR by semi-quantitative PCR and quantitative PCR (qPCR), along with the expression profiles of its ligands (Bdyp1 and Bdyp2) and ovarian development in time course. Finally, we further verified the function of BdVgR by silencing its expression by RNA interference (RNAi).

2. Results

2.1. Sequence and Structural Characteristics of VgR in B. dorsalis

The cDNA sequence of BdVgR (GenBank Accession No. JX469118) was generated from Seven-day-old female adult of the oriental fruit fly. The full-length was 6595 bp, consisting of a 290-bp 5′-untranslated region (UTR) and a 530-bp 3′-UTR, a putative open reading frame (ORF) of 5775 bp, in addition with a polyadenylation signal “AATAAA” at 481 bp downstream of the stop codon.
The ORF of BdVgR encoded a 1925 amino acid sequence, with the predicted molecular weight (Mw) of 214.72 kDa and the theoretical isoelectric point (pI) of 5.89. Analysis of BdVgR revealed that no signal peptide was identified in the N-terminal of the putative protein sequence (Figure 1). Based on the results of conserved domain analysis, BdVgR was supposed to be a member of the low-density lipoprotein receptor (LDLR) superfamily, containing four classical and conserved modules, such as ligand-binding domain (LBD), epidermal growth factor (EGF) precursor domain, transmembrane domain and cytoplasmic domain. In total, BdVgR possessed thirteen LDLR class A repeats (LDLRA) located in the two LBDs, with five and eight LDLRA repeats in the first and second LBD domain, respectively. Both of the LDBs were followed by an EGF-precursor domain. In addition, calcium-binding EGF-like domain, YWTD domain and LDLR class B repeats (LDLRB) were all identified in both of the EGF-precursor domains. According to the prediction of TMHMM server, the hydrophobic transmembrane domain was located at 1766–1788 aa, followed by the cytoplasmic domain at 1789–1924 aa with one di-leucine LL internalization signal motif (1834–1835 aa). However, O-linked sugar domain was not identified in BdVgR.

2.2. Sequences Comparison and Phylogenetic Analysis

The comparison of BdVgR with other insect VgRs indicated that the amino acid sequences had a similar structural feature, especially within the sequences from the same insect order (Figure 1). However, the total number of LDLRA repeats varied among different insect orders. The BdVgR sequence had a Dipteran typical 13 LDLRA repeats, whereas there were 11, 12 (10 for Bombus impatiens), 13, 13 and 8 LDLRA repeats in Lepidoptera, Hymenoptera, Blattaria, Homoptera and Coleoptera, respectively. Besides, the O-linked sugar domains were not conserved even in the same insect order (except for Blattaria). In addition, the multiple alignments showed that BdVgR shared high identity with the VgRs from Dipteran species. The identities were 41.94%, 39.37%, 81.65%, 52.71% and 61.53%, comparing to the BdVgR sequence to those generated from A. aegypti [19], A.gambiae, C. capitata, D. melanogaster [21] and M. domestica, respectively. However, the identities were much lower when comparing BdVgR with other insect VgRs. In addition, the phylogenetic tree suggested that BdVgR shared a closer ancestry relationship with CcVgR compared with MdVgR and DmYPR. As expected, the VgRs from Dipteran were clustered into a separate clade from other insects (Figure 2).
Figure 1. Diagrammatic comparison of typical domains of B. dorsalis vitellogenin receptor with other insect vitellogenin receptor. S, signal peptide; LBD, ligand binding domain; EGF, EGF-precursor domain; O, O-linked sugar domain; T, transmembrane domain; C, cytoplasmic domain; ID, identity; A, LDLRA; E, EGF-like repeat; Y, YWTD; BdVgR, Bactrocera dorsalis, JX469118; CcVgR, Ceratitis captitata, JAC05586; MdVgR, Musca domestica, XP_005190432; DmVgR, Drosophila melanogaster, AAB60217; AaVgR, Aedes aegypti, AAK15810; AgVgR, Anopheles gambiae, EAA06264; AfVgR, Apis florea, XP_003690500; AmVgR, A. mellifera, XP_006563364; BiVgR, Bombus impatiens, XP_003489577; MrVgR, Megachile rotundata, XP_003704660; SiVgR, Solenopsis invicta, AAP92450; HsVgR, Harpegnathos saltator, EFN84770; NvVgR, Nasonia vitripennis, XP_001602954; TcVgR, Tribolium castaneum, XP_968903; NlVgR, Nilaparvata lugens, ADE34166; PhVgR, Pediculus humanus corporis, EEB10383; BgVgR, Blattela germanica, CAJ 19121; RmVgR, Rhyparobia maderae, BAE93218; PaVgR, Periplaneta americana, BAC02725; AsVgR: Actias selene, AFV32171; ApVgR, Antheraea pernyi, AEJ88360; BmVgR, Bombyx mori, ADK94452; HaVgR, Helicoverpa armigera, AGF33811; SlVgR, Spodoptera litura, GU983858; DpVgR, Danaus plexippus, EHJ76019.
Figure 1. Diagrammatic comparison of typical domains of B. dorsalis vitellogenin receptor with other insect vitellogenin receptor. S, signal peptide; LBD, ligand binding domain; EGF, EGF-precursor domain; O, O-linked sugar domain; T, transmembrane domain; C, cytoplasmic domain; ID, identity; A, LDLRA; E, EGF-like repeat; Y, YWTD; BdVgR, Bactrocera dorsalis, JX469118; CcVgR, Ceratitis captitata, JAC05586; MdVgR, Musca domestica, XP_005190432; DmVgR, Drosophila melanogaster, AAB60217; AaVgR, Aedes aegypti, AAK15810; AgVgR, Anopheles gambiae, EAA06264; AfVgR, Apis florea, XP_003690500; AmVgR, A. mellifera, XP_006563364; BiVgR, Bombus impatiens, XP_003489577; MrVgR, Megachile rotundata, XP_003704660; SiVgR, Solenopsis invicta, AAP92450; HsVgR, Harpegnathos saltator, EFN84770; NvVgR, Nasonia vitripennis, XP_001602954; TcVgR, Tribolium castaneum, XP_968903; NlVgR, Nilaparvata lugens, ADE34166; PhVgR, Pediculus humanus corporis, EEB10383; BgVgR, Blattela germanica, CAJ 19121; RmVgR, Rhyparobia maderae, BAE93218; PaVgR, Periplaneta americana, BAC02725; AsVgR: Actias selene, AFV32171; ApVgR, Antheraea pernyi, AEJ88360; BmVgR, Bombyx mori, ADK94452; HaVgR, Helicoverpa armigera, AGF33811; SlVgR, Spodoptera litura, GU983858; DpVgR, Danaus plexippus, EHJ76019.
Ijms 16 18368 g001
Figure 2. A phylogenetic tree of vitellogenin receptors. The amino acid sequences of BdVgR 24 and other insect VgR sequences were selected to analyze the evolutionary relationship using the Neighbor-Joining method with a bootstrap value of 1000. The dot stands for protein sequence of VgR from B. dorsalis.
Figure 2. A phylogenetic tree of vitellogenin receptors. The amino acid sequences of BdVgR 24 and other insect VgR sequences were selected to analyze the evolutionary relationship using the Neighbor-Joining method with a bootstrap value of 1000. The dot stands for protein sequence of VgR from B. dorsalis.
Ijms 16 18368 g002

