Acessibilidade / Reportar erro

Molecular cloning and mRNA expression of the peptidoglycan recognition protein gene HcPGRP1 and its isoform HcPGRP1a from the freshwater mussel Hyriopsis cumingi

Abstract

Peptidoglycan recognition proteins (PGRPs) are innate immune molecules that have been structurally conserved throughout evolution in invertebrates and vertebrates. In this study, peptidoglycan recognition protein HcPGRP1 and its isoform HcPGRP1a were identified in the freshwater mussel Hyriopsis cumingii. The full-length cDNAs of HcPGRP1 (973 bp) and HcPGRP1a (537 bp) encoded polypeptides with 218 and 151 amino acids, respectively. Sequence analysis showed that HcPGRP1 had one C-terminal PGRP domain that was conserved throughout evolution. Phylogenetic analysis showed that HcPGRP1 clustered closely with EsPGRP4 of Euprymna scolopes. Real-time PCR showed that the mRNA transcripts of HcPGRP1 and HcPGRP1a were constitutively expressed in various tissues, with the highest level in hepatopancreas. Stimulation with lipopolysaccharide (LPS) and peptidoglycan (PGN) significantly up-regulated HcPGRP1 mRNA expression in hepatopancreas and foot, but not in gill, whereas HcPGRP1a expression was significantly up-regulated in all three tissues. Our results indicate that HcPGRP1 is both a constitutive and inducible protein that may be involved in immune responses (recognition and defense) against invaders.

expression pattern; gene cloning; Hyriopsis cumingi; innate immunity; peptidoglycan recognition protein


ANIMAL GENETICS

RESEARCH ARTICLE

Molecular cloning and mRNA expression of the peptidoglycan recognition protein gene HcPGRP1 and its isoform HcPGRP1a from the freshwater mussel Hyriopsis cumingi

Ye TaoI, II; Zi-yan YangI, III; Xin ZhangIV; Hong-juan WuI

ISchool of Environmental Science and Engineering, Huazhong University of Science and Technology, Wuhan, China

IIDepartment of Environmental Protection of Henan Province, Zhengzhou, China

IIISchool of Environment and Municipal Engineering, North China University of Water Resources and Electric Power, Zhengzhou, China

IVHenan Rural Energy & Environment Agency, Henan Provincial Department of Agriculture, Zhengzhou, China

Send correspondence to Send correspondence to Hong-juan Wu School of Environmental Science and Engineering, Huazhong University of Science and Technology Wuhan, 430074 Hubei Province, China E-mail: hongjuanwu@hust.edu.cn

ABSTRACT

Peptidoglycan recognition proteins (PGRPs) are innate immune molecules that have been structurally conserved throughout evolution in invertebrates and vertebrates. In this study, peptidoglycan recognition protein HcPGRP1 and its isoform HcPGRP1a were identified in the freshwater mussel Hyriopsis cumingii. The full-length cDNAs of HcPGRP1 (973 bp) and HcPGRP1a (537 bp) encoded polypeptides with 218 and 151 amino acids, respectively. Sequence analysis showed that HcPGRP1 had one C-terminal PGRP domain that was conserved throughout evolution. Phylogenetic analysis showed that HcPGRP1 clustered closely with EsPGRP4 of Euprymna scolopes. Real-time PCR showed that the mRNA transcripts of HcPGRP1 and HcPGRP1a were constitutively expressed in various tissues, with the highest level in hepatopancreas. Stimulation with lipopolysaccharide (LPS) and peptidoglycan (PGN) significantly up-regulated HcPGRP1 mRNA expression in hepatopancreas and foot, but not in gill, whereas HcPGRP1a expression was significantly up-regulated in all three tissues. Our results indicate that HcPGRP1 is both a constitutive and inducible protein that may be involved in immune responses (recognition and defense) against invaders.

Keywords: expression pattern, gene cloning, Hyriopsis cumingi, innate immunity, peptidoglycan recognition protein.

Introduction

The innate immune system is the first line of defense against invading microorganisms in vertebrates and the only line of defense in invertebrates and plants (Medzhitov and Janeway, 2002; Hoffmann, 2003). This system recognizes microorganisms through a series of pattern recognition receptors (PRRs) that have been highly conserved throughout evolution (Hoffmann et al., 1999; Janeway and Medzhitov, 2002). Peptidoglycan recognition proteins (PGRPs) are a type of PRR and play a very important role in various activities associated with the innate immune response of invertebrates and vertebrates, including pathogen recognition, degradation of the amidase activity of peptidoglycan (PGN), induction of phagocytosis, activation of Toll or immune deficiency (IMD) signal pathways and the prophenoloxidase cascade (Yoshida et al., 1996; Leulier et al., 2003; Mellroth et al., 2003; Bischoff et al., 2004; Chang et al., 2004; Takehana et al., 2004).

