GENE-39684; No. of pages: 10; 4C: Gene xxx (2014) xxx–xxx

Contents lists available at ScienceDirect

Gene journal homepage: www.elsevier.com/locate/gene

Two small heat shock protein genes in Apis cerana cerana: characterization, regulation, and developmental expression Zhaohua Liu a, Pengbo Yao b, Xingqi Guo b, Baohua Xu a,⁎ a b

College of Animal Science and Technology, Shandong Agricultural University, Taian, Shandong 271018, PR China College of Life Sciences, Shandong Agricultural University, Taian, Shandong 271018, PR China

a r t i c l e

i n f o

Article history: Received 5 October 2013 Received in revised form 24 April 2014 Accepted 8 May 2014 Available online xxxx Keywords: Apis cerana cerana Expression Gene cloning Heat shock Ecdysone

a b s t r a c t In the present study, we identified and characterized two small heat shock protein genes from Apis cerana cerana, named AccHsp24.2 and AccHsp23.0. An alignment analysis showed that AccHsp24.2 and AccHsp23.0 share high similarity with other members of the α-crystallin/sHSP family, all of which contain the conserved α-crystallin domain. The recombinant AccHsp24.2 and AccHsp23.0 proteins were shown to have molecular chaperone activity by the malate dehydrogenase thermal aggregation assay. Three heat shock elements were detected in the 5′flanking region of AccHsp24.2 and eleven in AccHsp23.0, and two Drosophila Broad-Complex genes for ecdysone steroid response sites were found in each of the genes. The presence of these elements suggests that the expression of these genes might be regulated by heat shock and ecdysone, which was confirmed by quantitative RT-PCR (RT-qPCR). The results revealed that the expression of the two genes could be induced by cold shock (4 °C) and heat shock (37 °C and 43 °C) in an analogous manner, and AccHsp24.2 was more susceptible than AccHsp23.0. In addition, the expression of the two genes was induced by high concentrations of ecdysone in vitro and in vivo. The accumulation of AccHsp24.2 and AccHsp23.0 mRNA was also detected in different developmental stages and tissues. In spite of the differential expression at the same stage, these genes shared similar developmental patterns, suggesting that they are regulated by similar mechanisms. © 2014 Elsevier B.V. All rights reserved.

1. Introduction Abbreviations: AccHsp23.0, Apis cerana cerana heat shock protein 23.0 gene; AccHsp24.2, Apis cerana cerana heat shock protein 24.2 gene; AccHsp27.6, Apis cerana cerana heat shock protein 27.6 gene; AccHsp23.0, Apis cerana cerana heat shock protein 23.0; AccHsp24.2, Apis cerana cerana heat shock protein 24.2; AccHsp27.6, Apis cerana cerana heat shock protein 27.6; ACD, α-crystallin domain; AgHsp20.9, Anopheles gambiae heat shock protein 20.9; AmHsp25.6, Apis mellifera heat shock protein 25.6; A2, 2-day-old adult worker; A10, 10-day-old adult worker; BmHsp20.1, Bombyx mori heat shock protein 20.1; bp, base pair(s); BR-C, Broad-Complex; BSA, bovine serum albumin; CD, conserved domain; cDNA, DNA complementary to RNA; DBRC, Drosophila Broad-Complex; DmHsp26, Drosophila melanogaster heat shock protein 26; EP, epidermis; FA, fat body; HE, head; HSE, heat shock element; HSF, heat shock factor; HSP, heat shock protein; IPTG, isopropyl β-D-thiogalactopyranoside; kDa, kilodalton(s); LhHsp20, Liriomyza huidobrensis heat shock protein 20; LmHsp20.6, Locusta migratoria heat shock protein 20.6; LsHsp21.7, Liriomyza sativae heat shock protein 21.7; L2, second instar larvae; L4, fourth instar larvae; L6, sixth instar larvae; McsHsp, Macrocentrus cingulum small heat shock protein; MDH, mitochondrial malate dehydrogenase; MG, midgut; MS, muscle; NvHsp21.7, Nasonia vitripennis heat shock protein 21.7; ORF, open reading frame; PBS, phosphate buffer solution; Pd, dark eyed pupae; pI, isoelectric point; Pp, pink eyed pupae; Pw, white eyed pupae; RACE, rapid amplification of cDNA ends; RNase, ribonuclease; RT-PCR, reverse transcription polymerase chain reaction; RT-qPCR, quantitative realtime polymerase chain reaction; SDS-PAGE, sodium dodecyl sulfate polyacrylamide gel electrophoresis; sHSP, small heat shock protein; TcHsp21.8, Tribolium castaneum heat shock protein 21.8; TH, thorax; UTR, untranslated region(s); UV, ultraviolet; 20E, 20hydroxyecdysone. ⁎ Corresponding author. E-mail address: [email protected] (B. Xu).