2.3. Tissue-Specific Expression Pattern of BdVgR

To elucidate the tissue-specific patterns of BdVgR, semi-quantitative PCR was-conducted to detected its expression abundance among different segments and tissues from seven-day-old female adults. As a positive control, α-tubulin was detected in all segments and tissues. However, the expression levels of BdVgR were quite different among those samples. Apparently, BdVgR was expected to be expressed in the ovaries exclusively (Figure 3).
Figure 3. Spatial-expression patterns of BdVgR in different body segments and tissues of seven-day-old female adults. The tissues and segment included the fat body (Fb), midgut (Mg), Malpighian tubule (Mt), head (Hd), thorax (Th), ovary (Ov). And α-tubulin (α-tub) was used as a reference gene. The thermal cycles for PCR amplification was 30 times, yielding 163 and 184 bp PCR products for BdVgR and α-tubulin, respectively.
Figure 3. Spatial-expression patterns of BdVgR in different body segments and tissues of seven-day-old female adults. The tissues and segment included the fat body (Fb), midgut (Mg), Malpighian tubule (Mt), head (Hd), thorax (Th), ovary (Ov). And α-tubulin (α-tub) was used as a reference gene. The thermal cycles for PCR amplification was 30 times, yielding 163 and 184 bp PCR products for BdVgR and α-tubulin, respectively.
Ijms 16 18368 g003

2.4. Ovary Growth, Developmental-Specific Expression Patterns of BdVgR, Bdyp1 and Bdyp2

To evaluate the potential relationship between the development of ovary and BdVgR, the diameters of ovaries and the expression profiles of BdVgR were monitored throughout the different ages of female adults, in addition with its ligands, Bdyp1 and Bdyp2 (Figure 4). As demonstrated in Figure 4A, no obvious enlargements were detected during the first four days. However, a dramatic enlargement was observed on the seventh day, kept increasing in the following days, and finally reached the maximum size on the 16th day (2.146 ± 0.0354 mm), which was 4.13 times larger than that on the first day (0.522 ± 0.0064 mm). The developmental expression patterns showed that the mRNA of BdVgR, Bdyp1 and Bdyp2 were detectable throughout all selected time points, but extremely low in the very beginning of the adult stage (Figure 4B–D). Synchronous with the dramatic change in ovarian development, the expression level of BdVgR was also significantly increased to its summit on the seventh day, which was 102.31 times of that in the first day adults (Figure 4B). However, there was a turning point on the tenth day. The expression level of BdVgR quickly down-regulated, decreasing to 21.85% compared to the expression level of the seventh day, but finally recovered to a certain high level during the following days. Interestingly, Bdyp1 and Bdyp2 had similar transcriptional patterns, as they started to increase steeply in the seventh day and reached their maximum levels in the tenth day. However, the highest expression level of BdVgR appeared on the seventh day, three days earlier than that in Bdyp1 and Bdyp2 (Figure 4C,D).
Figure 4. The growth of ovary and temporal-expression patterns of BdVgR, Bdyp1 and Bdyp2: (A) the developmental status of ovaries; (B) the temporal-expression pattern of BdVgR; (C) the temporal-expression pattern of Bdyp1; (D) the temporal-expression pattern of Bdyp2. The bars represented the mean ± SE. Different letters indicated significant difference with ANOVA (Least Significant Difference, LSD, p < 0.05).
Figure 4. The growth of ovary and temporal-expression patterns of BdVgR, Bdyp1 and Bdyp2: (A) the developmental status of ovaries; (B) the temporal-expression pattern of BdVgR; (C) the temporal-expression pattern of Bdyp1; (D) the temporal-expression pattern of Bdyp2. The bars represented the mean ± SE. Different letters indicated significant difference with ANOVA (Least Significant Difference, LSD, p < 0.05).
Ijms 16 18368 g004