The first PGRP, a 19-kDa protein, was discovered in the hemolymph and cuticle of a silkworm (Bomby xmori) (Yoshida et al., 1996). Subsequently, highly diversified PGRP homologs were identified in insects and mammals (Werner et al., 2000; Liu et al., 2001; Christophides et al., 2002), as well as in mollusks (Su et al., 2007). Currently, more than 20 PGRP genes have been identified from mollusks. Based on the protein primary structure and length, mollusk PGRPs have been divided into two classes: short PGRPs (PGRP-S) and long PGRPs (PGRP-L). Short PGRPs are small extracellular proteins (19-20 kDa) that usually contain a signal peptide sequence and a PGRP domain. In contrast, long PGRPs have long transcripts and are either intracellular or membrane-spanning proteins (30-90 kDa). In addition to the PGRP domain in the C-terminal, PGRP-Ls also contain a large N-terminal domain of unknown function (Guan and Mariuzza, 2007). Long PGRPs often have multiple splice forms, such as BgPGRP-LA from Biomphalaria glabrata that has three isoforms (-LA, LA1 and -LA2) (Zhang et al., 2007).

All PGRP genes contain a highly conserved PGRP domain that can recognize and bind PGN; the structure of this region is similar to that of bacteriophage T7 lysozyme that can hydrolyze the bond between N-acetylmuramic acid and the peptide of bacterial PGN (Kim et al., 2003; Dziarski, 2004). Many PGRPs have amidase activity and are characterized by a Zn2+ binding site. The active site consists of two histidines, one tyrosine and one cysteine residue, with the cysteine residue being crucial for amidase activity. In Drosophila, the mutation of cysteine (Cys80) in PGRP-SA to tyrosine (Tyr80) abolishes Toll activation by Gram-positive bacteria (Michel et al., 2001). In addition, the mutation of cysteine (Cys419) to alanine (Ala419) in human PGLYRP-2 results in a loss of amidase activity, indicating that Cys419 is important for this activity (Wang et al., 2003).

In view of the significant roles of PGRPs in innate immune defense, these proteins have also been studied in mollusks. AiPGRP from Argopecten irradians is induced primarily by PGN and may be involved in the scallop immune response to infection by Gram-positive bacteria (Ni et al., 2007). In addition to binding PGN, some PGRPs may also bind other pathogen-associated molecular pattern (PAMP) molecules. For example, CfPGRP-S1 mRNA expression was up-regulated after stimulation by PGN and LPS, indicating a possible role in the immune defense against Gram-positive and Gram-negative bacteria (Yang et al., 2012).

Hyriopsis cumingii, an economically important freshwater pearl mussel species used for monitoring aquatic environments, is widely cultured in southern China. In this study, HcPGRP1 and its isoform HcPGRP1a were cloned from H. cumingii. In addition, the tissue distribution and inducibility of transcripts in various tissues challenged with LPS or PGN was investigated. The findings reported here should be useful in understanding the immune significance of PGRP in mollusks.

Materials and Methods

Experimental animals

Pearl mussels (H. cumingii) ~2 years old (shell length: 55 ± 5 mm) obtained from an aquaculture center in Yueshan Village, Ezhou city, China, were maintained in aerated fresh lake water at 20 ± 1 ºC before the experiments. For the experiments, mussels were transferred to 10 L of lake water containing Microcystis aeruginosa 905 (purchased from the Institute of Hydrobiology, Chinese Academy of Sciences), with a microcystin-LR (MC-LR) concentration of 40-60 g/L each day. After five days, the hepatopancreas was excised from the mussels and homogenized in Trizol reagent to obtain the RNA used to construct a suppression subtractive hybridization (SSH) cDNA library and clone the full-length cDNA of PGRP.