The Chinese honeybee (Apis cerana cerana), a subspecies of the Eastern honey bee or A. cerana, is an important indigenous species that plays a critical role in agricultural economic development as a pollinator of flowering plants, particularly winter-flowering plants. A. cerana is highly resistant to cold, pests, and diseases and also plays an important role in combating soil degradation by pollinating wild plants, thereby ensuring that more biomass may be returned to the soil. Therefore, A. cerana cerana is of great significance to the ecological environment of China. However, because of deforestation, excessive pesticide usage, environmental pollution, and other reasons, there is concern with regard to the survival of A. cerana cerana (Yang, 2005). Heat shock proteins (HSPs) are protein chaperones that are ubiquitously expressed in most organisms; these proteins are also involved in some processes of embryogenesis, diapause, and morphogenesis (Hendrick and Hartl, 1993). Traditionally, the HSP superfamily has been classified based on the molecular weights of the proteins, with four types of HSPs, including HSP90, HSP70, HSP60, and a family of small HSPs (sHSPs) (Mizrahi et al., 2010). There are numerous studies concerning sHSPs in bacteria, algae, plants, amphibians, birds, and mammals. The four primary sHSPs of Drosophila melanogaster (Hsp22, Hsp23, Hsp26, and Hsp27) have been studied in detail: these proteins share significant sequence similarity and are coordinately expressed

http://dx.doi.org/10.1016/j.gene.2014.05.034 0378-1119/© 2014 Elsevier B.V. All rights reserved.

Please cite this article as: Liu, Z., et al., Two small heat shock protein genes in Apis cerana cerana: characterization, regulation, and developmental expression, Gene (2014), http://dx.doi.org/10.1016/j.gene.2014.05.034

2

Z. Liu et al. / Gene xxx (2014) xxx–xxx

following stress yet have distinct developmental expression patterns and intracellular localizations (Michaud et al., 2002). The sHSPs range in molecular weight from approximately 15 to 43 kDa and exhibit the greatest variation in sequence, size, and function compared to other HSP members (de Jong et al., 1998). Despite these differences, two or three conserved domains are present in almost all sHSPs, the most prominent of which is the α-crystallin domain. In the higher-order structure, the α-crystallin domain forms a conserved β-sandwich composed of two antiparallel β-sheets (PérezMorales et al., 2009). This conserved structure facilitates the assembly of sHSPs into oligomeric complexes of up to 800 kDa, which is crucial for their primary function as molecular chaperones, binding to partially denatured proteins to prevent irreversible protein aggregation during such stress as extreme temperature, oxidation, UV irradiation, heavy metals, and chemical intoxication (Reineke, 2005). In addition to stress responses, the chaperone function of sHSPs is also involved in normal physiological processes, including cell growth, differentiation, apoptosis (Arrigo, 1998), membrane fluidity (Tsvetkova et al., 2002), diapause (Gkouvitsas et al., 2008), and lifespan (Morrow et al., 2004a). According to previous reports, HSP90, HSP70, and HSP40 function in the maturation process of molting steroid hormone receptors (Brandt and Blagg, 2009; Zheng et al., 2010), and the genes encoding small HSPs may be regulated by ecdysone. The regulation of the D. melanogaster Hsp23 gene by ecdysterone has been analyzed by measuring the activities of hybrid Hsp-Escherichia coli β-galactosidase genes in transfected hormone-sensitive D. melanogaster cells (Mestril et al., 1986). Salivary gland culture experiments have shown that the hsp27 gene from Ceratitis capitata can also be regulated by 20-hydroxyecdysone (20E) (Kokolakis et al., 2008). However, there is limited information on the ecdysone regulation of the genes encoding small HSPs in other insects. We previously cloned and characterized one small HSP gene from A. cerana cerana (Liu et al., 2012). In the present study, another two small HSP genes were cloned, and the regulation of these genes by heat shock and ecdysone was examined. Moreover, the recombinant proteins showed molecular chaperone activity in vitro by inhibiting the thermal aggregation of the mitochondrial malate dehydrogenase enzyme. 2. Experimental procedures 2.1. Animals and treatments Worker bees of A. cerana cerana were obtained from the Technology Park of Shandong Agricultural University. The entire bodies of fourth (L4)- and sixth (L6)-instar larvae, white (Pw)-, pink (Pp)-, and dark (Pd)-eyed pupae from the hive, and adult workers (2 and 10 d after emergence) were obtained by labeling newly emerged bees with paint. The head (HE), thorax (TH), midgut (MG), muscle (MS), epidermis (EP), and fat body (FA) of newly emerged bees were dissected on ice, frozen immediately in liquid nitrogen, and stored at −80 °C. Adult bees (12 d) were also divided into groups (each group contains 30 individuals) and subjected to different temperatures. Second (L2)-instar larvae were maintained in a 24-well plate and reared on an artificial diet at 34 °C and 95% humidity. After treated with different artificial diets for 96 h, the larvae were flash-frozen in liquid nitrogen at the indicated time points and stored at −80 °C. 2.2. RNA extraction, cDNA synthesis, and DNA isolation Total RNA was extracted using the Trizol reagent (Invitrogen, Carlsbad, CA, USA) according to the manufacturer's instructions. The RNA was digested with RNase-free DNase I to remove any genomic DNA, and cDNA was synthesized using a reverse transcription system (TransGen Biotech, Beijing, China). Genomic DNA was isolated using an EasyPure Genomic DNA Extraction Kit in accordance