2.5. Silencing BdVgR Expression by RNA Interference

To verify the potential role of BdVgR in yolk protein uptake and ovary development, BdVgR-targeted dsRNA was synthesized and injected into the four-day-old female adults. The RNAi effects were determined by ovarian dissection and qPCR on the third day after injection, with the indicators of ovarian diameter and expression levels. Obviously, the development of ovaries was significantly arrested in dsBdVgR-treated group (Figure 5A). The diameters of ovaries dramatically reduced (1.084 ± 0.0361 mm), ranging from 0.760 to 1.483 mm (p = 0.039). Conversely, the diameter of ovaries from the dsGFP-treated group (1.751 ± 0.0243 mm), ranging from 1.490 to 1.927 mm, showed no significant difference with the seven-day-old control insects (1.761 ± 0.0206 mm, p =0.37). Coincident with the decreased size of ovaries, there was a significantly reduced BdVgR transcript level in female adults (43.7%) derived from dsBdVgR-treated group (p = 0.002) (Figure 5B). In addition, the expression profile of Bdyp1 was up-regulated in the dsBdVgR treated group (p = 0.010) (Figure 5C), and Bdyp2 was down-regulated (p = 0.021) (Figure 5D).
Figure 5. The ovary diameter and expression levels of BdVgR, Bdyp1 and Bdyp2 72 h after injected with dsRNA. (A) The diameters of ovaries; (B) The mRNA level of BdVgR; (C) The mRNA level of Bdyp1; (D) The mRNA level of Bdyp2. The bars represent the mean ± SE. The asterisks indicate significant difference with t-test, * p < 0.05 or ** p < 0.01.
Figure 5. The ovary diameter and expression levels of BdVgR, Bdyp1 and Bdyp2 72 h after injected with dsRNA. (A) The diameters of ovaries; (B) The mRNA level of BdVgR; (C) The mRNA level of Bdyp1; (D) The mRNA level of Bdyp2. The bars represent the mean ± SE. The asterisks indicate significant difference with t-test, * p < 0.05 or ** p < 0.01.
Ijms 16 18368 g005

3. Discussion

Here, the complete VgR sequence was obtained from B. dorsalis. This was the first report of VgR from a Tephritidae to date. As expected, BdVgR harbored several typical functional motifs of VgRs and the LDLR superfamily, namely, ligand binding domains, EGF-precursor domains, transmembrane domain and cytoplasmic tail domain, and showed similarity to the sequences (VgRs and YPR) in insect, particularly to those sequences in Diptera, perhaps due to the evolutionary conservation in their ligands (Figure 1). Although those motifs are conserved in insect, there were some special traits in BdVgR sequence. Similar to the VgR sequences in Diptera and Blattaria, BdVgR also had a typical 13 cysteine-rich LDLRA repeats with classical arrangement, which was five and eight repeats in the first and second LBD, respectively [26,27,28]. However, the number of LDLRA repeats and arrangement are quite different from other insect orders and species. For example, there are four- and seven-LDLRA repeats in Lepidoptera, four/two and eight repeats in Hymenoptera, single eight-repeats in Coleoptera (Tribolium castaneum), and four and eight repeats in ticks [5,29] and vertebrates [9]. In addition, both of the EGF-precursor domains contained a calcium-binding site. However, no O-link sugar domain appeared in the C-terminal of BdVgR sequence, although it was supposed to play a role in improving VgR stability and regulating the signal pathway [2] and very conserved in Blattaria [26,28] and present in A. aegypti [19], D. melanogaster [21], M. domestica and other insects. Besides, only one LL motif was found in the cytoplasmic tail of BdVgR, and no NPXY motif was found as described in the fire ant [30] and cockroaches [26,27,28].
Unlike its ligands, which are widely detected in the fat body, ovary, head and thorax in the female adult [25], the BdVgR mRNA were exclusively expressed in ovary and no traces were detected in the other tissues or segments (Figure 3). The result is along with the reports in other insect and vertebrates, such as D. melanogaster [21], A. aegypti [19], Bombyx mori [31], Antheraea pernyi [32], Spodoptera litura [33], Periplaneta americana [28], Leucophaea maderae [26], Solenopsis invicta [30], Thunnus thynnus [34] and Oncorhyn chusmykiss [35], consistent with its specific distribution and conserved role for egg development. However, recent studies point out that the VgR is no longer exclusively located in ovaries of female adults: the fat body [32], hypopharyngeal glands [36], head and even embryos [37] are found to be the extraovarian tissues or segment that express VgR; this may due to the alternative functions of their ligands in food storage [37], immunity [38] and resistance [39].
The relationship between expression level of VgR and ovarian development has been extensively studied in a variety of insects, such as, cockroach [26,27,28], fire ant [30], armyworm [33] and vertebrates, like shrimp [9], trout [35] and chicken [40]. The developmental-specific analysis indicated that the BdVgR shared a similar mRNA expression pattern with other insects, which expressed at all ovarian developmental stages, extremely low in the very beginning of previtellogenic stage, increased dramatically before fully vitellogenic period. To be more exact, the transcript pattern of BdVgR was up-regulated with the process of sexual maturation (Day 1–7), but quick declined when the ovaries fully developed and oviposition initiated (Day 9–10) (Figure 4), in coordination with the expressions of Bdyp1 and Bdyp2. The results demonstrated that the enlargement of ovary depends on vitellogenesis and endocytosis process. The over-expressed BdVgR may work as a precondition for Vgs to endocytosis effectively in female adults. Interestingly, comparing the high expression level of BdVgR, the transcript levels of Bdyp1 and Bdyp2 were relative low from the tenth day (after egg-laying), which may indicate that other ligands rather than yolk proteins play a more important role thereafter.
RNAi is now used as a powerful tool in gene functional study in biology. Therefore, we conducted an RNAi experiment to verify the function of BdVgR in the female adults. Considering the developmental expression pattern and the persistence of the knock-down effect of RNAi in B. dorsalis [41,42,43], four-day-old females and 0.7 μg exogenous dsRNA per insect were selected as the appropriate stage and the optimal dose for microinjection based on the preliminary trials. Clearly, an effective gene silence was achieved, causing a significant reduction in mRNA expression of BdVgR and inhibition of ovarian size and development (Figure 5A,B), indicating its role in transporting the nutrients into the growing oocytes and egg development. The similar results are also found in cockroach [27], silkworm [31], armyworm [33], fire ant [44] and tick [7]. Besides gene suppression and ovarian developmental delay, egg formation failure, less egg laying, oviposition delay, and accumulation of Vgs are also regarded as the equivalent phenotypes caused by VgR knockdown. The expressions of Bdyp1 and Bdyp2 have also have been influenced (Figure 5C,D). The transcript level of Bdyp1 was up-regulated and Bdyp2 was down-regulated in the dsBdVgR-treated group, this may be owing to the different regulation mechanisms of the two genes in this species [25]. Taken together, all of those results demonstrated BdVgR’s key role in transporting yolk protein and ovary development in female adults of the oriental fruit fly.