SSH cDNA library construction and EST analysis

The differential expression of hepatopancreas genes in mussels exposed to MC-LR was assessed with an SSH cDNA library constructed using a PCR-Select cDNA subtraction kit (Clontech, USA). About 400 positive clones were sequenced and 98 high quality sequences were identified; these sequences corresponded to genes involved in apoptosis, signal transduction, cytoskeletal remodel, innate immunity, material and energy metabolism, translation and transcription (Yang et al., 2012). BLAST analysis of all the EST sequences revealed that one EST (EST No.j-7; 406 bp) was homologous to PGRP 4 in E. scolopes (score 150 bits, expect 2e-42, identities 63%, positives 80%, GenBank No. AAY27976); this EST was used to design primers for cloning the HcPGRP1 gene.

Cloning of the full-length cDNA of HcPGRP1

Total RNA was extracted from the hepatopancreas of MC-LR-challenged mussels using Trizol reagent according to the manufacturer's instructions (Invitrogen, USA). cDNA was then synthesized and amplified using a RevertAid first strand cDNA synthesis kit (MBI Fermentas, Germany) according to the manufacturer's instructions. Two groups of specific primers (5' GSP1, 5' GSP2 and 3' GSP1, 3' GSP2) were designed (Table 1) from the partial PGRP sequence identified in the SSH cDNA library and were used along with adaptor primers (UPM) for 5' and 3' rapid amplification of cDNA ends (RACE). The RACE reactions were done using a SMART RACE cDNA amplification kit (Clontech) according to the manufacturer's instructions. Two rounds of PCR amplification were done using the following conditions: 94 ºC for 3 min, followed by seven cycles of 94 ºC for 30 s, 66 ºC (first round) or 68 ºC (second round) for 30 s, 72 ºC for 90 s and 28 cycles of 94 ºC for 30 s, 63 ºC (first round) or 65 ºC (second round) for 30 s, 72 ºC for 90 s, and a final extension at 72 ºC for 10 min. The PCR products were separated by electrophoresis on agarose gels and the desired fragments were recovered with an E.Z.N.A® Gel Extraction kit (Omega), cloned into the pMD-18 T vector (TaKaRa, Japan) and transformed into competent Escherichia coli TOP10. The positive recombinants were identified by blue-white color selection in ampicillin-containing LB plates and then sequenced at Shanghai Sangon Biological Engineering Technology & Services Co. Ltd. (China).

Cloning of HcPGRP1 isoform cDNA

To obtain a possible splicing variant of HcPGRP1 in H. cumingii, the primer pair isoPGRP1F/isoPGRP1R (Table 1) was designed based on the full cDNA sequence of HcPGRP1. A mixture of cDNA templates from different organs of grass carp was used to clone the splicing isoform of HcPGRP1a.

Sequence analysis

Sequence similarity with other known proteins was analyzed using the BLAST program. The alignment of multiple sequences was done with Clustal W (version 1.83). The theoretical signal peptide was predicted using the SignalP 3.0 Server. The transmembrane domain was predicted with the TMHMM program. Possible N-glycosylation sites were identified using NetNGlyc 1.0 and a Pfam protein family search was done with the Pfam HMM search program. Physico-chemical parameters were analyzed using the ProtParam program. A phylogenetic tree was constructed based on the deduced amino acid sequences using the neighbor-joining (NJ) method with 1000 bootstrap replicates and MEGA v4.0 software.

Gene expression pattern of HcPGRP1 and HcPGRP1a using real-time quantitative PCR

To investigate the tissue distribution of HcPGRP1 and HcPGRP1a transcripts, total RNA was extracted from various tissues (hemocytes, hepatopancreas, gonad, kindey, intestinal, gill, mantle, adductor muscle and foot) of healthy mussels. Total RNA was also extracted from hepatopancreas, gonad, kidney, gill and foot tissues of H. cumingii challenged with LPS or PGN for various intervals.

One hundred and twenty mussels were used for the stimulation experiment. The mussels were randomly divided into three groups (n = 40 each). The sample groups were injected with 100 µL of LPS (1 mg/mL in PBS) from E. coli 055:B5 (Sigma-Aldrich) or 100 L of PGN (1 mg/mL in PBS) from Staphylococcus aureus (Sigma-Aldrich), while the control group was injected with 100 µL of PBS. Tissue samples were obtained 3, 6, 12, 18, 24 and 36 h after challenge with PBS, LPS or PGN. Total RNA was extracted using Trizol reagent (Invitrogen) as described by the manufacturer. After treatment with RNase-free DNase, 2 µg of RNA from each sample was reverse-transcribed with a RevertAidTM First Strand cDNA synthesis kit (Fermentas) at 42 ºC using an oligo (dT) 18 primer.