with the manufacturer's instructions (TransGen Biotech, Beijing, China). 2.3. PCR primers and conditions The primers used in this study are listed in Table 1. The amplification conditions are shown in Table 2. 2.4. Isolation of the AccHsp24.2 and AccHsp23.0 genes The primer pairs P1 + P2 and P3 + P4 were designed based on a conserved DNA sequence among homologous genes from Apis mellifera, Nasonia vitripennis, and Macrocentrus cingulum and used to clone the internal conserved fragments of AccHsp24.2 and AccHsp23.0 by RT-PCR. The 5P1, 5P2, 3P1, and 3P2 primers and 5P3, 5P4, 3P3, and 3P4 primers were designed based on these internal fragment sequences for 5′ and 3′ RACE (rapid amplification of cDNA ends), respectively. Following RACEPCR, the full-length cDNA sequences were derived through the assembly of the 5′- and 3′-end sequences and the internal fragment. Subsequently, end-to-end PCRs (QP1 + QP2 and QP3 + QP4) were used to amplify the full-length cDNA sequence of the two genes. The genomic DNA was also cloned (GP1 + GP2 and GP3 + GP4) using genomic DNA as the template. All the PCRs were performed as previously described (Wang et al., 2008). 2.5. Cloning of the 5′-flanking region of AccHsp24.2 and AccHsp23.0 To clone the 5′-flanking regions of AccHsp24.2 and AccHsp23.0, the restriction endonucleases DraI and SspI, respectively, were used in inverse PCR. Genomic DNA was digested with the indicated restriction endonuclease at 37 °C overnight, ligated using T4 DNA ligase (TaKaRa, Dalian, China), and then used as templates. The PS1 + PX1 and PS2 + PX2 primers and the PS3 + PX3 and PS4 + PX4 primers were designed based on the genomic sequence of AccHsp24.2 and AccHsp23.0 to obtain the 5′-flanking regions. The PCR conditions are shown in Table 2. The PCR products were cloned into the pMD18-T vector and sequenced. The transcription factor database TRANSFAC R.3.4 (Tsunoda and Takagi, 1999) and Matinspector were used to search for transcription factor binding sites in the 5′-flanking region. 2.6. Bioinformatic and phylogenetic analyses Conserved domains in AccHsp24.2 and AccHsp23.0 were detected using the NCBI bioinformatics tools (http://blast.ncbi.nlm.nih.gov/ Blast.cgi). In addition, the DNAman software was used to search for open reading frames (ORFs) and to perform multiple sequence alignments. The PSIPRED Protein Sequence Analysis Workbench (http:// bioinf.cs.ucl.ac.uk/psipred/) was used to predict secondary structures, and the Protein Data Bank (http://www.pdb.org/) was searched for protein 3D structure models. Phylogenetic and molecular evolutionary analyses were performed using the neighbor-joining method with the Molecular Evolutionary Genetics Analysis (MEGA version 4.1) software. 2.7. Expression analysis using real-time PCR To determine the expression profile of AccHsp24.2 and AccHsp23.0, real-time PCR was performed with the SYBR® PrimeScript™ RT-PCR Kit (TaKaRa, Dalian, China) using a CFX 96™ real-time system (BioRad, USA). The AccHsp24.2 and AccHsp23.0 transcripts were amplified using two pairs of specific primers: RP1 + RP2 and RP3 + RP4, respectively. The primers β-actin-s and β-actin-x were used to amplify the β-actin transcripts from A. cerana cerana (GenBank accession no. XM_640276). The following real-time PCR amplification conditions were used: an initial denaturation at 95 °C for 30 s; 41 cycles of 95 °C for 5 s, 53 °C for 15 s, and 72 °C for 15 s; and a single melt cycle from 65 °C to 95 °C. The relative expression was analyzed