4. Experimental Section

4.1. Insects Rearing, Sample Preparation and Ovarian Diameter Measurement

The oriental fruit flies were maintained in the laboratory under 27 ± 1 °C, 70%–80% relative humidity, 14:10 h (L:D) photoperiod conditions and reared as described in our previous study [45]. Sufficient newly emerged (1-day-old) virgin females were selected randomly and maintained in the same controlled condition. The females were collected at 3-day intervals from the first day after eclosion to the following nineteen days.
Tissues (including the fat body, midgut, Malpighian tubules and ovaries) and body segments (heads and thoraxes) were dissected from 7-day-old females in Ringer’s saline under a binocular stereoscope (Olympus SZX12, Tokyo, Japan). Tissue and body segment samples were all kept in RNAstore Reagent (Tiangen Biotech, Beijing, China) and stored at −80 °C until RNA isolation.
Fifteen female adults from different time points were randomly collected and carefully dissected for ovarian diameter measurement. The diameters of the ovaries were measured with an M165C microscope equipped with LAS v3.7 (Leica Microsystems, Wetzlar, Germany). Statistical difference of the ovarian diameter was analyzed with ANOVA by SAS 8.01 program (SAS Institute Inc., Cary, NC, USA) and reported as mean ± SE (LSD or t-test, p < 0.05).

4.2. RNA Extraction and Cloning of BdVgR

Total RNA from different tissues was isolated with RNeasy® Plus Mini Kit (Qiagen GmbH, Hilden, Germany). Total RNA from the whole flies and segments was extracted with TRIzol reagent (Invitrogen, Carlsbad, CA, USA) and then treated with RQ1 RNase-Free DNase (Promega, Madison, WI, USA) to eliminate genomic DNA contamination. All of the RNA samples were detected with agarose gel electrophoresis and measured with NanoVue spectrophotometer (GE Healthcare Bio-Science, Uppsala, Sweden) to ensure the integrity, quality and concentration. The PrimeScript™ RT reagent Kit (TaKaRa, Dalian, China) was applied to synthesize first strand cDNA with 0.5 μg total RNA. All manipulations were followed the manufacturer’s protocols.
Based on the transcriptome data of B. dorsalis [46], four pairs of gene-specific primers (BdVgR-F/R1-4, Table S1) were designed to amplify the fragments of BdVgR. Further, another two pairs of gene-specific primers (BdVgR-3R1/2, BdVgR-5R1/2, Table S1) were designed to determine the 5′- and 3′-flanking regions of BdVgR using SMARTer™ RACE cDNA Amplification Kit (Clontech, Mountain View, CA, USA). All gene-specific primers were examined by DNAMAN 5.2.2 (Lynnon, PQ, Canada) to check their self-complementarities and melting temperatures.
The 25 μL PCR reaction system included: 1.0 μL of cDNA, 1.0 μL of forward and backward primer (10 μM), 2.0 μL of dNTPs (2.5 mM each), 2.5 μL of Mg2+ (25 mM), 2.5 μL of 10× PCR buffer (Mg2+ free), 15.0 μL nuclease-free water and 0.25 μL of rTaq (5 U/μL) (Takara). All the PCR reactions were conducted as the following thermal cycles: 3 min at 94 °C, 33 cycles of 30 s at 94 °C, 30 s at 55–58 °C, 2 min at 72 °C, and final extension at 72 °C for 10 min. The purified PCR products were cloned into the pGEM®-T Easy vector (Promega, Madison, WI, USA), transformed into DH5α competent cells (Biomed, Beijing, China) and finally sequenced (Invitrogen, Shanghai, China).

4.3. Bioinformatics Analysis of BdVgR

SignalP 4.1 was used to predict the putative signal peptide (http://www.cbs.dtu.dk/services/SignalP/) [47]. DNAMAN 5.2.2 was adopted to assemble the full-length of cDNA, predict the molecular weight and isoelectric point of the deduced polypeptide. The conserved motifs were analyzed with the Conserved Domain Database in NCBI (http://www.ncbi.nlm.nih.gov/cdd/) [48] and SMART database (http://smart.embl.de/) [49]. The GPP prediction server was applied to predict the O-linked sugar domain (http://comp.chem.nottingham.ac.uk/glyco/) [50]. The transmembrane region was predicted by the TMHMM Server v.2.0 (http://www.cbs.dtu.dk/services/TMHMM/). MEGA 5.04 was applied to construct the phylogenetic tree using the Neighbor-Joining method with a bootstrap value of 1000 [51].