HcPGRP1 expression was examined by quantitative real-time PCR (qRT-PCR) with a SYBR Green supermix (Bio-Rad, USA) in a CFX96 C1000 thermal cycler (Bio-Rad, USA). Two pairs of gene-specific primers (RT-PGRP1F and RT-PGRP1R; RT-PGRP1aF and RT-PGRP1aR) (Table 1) designed by Primer premier 5.0 for HcPGRP1 and HcPGRP1a were used to amplify PCR products of 209 bp and 223 bp, respectively. Two β-actin primers (actin-F and actin-R; Table 1) were used to amplify a 218-bp gene fragment as an internal control for qRT-PCR. The amplification efficiency of all primers was determined using standard curves and primers with an efficiency of 0.95-1.0 chosen for this study. The qRT-PCR amplifications were done in triplicate along with the internal control gene in 96-well plates. All analyses were based on the CT values of the PCR products.

The 2-ΔΔCT method was used to analyze the level of HcPGRP1 and HcPGRP1a expressi±d Schmittgen, 2001). The relative expression of the target gene was normalized to the expression of β-actin and then expressed as a fold change relative to the corresponding control group. All data showing the relative mRNA expression are the mean of at least three independent experiments. Statistical comparisons were done using Students t-test, with p < 0.05 indicating significance.

Results

Molecular features of HcPGRP1 and HcPGRP1a cDNA

The cDNA sequence of H. cumingii PGRP was confirmed by blastn analysis on NCBI and was designated as HcPGRP1 (GenBank acc. no. KF479260) (Figure 1A). The full-length nucleotide sequence (973 bp) of HcPGRP1 cDNA consisted of a 5' terminal untranslated region (UTR) of 194 bp, a 3' UTR of 122 bp with a poly (A) tail and an open reading frame (ORF) of 657 bp. The ORF encoded a polypeptide of 218 amino acids with a theoretical isoelectric point of 6.29 and predicted molecular mass of 24.98 kDa. The N-terminus of HcPGRP1 had a putative signal peptide with a putative cleavage site located after position 36 (MFQ-EG) (Figure 1A). An N-glycosylation site was predicted at position 44 (NVT) and an amidase_II domain (also known as a PGRP domain or T phage lysozyme homology domain) was predicted at the C-terminal end of HcPGRP1.


Based on the full-length cDNA sequence, a pair of primers was designed to obtain an alternative splicing isoform of HcPGRP1, which was named HcPGRP1a (GenBank acc. no. KF479261) (Figure 1B). The complete cDNA sequence of HcPGRP1a was 537 bp, with an open reading frame (ORF) of 458 bp encoding a polypeptide of 151 amino acids. Alignment of the cDNA sequences of HcPGRP1 and HcPGRP1a showed that alternative splicing in HcPGRP1 is mainly generated by deletion of the sequence between R37-S103 (RPRDIKCNVTLVTREEWHARPTRHTEHMNTPVGIVFIHHTAMAECDDQHTCTVEMQKIQNFHMDIRS) while it retained the entire signal peptide sequence and part of the PGRP domain. Table 2 shows the most important biochemical parameters of HcPGRP1 and HcPGRP1a.

Sequence analysis and phylogenetic characterization of HcPGRP1

BLAST analysis showed that the sequence identity of HcPGRP1 with other homologs ranged from 32% to 56%. The highest percentages of identity and similarity (56% and 70%, respectively) were with the EsPGRP4 homolog (GenBank accession no. AAY27976) from the Pacific oyster E. scolopes. Based on the multiple sequence alignments between HcPGRP1 and other animal PGRPs (Figure 2), four catalytic residues (His74, Tyr109, His183 and Cys191) for the T7 lysozyme Zn2+ binding were found to be highly conserved in HcPGRP1.


A phylogenetic tree was constructed using the neighbor-joining (NJ) method with 1000 bootstraps based on the multiple alignments (Figure 3). Based on the overall amino acid sequences of 25 PGRPs from other aquatic mollusks, the phylogenetic tree showed that HcPGRP1 and EsPGRP4 from E. scolopes were grouped together in a minimum cluster and then clustered with the other PGRP subfamily from E. scolopes, suggesting there was a close evolutionary relationship between these proteins.