Please cite this article as: Liu, Z., et al., Two small heat shock protein genes in Apis cerana cerana: characterization, regulation, and developmental expression, Gene (2014), http://dx.doi.org/10.1016/j.gene.2014.05.034

Z. Liu et al. / Gene xxx (2014) xxx–xxx

3

Table 1 Primers used in this study. Abbreviation

Primer sequence (5′–3′)

Description

P1 P2 P3 P4 5P1 5P2 3P1 3P2 5P3 5P4 3P3 3P4 AAP AUAP B26 B25 QP1 QP2 QP3 QP4 GP1 GP2 GP3 GP4 PS1 PX1 PS2 PX2 PS3 PX3 PS4 PX4 RP1 RP2 RP3 RP4 β-Actin-s β-Actin-x EP1 EP2 EP3 EP4

ATGTCATTGTTGCCGCTG GCTCTTCATGCTGATTCG ATGTCTCTGATTCCATTG TCACCATCGAGCAGACGGG GATCGTCCACATCGTATC ATGTCATTGTTGCCGCTG GCTCTTCATGCTGATTCG GCAAACAGCTGACGAGAAG CCTCATCGACTACTTGATC ATGTCTCTGATTCCATTG GGAGCAACCAAAGATCCAG TCACCATCGAGCAGACGGG GGCCACGCGTCGACTAGTAC(G)14 GGCCACGCGTCGACTAGTAC GACTCTAGACGACATCGA(T)18 GACTCTAGACGACATCGA GTTGAACTTCAAACATTGG CTATGTTAATATGTATAGACTG CACTCGAGAGAGAAGCTTGAC GTGTTCCTGCTTCTTACTGC GTTGAACTTCAAACATTGG CTATGTTAATATGTATAGACTG CACTCGAGAGAGAAGCTTGAC GTGTTCCTGCTTCTTACTGC CAACAGAGCTATGTTGGAT CACATCGCATCTGGGATC GATTGGTCGTGGTAGAGG TTCTGCCTGGTGGGCAG GCAACCAAAGATCCAGAGCG GGTTTGGGAATATATCCCG CGGCATTGAAGGAGAACACG CTGATTGGTGGGAAGACCTG GATTGGTCGTGGTAGAGG GCTCTTCGTGCTGATTCG CTGATTGGTGGGAAGACCTG CTTGGGGTGAACTTCTGCG GTTTTCCCATCTATCGTCGG TTTTCTCCATATCATCCCAG GGTACCATGTCATTATTGCCG AAGCTTTTTTTTCTCTGGCATC GGTACCATGTCTTTGATTCCA AAGCTTATTTTCTTCTTTCTTC