4.4. Semi-Quantitative PCR and Quantitative PCR (qPCR)

Semi-quantitative PCR was performed to compare the transcript abundance of BdVgR in different tissues and segments with gene-specific primers (BdVgR-qF/R) (Table S2), using α-tubulin (α-tubulin-qF/R, GenBank Accession number: GU269902, Table S2) as an internal control [52]. The PCR amplification condition was 95 °C for 3 min and 30 cycles of 95 °C for 30 s, 60 °C for 30 s, 72 °C for 30 s, and 72 °C for 5 min for final extension. The 25 μL PCR reaction system included: 1.0 μL of cDNA, 1.0 μL of forward and backward primer (5 μM), 2.0 μL of dNTPs (2.5 mM each), 2.5 μL of Mg2+ (25 mM), 2.5 μL of 10× PCR buffer (Mg2+ free), 15.0 μL nuclease-free water and 0.25 μL of rTaq (5 U/μL) (Takara).
qPCR was further applied to investigate the transcript levels of BdVgR, Bdyp1 (GenBank Accession No. AF368053) and Bdyp2 (GenBank Accession No. AF368054) at different developmental stages of female adults, with gene-specific primers (BdVgR-qF/R, Bdyp1-qF/R and Bdyp1-qF/R, Table S2) and α-tubulin (α-tubulin-qF/R, Table S2) as an internal control. The qPCR reaction was performed in the StepOne Plus Real-Time PCR System (ABI, Carlsbad, CA, USA) with 1.0 μL cDNA, 1.0 μL forward and backward primer (10 μM), 7.0 μL GoTaq® qPCR Master Mix (Promega) and 10.0 μL nuclease-free ddH2O. The reaction condition was: 95 °C for 2 min, 40 cycles of 95 °C for 15 s and 60 °C for 30 s, followed by melting curve analysis. The relative mRNA levels were determined using the 2−ΔΔCt method [53]. Three independent replications were carried out for each reaction sample. The statistical analysis and data were conducted and presented as described above (SAS, p < 0.05).

4.5. RNA Interference

To obtain the BdVgR-targeting dsRNA, dsVgR, a 558 bp fragment (include T7 promoter region) located at the LBD (from 1012 to 1184 aa) was synthesized with the primers BdVgR-dsTF/TR (Table S2) as in previous study and finally eluted in DEPC water [41]. To examine the RNA interference effect of BdVgR, a total amount of 0.7 μg dsVgR was injected into the ventral abdomen of each 4-day-old virgin female adult using Nanoject II Auto-Nanoliter Inject (Drummond Scientific, Broomall, PA, USA). In addition, the dsRNA of GFP, dsGFP (Table S2), was also synthesized and injected as a negative control. After injection, all adults were removed and maintained in the same controlled condition and supplied with artificial diet. Sixty insects were treated in each group, and the experiments were independently repeated for three times. After 72 h, 6 female adults from each group (dsVgR and dsGFP) were selected randomly for RNA extraction and transcript level detection with qPCR. Besides, fifteen female adults were harvested for ovary dissection and ovarian diameter measurement. The expression levels and ovarian diameter were evaluated and reported as described above. The t-test was employed to evaluate the differences between these two groups (SAS, p < 0.05 or 0.01).

5. Conclusions

In summary, our findings showed that the full-length of VgR was obtained and identified from the oriental fruit fly. The molecular characteristics, expression patterns and function of BdVgR indicated its typical conserved molecular structure and essential function in yolk protein transportation and ovary development. Obviously, this research deepened our understanding of the mechanism of reproduction of B. dorsalis, providing a potential target in pest control. However, the hormone regulatory mechanism, for example, ecdysone, juvenile hormone and insulin are still unexplained. Thus, further studies will focus on this aspect, aiming to point out which hormone is the most important in regulating reproductive process.

Supplementary Materials

Supplementary materials can be found at https://www.mdpi.com/1422-0067/16/08/18368/s1.

Acknowledgments

This study was supported in part by the grants from the Special Fund for Agro-scientific Research in the Public Interest (201203038), the Natural Science Foundation of Chongqing (CSTC, 2013jjB0176), the earmarked fund for the Modern Agro-industry (Citrus) Technology Research System and the Fundamental Research Funds for the Central Universities (XDJK2013A017) of China.