Tissue distribution of HcPGRP1 and HcPGRP1a mRNA

SYBR Green real-time PCR analysis was used to study the distribution of mRNA expression of the two PGRPs, with the housekeeping gene β-actin used as an internal standard. HcPGRP1 and HcPGRP1a mRNA was constitutively expressed at different levels in a variety of tissues (Figure 4). The level of HcPGRP1 mRNA expression was highest in hepatopancreas followed by intestine, and lowest in hemocytes. The expression of HcPGRPS1a was also highest in hepatopancreas followed by intestine, but was lowest in adductor muscle.


Expression pattern of HcPGRP1 and HcPGRP1a in mussels challenged with LPS or PGN

Time-course experiments were used to investigate the temporal variation in HcPGRPS1 and HcPGRP1a transcription in vivo for up to 36 h after a challenge with LPS or PGN. Based on the initial expression patterns determined by RT-PCR, three H. cumingii tissues (hepatopancreas, gill and foot) were selected to evaluate the patterns of HcPGRP1 and HcPGRP1a gene expression. The β-actin gene was used as an internal control.

Figure 5A-C shows that stimulation with LPS significantly up-regulated HcPGRP1 gene expression in hepatopancreas and foot, with the highest expression occurring 18 h and 3 h post-stimulation, respectively (p < 0.05); there was no significant change in HcPGRP1 expression in gill. In contrast, HcPGRP1a expression was significantly up-regulated in hepatopancreas, gill and foot, with the greatest increase occurring 12 h, 12 h and 6 h post-stimulation, respectively (p < 0.05). In mussels stimulated with PGN (Figure 5D-F), the expression of both HcPGRP1 and HcPGRP1a was significantly up-regulated in hepatopancreas, gill and foot. HcPGRP1 mRNA expression was highest in hepatopancreas, gill and foot at 12 h, 12 h and 24 h post-stimulation, respectively (p < 0.05), whereas that of HcPGRP1a mRNA was highest at 12 h, 6 h and 18 h, respectively (p < 0.05).


Discussion

In this study, a PGRP gene (designated as HcPGRP1) and a splice variation (HcPGRP1a) were cloned from H. cumingi. Both of these genes and the previously cloned HcPGRPS1 (Yang et al., 2013) belonged to the PGRP gene family of H. cumingi. HcPGRP1 had a predicted molecular mass of 24.98 kDa with a signal peptide of 36 amino acids but no transmembrane domain, indicating that it was most likely to be a secretory protein. BLAST and phylogenetic analyses showed that HcPGRP1 was homologous to EsPGRP4 from the mollusk E. scolopes (Goodson et al., 2005) and clustered closely with other PGRPs from E. scolopes, suggesting that there was a close evolutionary relationship among these proteins. BLAST analysis also revealed that the HcPGRP1 gene shared 50% identity with HcPGRPS1 and < 56% identity with EsPGRP4, while the phylogenetic analysis placed HcPGRP1 and HcPGRPS1 on different evolutionary branches. This result suggests that these two proteins from the same organism belong to different subfamilies.

Multiple alignments with other short PGRPs showed that the four amino acids (His17, Tyr46, His122 and Cys130) that are required for the T7 lysozyme Zn2+ binding and amidase activity were conserved in HcPGRP1 (His74, Tyr109, His183 and Cys191) (Cheng et al., 1994); these sites have also been found in HcPGRPS1 from H. cumingi (Yang et al., 2013), CgPGRPS3 from Crassostrea gigas (Itoh and Takahashi, 2008), AiPGRP from A. irradias (Ni et al., 2007) and DmPGRPSC2 from Drosophila (Bischoff et al., 2006). This similarity suggests that these molecules might serve as amidases involved in the elimination of PGN during the immune response against bacteria.

Mollusk PGRPs may have developed special functions during their evolution. For example, EsPGRPs 1, 2, 3 and 4 from E. scolopes may serve as signal transducers to trigger the Toll/NF-κB phosphorylation cascade. However, the replacement of Cys160 by Ser160 in EsPGRP4 results in a loss of catalytic activity and limited PGN recognition or signal transduction functions (Goodson et al., 2005). Some PGRPs specifically or preferentially recognize Dap-type or Lys-type PGNs, with this specificity being determined by the three amino acids in the PGN binding groove. For example, in mollusks, rCfPGRPS1 shows high affinity for Lys-type PGN (Yang et al., 2010) and HcPGRPS1 displays PGN-binding activity towards DAP-type and Lys-type PGN (Yang et al., 2013), whereas human PGLYRP-1 (Gly68, Trp69 and Arg88) and Drosophila PGRP-LE (Gly234, Trp235 and Arg254) bind Dap-type PGN (Swaminathan et al., 2000; Kumar et al., 2005; Chang et al., 2006). As shown here, HcPGRP1 retained only two of the four amino acids (Trp104 and Arg123) that are normally conserved, which suggested that the binding specificity of this protein was for DAP-type PGN.