Author Contributions

Lin Cong, Wen-Jia Yang, Xuan-Zhao Jiang and Jin-Jun Wang conceived and designed the experiments; Lin Cong performed the experiments; Lin Cong, Jin-Zhi Niu, Guang-Mao Shen, Chun Ran and Jin-Jun Wang analyzed the data; and Lin Cong and Jin-Jun Wang wrote the paper.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Sappington, T.W.; Raikhel, A.S. Molecular characteristics of insect vitellogenins and vitellogenin receptors. Insect Biochem. Mol. Biol. 1998, 28, 277–300. [Google Scholar]
  2. Rodenburg, K.W.; Smolenaars, M.M.; van Hoof, D.; van der Horst, D.J. Sequence analysis of the non-recurring C-terminal domains shows that insect lipoprotein receptors constitute a distinct group of LDL receptor family members. Insect Biochem. Mol. Biol. 2006, 36, 250–263. [Google Scholar]
  3. Tufail, M.; Takeda, M. Insect vitellogenin/lipophorin receptors: Molecular structures, role in oogenesis, and regulatory mechanisms. J. Insect Physiol. 2009, 55, 88–104. [Google Scholar] [CrossRef]
  4. Zhong, R.; Ding, T.B.; Niu, J.Z.; Xia, W.K.; Liao, C.Y.; Dou, W.; Wang, J.J. Molecular characterization of vitellogenin and its receptor genes from citrus red mite, Panonychus citri (McGregor). Int. J. Mol. Sci. 2015, 16, 4759–4773. [Google Scholar] [CrossRef]
  5. Smith, A.D.; Kaufman, W.R. Molecular characterization of the vitellogenin receptor from the tick, Amblyomma hebraeum (Acari: Ixodidae). Insect Biochem. Mol. Biol. 2013, 43, 1133–1141. [Google Scholar] [CrossRef]
  6. Boldbaatar, D.; Battsetseg, B.; Matsuo, T.; Hatta, T.; Umemiya-Shirafuji, R.; Xuan, X.N.; Fujisaki, K. Tick vitellogenin receptor reveals critical role in oocyte development and transovarial transmission of Babesia parasite. Biochem. Cell Biol. 2008, 86, 331–344. [Google Scholar] [CrossRef]
  7. Mitchell, R.D.; Ross, E.; Osgood, C.; Sonenshine, D.E.; Donohue, K.V.; Khalil, S.M.; Thompson, D.M.; Roe, R.M. Molecular characterization, tissue-specific expression and RNAi knockdown of the first vitellogenin receptor from a tick. Insect Biochem. Mol. Biol. 2007, 37, 375–388. [Google Scholar] [CrossRef] [PubMed]
  8. Lee, J.H.; Kim, B.K.; Seo, Y.I.; Choi, J.H.; Kang, S.W.; Kang, C.K.; Park, W.G.; Kim, H.W. Four cDNAs encoding lipoprotein receptors from shrimp (Pandalopsis japonica): Structural characterization and expression analysis during maturation. Comp. Biochem. Physiol. B Biochem. Mol. Biol. 2014, 169, 51–62. [Google Scholar] [CrossRef] [PubMed]
  9. Roth, Z.; Khalaila, I. Identification and characterization of the vitellogenin receptor in Macrobrachium rosenbergii and its expression during vitellogenesis. Mol. Reprod. Dev. 2012, 79, 478–487. [Google Scholar] [CrossRef] [PubMed]
  10. Tiu, S.H.K.; Benzie, J.; Chan, S.M. From hepatopancreas to ovary: molecular characterization of a shrimp vitellogenin receptor involved in the processing of vitellogenin. Biol. Reprod. 2008, 79, 66–74. [Google Scholar] [CrossRef] [PubMed]
  11. Warrier, S.; Subramoniam, T. Receptor mediated yolk protein uptake in the crab Scylla serrata: Crustacean vitellogenin receptor recognizes related mammalian serum lipoproteins. Mol. Reprod. Dev. 2002, 61, 536–548. [Google Scholar] [CrossRef] [PubMed]
  12. Grant, B.; Hirsh, D. Receptor-mediated endocytosis in the Caenorhabditis elegans oocyte. Mol. Biol. Cell 1999, 10, 4311–4326. [Google Scholar] [CrossRef] [PubMed]
  13. Mizuta, H.; Luo, W.S.; Ito, Y.; Mushirobira, Y.; Todo, T.; Hara, A.; Reading, B.J.; Sullivan, C.V.; Hiramatsu, N. Ovarian expression and localization of a vitellogenin receptor with eight ligand binding repeats in the cutthroat trout (Oncorhynchus clarki). Comp. Biochem. Physiol. B Biochem. Mol. Biol. 2013, 166, 81–90. [Google Scholar] [CrossRef] [PubMed]
  14. Dominguez, G.A.; Quattro, J.M.; Denslow, N.D.; Kroll, K.J.; Prucha, M.S.; Porak, W.F.; Grier, H.J.; Sabo-Attwood, T.L. Identification and transcriptional modulation of the largemouth bass, Micropterus salmoides, vitellogenin receptor during oocyte development by insulin and sex steroids. Biol. Reprod. 2012, 87. [Google Scholar] [CrossRef] [PubMed]
  15. Pousis, C.; Santamaria, N.; Zupa, R.; de Giorgi, C.; Mylonas, C.C.; Bridges, C.R.; de la Gandara, F.; Vassallo-Agius, R.; Bello, G.; Corriero, A. Expression of vitellogenin receptor gene in the ovary of wild and captive Atlantic bluefin tuna (Thunnus thynnus). Anim. Reprod. Sci. 2012, 132, 101–110. [Google Scholar] [CrossRef] [PubMed]
  16. Prat, F.; Coward, K.; Sumpter, J.P.; Tyler, C.R. Molecular characterization and expression of two ovarian lipoprotein receptors in the rainbow trout, Oncorhynchus mykiss. Biol. Reprod. 1998, 58, 1146–1153. [Google Scholar] [CrossRef] [PubMed]
  17. Stifani, S.; Nimpf, J.; Schneider, W.J. Vitellogenesis in Xenopus laevis and chicken: Cognate ligands and oocyte receptors. J. Biol. Chem. 1990, 265, 882–888. [Google Scholar] [PubMed]
  18. Stifani, S.; Barber, D.L.; Nimpf, J.; Schneider, W.J. A single chicken oocyte plasma membrane protein mediates uptake of very low density lipoprotein and vitellogenin. Proc. Natl. Acad. Sci. USA 1990, 87, 1955–1959. [Google Scholar] [CrossRef] [PubMed]
  19. Sappington, T.W.; Kokoza, V.A.; Cho, W.L.; Raikhel, A.S. Molecular characterization of the mosquito vitellogenin receptor reveals unexpected high homology to the Drosophila yolk protein receptor. Proc. Natl. Acad. Sci. USA 1996, 93, 8934–8939. [Google Scholar] [CrossRef] [PubMed]
  20. Cho, K.C.; Raikhel, A.S. Organization and developmental expression of the mosquito vitellogenin receptor gene. Insect Mol. Biol. 2001, 10, 465–474. [Google Scholar] [CrossRef] [PubMed]
  21. Schonbaum, C.P.; Lee, S.; Mahowald, A.P. The Drosophila yolkless gene encodes a vitellogenin receptor belonging to the low density lipoprotein receptor superfamily. Proc. Natl. Acad. Sci. USA 1995, 92, 1485–1489. [Google Scholar] [CrossRef] [PubMed]
  22. Clarke, A.R.; Armstrong, K.F.; Carmichael, A.E.; Milne, J.R.; Raghu, S.; Roderick, G.K.; Yeates, D.K. Invasive phytophagous pests arising through a recent tropical evolutionary radiation: The Bactrocera dorsalis complex of fruit flies. Annu. Rev. Entomol. 2005, 50, 293–319. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  23. Jin, T.; Zeng, L.; Lin, Y.; Lu, Y.; Liang, G. Insecticide resistance of the oriental fruit fly, Bactrocera dorsalis (Hendel) (Diptera: Tephritidae), in mainland China. Pest Manag. Sci. 2011, 67, 370–376. [Google Scholar] [CrossRef] [PubMed]
  24. Hsu, J.C.; Feng, H.T. Development of resistance to spinosad in oriental fruit fly (Diptera: Tephritidae) in laboratory selection and cross-resistance. J. Econ. Entomol. 2006, 99, 931–936. [Google Scholar] [CrossRef] [PubMed]
  25. Chen, S.L.; Lin, C.P.; Lu, K.H. cDNA isolation, expression, and hormonal regulation of yolk protein genes in the oriental fruit fly, Bactrocera dorsalis (Hendel) (Diptera: Tephritidae). J. Insect Physiol. 2012, 58, 763–770. [Google Scholar] [CrossRef] [PubMed]
  26. Tufail, M.; Takeda, M. Molecular cloning and developmental expression pattern of the vitellogenin receptor from the cockroach, Leucophaea maderae. Insect Biochem. Mol. Biol. 2007, 37, 235–245. [Google Scholar] [CrossRef] [PubMed]
  27. Ciudad, L.; Piulachs, M.D.; Belles, X. Systemic RNAi of the cockroach vitellogenin receptor results in a phenotype similar to that of the Drosophila yolkless mutant. FEBS J. 2006, 273, 325–335. [Google Scholar] [CrossRef] [PubMed]
  28. Tufail, M.; Takeda, M. Molecular cloning, characterization and regulation of the cockroach vitellogenin receptor during oogenesis. Insect Mol. Biol. 2005, 14, 389–401. [Google Scholar] [CrossRef] [PubMed]
  29. Boldbaatar, D.; Umemiya-Shirafuji, R.; Liao, M.; Tanaka, T.; Xuan, X.N.; Fujisaki, K. Multiple vitellogenins from the Haemaphysalis longicornis tick are crucial for ovarian development. J. Insect Physiol. 2010, 56, 1587–1598. [Google Scholar] [CrossRef] [PubMed]
  30. Chen, M.E.; Lewis, D.K.; Keeley, L.L.; Pietrantonio, P.V. cDNA cloning and transcriptional regulation of the vitellogenin receptor from the imported fire ant, Solenopsis invicta Buren (Hymenoptera: Formicidae). Insect Mol. Biol. 2004, 13, 195–204. [Google Scholar] [CrossRef] [PubMed]
  31. Lin, Y.; Meng, Y.; Wang, Y.X.; Luo, J.; Katsuma, S.; Yang, C.W.; Banno, Y.; Kusakabe, T.; Shimada, T.; Xia, Q.Y. Vitellogenin receptor mutation leads to the oogenesis mutant phenotype “scanty vitellin” of the silkworm, Bombyx mori. J. Biol. Chem. 2013, 288, 13345–13355. [Google Scholar] [CrossRef] [PubMed]
  32. Liu, Q.N.; Zhu, B.J.; Liu, C.L.; Wei, G.Q.; Wang, Z.G. Characterization of vitellogenin receptor (VgR) from the Chinese oak silkworm, Antheraea pernyi. Bull. Insectol. 2011, 64, 167–174. [Google Scholar]
  33. Shu, Y.H.; Wang, J.W.; Lu, K.; Zhou, J.L.; Zhou, Q.; Zhang, G.R. The first vitellogenin receptor from a Lepidopteran insect: Molecular characterization, expression patterns and RNA interference analysis. Insect Mol. Biol. 2011, 20, 61–73. [Google Scholar] [CrossRef] [PubMed]
  34. Pousis, C.; de Giorgi, C.; Mylonas, C.; Bridges, C.; Zupa, R.; Vassallo-Agius, R.; de la Gándara, F.; Dileo, C.; de Metrio, G.; Corriero, A. Comparative study of liver vitellogenin gene expression and oocyte yolk accumulation in wild and captive Atlantic bluefin tuna (Thunnus thynnus). Anim. Reprod. Sci. 2011, 123, 98–105. [Google Scholar] [CrossRef] [PubMed]
  35. Davail, B.; Pakdel, F.; Bujo, H.; Perazzolo, L.M.; Waclawek, M.; Schneider, W.J.; Le Menn, F. Evolution of oogenesis: The receptor for vitellogenin from the rainbow trout. J. Lipid Res. 1998, 39, 1929–1937. [Google Scholar] [PubMed]
  36. Amdam, G.V.; Norberg, K.; Hagen, A.; Omholt, S.W. Social exploitation of vitellogenin. Proc. Natl. Acad. Sci. USA 2003, 100, 1799–1802. [Google Scholar] [CrossRef] [PubMed]
  37. Guidugli-Lazzarini, K.R.; do Nascimento, A.M.; Tanaka, E.D.; Piulachs, M.D.; Hartfelder, K.; Bitondi, M.G.; Simoes, Z.L.P. Expression analysis of putative vitellogenin and lipophorin receptors in honey bee (Apis mellifera L.) queens and workers. J. Insect Physiol. 2008, 54, 1138–1147. [Google Scholar] [CrossRef] [PubMed]
  38. Singh, N.K.; Pakkianathan, B.C.; Kumar, M.; Prasad, T.; Kannan, M.; König, S.; Krishnan, M. Vitellogenin from the silkworm, Bombyx mori: An effective anti-bacterial agent. PLoS ONE 2013, 8, e73005. [Google Scholar] [CrossRef] [PubMed]
  39. Zhang, W.N.; Xiao, H.J.; Liang, G.M.; Guo, Y.Y.; Wu, K. Tradeoff between reproduction and resistance evolution to Bt-toxin in Helicoverpa armigera: regulated by vitellogenin gene expression. Bull. Entomol. Res. 2014, 104, 444–452. [Google Scholar] [CrossRef] [PubMed]
  40. Bujo, H.; Lindstedt, K.A.; Hermann, M.; Dalmau, L.M.; Nimpf, J.; Schneider, W.J. Chicken oocytes and somatic cells express different splice variants of a multifunctional receptor. J. Biol. Chem. 1995, 270, 23546–23551. [Google Scholar] [CrossRef] [PubMed]
  41. Yang, W.J.; Xu, K.K.; Cong, L.; Wang, J.J. Identification, mRNA expression, and functional analysis of chitin synthase 1 gene and its two alternative splicing variants in oriental fruit fly, Bactrocera dorsalis. Int. J. Biol. Sci. 2013, 9, 331–342. [Google Scholar] [CrossRef] [PubMed]
  42. Suganya, R.; Chen, S.L.; Lu, K.H. cDNA cloning and characterization of S6 kinase and its effect on yolk protein gene expression in the oriental fruit fly Bactrocera dorsalis (Hendel). Arch. Insect Biochem. Physiol. 2011, 78, 177–189. [Google Scholar] [CrossRef] [PubMed]
  43. Chen, S.L.; Dai, S.M.; Lu, K.H.; Chang, C. Female-specific doublesex dsRNA interrupts yolk protein gene expression and reproductive ability in oriental fruit fly, Bactrocera dorsalis (Hendel). Insect Biochem. Mol. Biol. 2008, 38, 155–165. [Google Scholar] [CrossRef] [PubMed]
  44. Lu, H.L.; Vinson, S.B.; Pietrantonio, P.V. Oocyte membrane localization of vitellogenin receptor coincides with queen flying age, and receptor silencing by RNAi disrupts egg formation in fire ant virgin queens. FEBS J. 2009, 276, 3110–3123. [Google Scholar] [CrossRef] [PubMed]
  45. Cong, L.; Yang, W.J.; Shen, G.M.; Dou, W.; Wang, J.J. Molecular characterization of the cDNA encoding ecdysone receptor isoform B1 and its expression in the oriental fruit fly, Bactrocera dorsalis (Diptera: Tephritidae). Fla. Entomol. 2012, 95, 650–658. [Google Scholar] [CrossRef]
  46. Shen, G.M.; Dou, W.; Niu, J.Z.; Jiang, H.B.; Yang, W.J.; Jia, F.X.; Hu, F.; Cong, L.; Wang, J.J. Transcriptome analysis of the oriental fruit fly (Bactrocera dorsalis). PLoS ONE 2011, 6, e29127. [Google Scholar] [CrossRef] [PubMed]
  47. Petersen, T.N.; Brunak, S.; Heijne, G.V.; Nielsen, H. SignalP 4.0: Discriminating signal peptides from transmembrane regions. Nat. Methods 2011, 8, 785–786. [Google Scholar] [CrossRef] [PubMed]
  48. Marchler-Bauer, A.; Derbyshire, M.K.; Gonzales, N.R.; Lu, S.; Chitsaz, F.; Geer, L.Y.; Geer, R.C.; He, J.; Gwadz, M.; Hurwitz, D.I.; et al. CDD: NCBI’s conserved domain database. Nucleic. Acids Res. 2015, 43. [Google Scholar] [CrossRef] [PubMed]
  49. Letunic, I.; Doerks, T.; Bork, P. SMART: Recent updates, new developments and status in 2015. Nucleic Acids Res. 2015, 43. [Google Scholar] [CrossRef] [PubMed]
  50. Hamby, S.E.; Hirst, J.D. Prediction of glycosylation sites using random forests. BMC Bioinformatics 2008, 9. [Google Scholar] [CrossRef] [PubMed]
  51. Tamura, K.; Peterson, D.; Peterson, N.; Stecher, G.; Nei, M.; Kumar, S. MEGA5: Molecular evolutionary genetics analysis using maximum likelihood, evolutionary distance, and maximum parsimony methods. Mol. Biol. Evol. 2011, 28, 2731–2739. [Google Scholar] [CrossRef] [PubMed]
  52. Shen, G.M.; Jiang, H.B.; Wang, X.N.; Wang, J.J. Evaluation of endogenous references for gene expression profiling in different tissues of the oriental fruit fly Bactrocera dorsalis (Diptera: Tephritidae). BMC Mol. Biol. 2011, 11. [Google Scholar] [CrossRef] [PubMed]
  53. Pfaffl, M.W. A new mathematical model for relative quantification in real-time RT-PCR. Nucleic Acids Res. 2011, 29, 2002–2007. [Google Scholar] [CrossRef]