Alternative splicing, a process by which multiple different functional messenger RNAs can be synthesized from a single gene, plays a key role in the expansion of proteomic and regulatory complexity (Nilsen and Graveley, 2010). Some previous reports had shown that multiple alternative splicing isoforms exist in PGRP gene families, e.g., Drosophila has 13 PGRP genes that encode 19 proteins, Anopheles has seven PGRP genes that encode nine proteins (Werner et al., 2000; Christophides et al., 2002). In addition, B. glabrata BgPGRP-LA has three isoforms (-LA, LA1 and-LA2) (Zhang et al., 2007). As shown here, we cloned a splice variation (HcPGRP1a) of HcPGRP1 that lacked part of the PGRP domain. Alternative splicing isoforms probably interact with normal forms to produce stimulatory and inhibitory effects (Rosenstiel et al., 2006; Chang et al., 2011).

Mollusk PGRPs show highly variable expression in various tissues. For example, bay scallop (A. irradians) AiPGRP is predominantly expressed in hemocytes (Ni et al., 2007), whereas Pacific oyster (C. gigas) CgPGRP-S1L and CgPGRP-S3 are mainly expressed in mantle and digestive diverticula, respectively (Itoh and Takahashi, 2008). In H. cumingi, the pattern of HcPGRP1 and HcPGRP1a expression was similar to that of H. cumingi HcPGRPS1 (Yang et al., 2013) and Solen grandis SgPGRP-S1 (Wei et al., 2012); both genes were detected in all of the tissues examined, with the highest expression in hepatopancreas. The selective expression of PGRPs in different tissues suggests that they have different functions in the body. The responsiveness of PGRPs after exposure to microorganisms also varies considerably among host species. Stimulation with PGN, LPS and glucan markedly up-regulates the expression of CfPGRP-S1 in hemocytes of Chlamys farreri, indicating that this inducible protein is involved in the immune response to invading microbes (Yang et al., 2010). Similarly, SgPGRP-S1 and SgPGRP-S2 from S. grandis and HcPGRPS1 from H. cumingi were significantly induced in response to stimulation by LPS or PGN (Wei et al., 2012; Yang et al., 2013). As shown here, microbial ligands (LPS and PGN) also induced the expression of HcPGRP1 and HcPGRP1a in hepatopancreas, gill and foot, indicating the importance of HcPGRP1 as innate pattern recognition receptors in immune defense. Interestingly, the increase in HcPGRP1 and HcPGRP1a expression in hepatopancreas and foot tissue in response to LPS was greater than for PGN, which suggests that LPS may be the potential ligand for HcPGRP1.

Acknowledgments

This study was supported by the National Natural Science Foundation of China (Grant no. 31072209).

Internet Resources

Received: November 21, 2013;

Accepted: April 10, 2014.

Associate Editor: Célia Maria de Almeida Soares

License information: This is an open-access article distributed under the terms of the Creative Commons Attribution License, which permits unrestricted use, distribution, and reproduction in any medium, provided the original work is properly cited.

  • Bischoff V, Vignal C, Boneca IG, Michel T, Hoffmann JA and Royet J (2004) Function of the Drosophila pattern-recognition receptor PGRP-SD in the detection of Gram-positive bacteria. Nat Immunol 5:1175-1180.
  • Bischoff V, Vignal C, Duvic B, Boneca IG, Hoffmann JA and Royet J (2006) Downregulation of the Drosophila immune response by peptidoglycan-recognition proteins SC1 and SC2. PLoS Pathog 2:e14.
  • Chang CI, Pili-Floury S, Hervé M, Parquet C, Chelliah Y, Lemaitre B, Mengin-Lecreulx D and Deisenhofer J (2004) A Drosophila pattern recognition receptor contains a peptidoglycan docking groove and unusual L,D- carboxypeptidase activity. PLoS Biol 2:1293-1302.
  • Chang CI, Chelliah Y, Borek D, Mengin-Lecreulx D and Deisenhofer J (2006) Structure of tracheal cytotoxin in complex with a heterodimeric pattern-recognition receptor. Science 311:1761-1764.
  • Chang MX, Wang TH, Nie P, Zou J and Secombes CJ (2011) Cloning of two rainbow trout nucleotide-binding oligomerization domain containing 2 (NOD2) splice variants and functional characterization of the NOD2 effector domains. Fish Shellfish Immunol 30:118-127.
  • Cheng X, Zhang X, Pflugrath JW and Studier EW (1994) The structure of bacteriophage T7 lysozyme, a zinc amidase and an inhibitor of T7 RNA polymerase. Proc Natl Acad Sci USA 91:4034-4038.
  • Christophides GK, Zdobnov E, Barillas-Mury C, Birney E, Blandin S, Blass C, Brey PT, Collins FH, Danielli A, Dimopoulos G, et al. (2002) Immunity related genes and gene families in Anopheles gambiae. Science 298:159-165.
  • Dziarski R (2004) Peptidoglycan recognition proteins (PGRPs). Mol Immunol 40:877-886.
  • Goodson MS, Kojadinovic M, Troll JV, Scheetz TE, Casavant TL, Soares MB and McFall-Ngai MJ (2005) Identifying components of the NF-kappaB pathway in the beneficial Euprymna scolopes-Vibrio fischeri light organ symbiosis. Appl Environ Microbiol 71:6934-6946.
  • Guan R and Mariuzza RA (2007) Peptidoglycan recognition proteins of the innate immune system. Trends Microbiol 15:127-134.
  • Hoffmann JA (2003) The immune response of Drosophila Nature 426:33-38.
  • Hoffmann JA, Kafatos FC, Janeway CA and Ezekowitz RAB (1999) Phylogenetic perspectives in innate immunity. Science 284:1313-1318.
  • Naoki I and Keisuke GT (2008) Distribution of multiple peptidoglycan recognition proteins in the tissues of Pacific oyster, Crassostrea gigas. Comp Biochem Physiol B 150:409-417.
  • Janeway C and Medzhitov R (2002) Innate immune recognition. Annu Rev Immunol 20:197-216.
  • Kim MS, Byun M and Oh BH (2003) Crystal structure of peptidoglycan recognition protein LB from Drosophila melanogaster. Nat Immunol 4:787-793.
  • Kumar S, Roychowdhury A, Ember B, Wang Q, Guan RJ, Mariuzza RA and Boons GJ (2005) Selective recognition of synthetic lysine and meso-diaminopimelic acid-type peptidoglycan fragments by human peptidoglycan recognition proteins I alpha and S. J Biol Chem 280:37005-37012.
  • Leulier F, Parquet C, Pili-Floury S, Ryu JH, Caroff M, Lee WJ, Mengin-Lecreulx D and Lemaitre B (2003) The Drosophila immune system detects bacteria through specific peptidoglycan recognition. Nat Immunol 4:478-484.
  • Liu C, Xu ZJ, Gupta D and Dziarski R (2001) Peptidoglycan recognition proteins: A novel family of four human innate immunity pattern recognition molecules. J Biol Chem 276:34686-34694.
  • Livak KJ and Schmittgen TD (2001) Analysis of relative gene expression data using realtime quantitative PCR and the 2(T)(-Delta Delta C) method. Methods 25:402-408.
  • Medzhitov R and Janeway CA (2002) Decoding the patterns of self and nonself by the innate immune system. Science 296:298-300.
  • Mellroth P, Karlsson J and Steiner H (2003) A scavenger function for a Drosophila peptidoglycan recognition protein. J Biol Chem 278:7059-7064.
  • Michel T, Reichhart JM, Hoffmann JA and Royet J (2001) Drosophila toll is activated by Gram-positive bacteria through a circulating peptidoglycan recognition protein. Nature 414:756-759.
  • Ni D, Song L, Wu L, Wu LT, Chang YQ, Yu YD, Qiu LM and Wang LL (2007) Molecular cloning and mRNA expression of peptidoglycan recognition protein (PGRP) gene in bay scallop (Argopecten irradians, Lamarck 1819). Dev Comp Immunol 31:548-558.