Share and Cite

MDPI and ACS Style

Cong, L.; Yang, W.-J.; Jiang, X.-Z.; Niu, J.-Z.; Shen, G.-M.; Ran, C.; Wang, J.-J. The Essential Role of Vitellogenin Receptor in Ovary Development and Vitellogenin Uptake in Bactrocera dorsalis (Hendel). Int. J. Mol. Sci. 2015, 16, 18368-18383. https://doi.org/10.3390/ijms160818368

AMA Style

Cong L, Yang W-J, Jiang X-Z, Niu J-Z, Shen G-M, Ran C, Wang J-J. The Essential Role of Vitellogenin Receptor in Ovary Development and Vitellogenin Uptake in Bactrocera dorsalis (Hendel). International Journal of Molecular Sciences. 2015; 16(8):18368-18383. https://doi.org/10.3390/ijms160818368

Chicago/Turabian Style

Cong, Lin, Wen-Jia Yang, Xuan-Zhao Jiang, Jin-Zhi Niu, Guang-Mao Shen, Chun Ran, and Jin-Jun Wang. 2015. "The Essential Role of Vitellogenin Receptor in Ovary Development and Vitellogenin Uptake in Bactrocera dorsalis (Hendel)" International Journal of Molecular Sciences 16, no. 8: 18368-18383. https://doi.org/10.3390/ijms160818368

Article Metrics

Back to TopTop