  • Nilsen TW and Graveley BR (2010) Expansion of the eukaryotic proteome by alternative splicing. Nature 28:457-463.
  • Rosenstiel P, Huse K, Till A, Hampe J, Hellmig S, Sina C, Billmann S, von Kampen O, Waetzig GH, Platzer M, et al. (2006) A short isoform of NOD2/CARD15, NOD2-S, is an endogenous inhibitor of NOD2/receptor-interacting protein kinase 2-induced signaling pathways. Proc Natl Acad Sci USA 103:3280-3285.
  • Su J, Ni D, Song LS, Zhao JM and Qiu LM (2007) Molecular cloning and characterization of a short type peptidoglycan recognition protein (CfPGRP-S1) cDNA from Zhikong scallop Chlamys farreri. Fish Shellfish Immunol 23:646-656.
  • Swaminathan CP, Brown PH, Roychowdhury A, Wang Q, Guan R, Silverman N, Werner T, Liu G, Kang D, Ekengren S, et al. (2000) A family of peptidoglycan recognition proteins in the fruit fly Drosophila melanogaster. Proc Natl Acad Sci USA 97:13772-13777.
  • Takehana A, Yano T, Mita S, Kotani A, Oshima Y and Kurata S (2004) Peptidoglycan recognition protein (PGRP)-LE and PGRP-LC act synergistically in Drosophila immunity. EMBO J 23:4690-4700.
  • Wang ZM, Li X, Cocklin RR, Wang MH, Wang M, Fukase K, Inamura S, Kusumoto S, Gupta D and Dziarski R (2003) Human peptidoglycan recognition protein-L is an N-acetylmuramoyl-L-alanine amidase. J Biol Chem 278:49044-49052.
  • Wei XM, Yang JM, Yang DL, Xu J, Liu XQ, Yang JL, Fang JH and Qiao HJ (2012) Molecular cloning and mRNA expression of two peptidoglycan recognition protein (PGRP) genes from mollusk Solen grandis. Fish Shellfish Immunol 32:178-185.
  • Werner T, Liu G, Kang D, Ekengren S, Steiner H and Hultmark D (2000) A family of peptidoglycan recognition proteins in the fruit fly Drosophila melanogaster. Proc Natl Acad Sci USA 97:13772-13777.
  • Yang JL, Wang W, Wei XM, Qiu LM, Wang LL, Zhang H and Song LS (2010) Peptidoglycan recognition protein of Chlamys farreri (CfPGRP-S1) mediates immune defenses against bacterial infection. Dev Comp Immunol 34:1300-1307.
  • Yang ZY, Wu HJ and Li Y (2012) Toxic effect on tissues and differentially expressed genes in hepatopancreas identified by suppression subtractive hybridization of freshwater pearl mussel (Hyriopsis cumingii) following microcystin-LR challenge. Environ Toxicol 27:393-403.
  • Yang ZY, Li JH, Wu HJ, Li Y and Wang XY (2013) Molecular cloning and functional characterization of a short peptidoglycan recognition protein (HcPGRPS1) from the freshwater mussel, Hyriopsis cumingi Mol Immunol 56:729-738.
  • Yoshida H, Kinoshita K and Ashida M (1996) Purification of peptidoglycan recognition protein from hemolymph of the silkworm, Bombyx mori. J Biol Chem 271:13854-13860.
  • Zhang SM, Zeng Y and Loker ES (2007) Characterization of immune genes from the schistosome host snail Biomphalaria glabrata that encode peptidoglycan recognition proteins and Gram-negative bacteria binding protein. Immunogenetics 59:883-898.
  • BLAST program; http://www.ncbi.nlm.nih.gov/BLAST
  • Theoretical signal peptide prediction SignalP 3.0 Server, http://www.cbs.dtu.dk/services/SignalP/
  • Transmembrane domain prediction TMHMM program, http://www.cbs.dtu.dk/services/TMHMM-2.0/
  • N-glycosylation sites prediction NetNGlyc 1.0 program http://www.cbs.dtu.dk/services/NetNGlyc/
  • Pfam HMM search program, http://pfam.janelia.org
  • Physico-chemical parameter analysis ProtParam program, http://www.expasy.ch/tools/protparam.html
  • Send correspondence to

    Hong-juan Wu
    School of Environmental Science and Engineering, Huazhong University of Science and Technology
    Wuhan, 430074 Hubei Province, China
    E-mail:
  • Publication Dates

    • Publication in this collection
      22 Sept 2014
    • Date of issue
      Sept 2014

    History

    • Received
      21 Nov 2013
    • Accepted
      10 Apr 2014
    Sociedade Brasileira de Genética Rua Cap. Adelmio Norberto da Silva, 736, 14025-670 Ribeirão Preto SP Brazil, Tel.: (55 16) 3911-4130 / Fax.: (55 16) 3621-3552 - Ribeirão Preto - SP - Brazil
    E-mail: editor@gmb.org.br