cDNA sequence primer of AccHsp24.2, forward cDNA sequence primer of AccHsp24.2, reverse cDNA sequence primer of AccHsp23.0, forward cDNA sequence primer of AccHsp23.0, reverse 5′ RACE reverse primer of AccHsp24.2, outer 5′ RACE reverse primer of AccHsp24.2, inner 3′ RACE forward primer of AccHsp24.2, outer 3′ RACE forward primer of AccHsp24.2, inner 5′ RACE reverse primer of AccHsp23.0, outer 5′ RACE reverse primer of AccHsp23.0, inner 3′ RACE reverse primer of AccHsp23.0, outer 3′ RACE reverse primer of AccHsp23.0, inner Abridged anchor primer Abridged universal amplification primer 3′ RACE universal adaptor primer 3′ RACE universal primer Full-length cDNA primer of AccHsp24.2, forward Full-length cDNA primer of AccHsp24.2, reverse Full-length cDNA primer of AccHsp23.0, forward Full-length cDNA primer of AccHsp23.0, reverse Genomic sequence primer of AccHsp24.2, forward Genomic sequence primer of AccHsp24.2, reverse Genomic sequence primer of AccHsp23.0, forward Genomic sequence primer of AccHsp23.0, reverse Inverse PCR forward primer, outer of AccHsp24.2 Inverse PCR reverse primer, outer of AccHsp24.2 Inverse PCR forward primer, inner of AccHsp24.2 Inverse PCR reverse primer, inner of AccHsp24.2 Inverse PCR forward primer, outer of AccHsp23.0 Inverse PCR reverse primer, outer of AccHsp23.0 Inverse PCR forward primer, inner of AccHsp23.0 Inverse PCR reverse primer, inner of AccHsp23.0 Real time-PCR primer of AccHsp24.2, forward Real time-PCR primer of AccHsp24.2, reverse Real time-PCR primer of AccHsp23.0, forward Real time-PCR primer of AccHsp23.0, reverse Standard control primer, forward Standard control primer, reverse Protein expression primer, forward of AccHsp24.2 Protein expression primer, reverse of AccHsp24.2 Protein expression primer, forward of AccHsp23.0 Protein expression primer, reverse of AccHsp23.0

using the comparative CT method (2− ΔΔCT method). Three individual samples were prepared from each sample, and all the samples were analyzed in triplicate.

into pEasy-T3 and digested with the restriction endonucleases KpnI and HindIII; the sequences were then ligated into the expression vector pET30a (+), which was digested with the same restriction endonucleases. We next transformed each expression plasmid, pET-30a (+)-AccHsp24.2 and pET-30a (+)-AccHsp23.0, into the BL21 (DE3) E. coli strain. After intermediate culture and isopropyl-β-D-thiogalactopyranoside (IPTG) induction, the bacterial cells were harvested by centrifugation, resuspended, and sonicated. After centrifugation of the lysate, the pellet was resuspended in PBS and subjected to 15% SDS-PAGE for expression analysis. The recombinant proteins were purified using MagneHis(™) Protein

2.8. Construction of expression plasmids, recombinant protein expression, and purification To express the recombinant AccHsp24.2 and AccHsp23.0 proteins in BL21 cells, specific primers (EP1 + EP2 and EP3 + EP4, respectively) were designed to amplify the entire ORF. Each ORF sequence was cloned

Table 2 PCR amplification conditions used in this study. Primes pair

Amplification conditions

P1 + P2/P3 + P4 5P1 + B26/5P3 + B26 5P2 + B25/5P4 + B25 3P1 + AAP/3P3 + AAP 3P2 + AUAP/3P4 + AUAP QP1 + QP2/QP3 + QP4 GP1 + GP2/GP3 + GP4 PS1 + PX1/PS3 + PX3 PS2 + PX2/PS4 + PX4 EP1 + EP2/EP3 + EP4

6 6 6 6 6 6 6 6 6 6

min at 94 min at 94 min at 94 min at 94 min at 94 min at 94 min at 94 min at 94 min at 94 min at 94

°C, 40 °C, 40 °C, 40 °C, 40 °C, 40 °C, 40 °C, 40 °C, 50 °C, 50 °C, 40

s at 94 s at 94 s at 94 s at 94 s at 94 s at 94 s at 94 s at 94 s at 94 s at 94

°C, 40 °C, 40 °C, 40 °C, 40 °C, 40 °C, 40 °C, 40 °C, 50 °C, 50 °C, 40

s at 52 s at 53 s at 58 s at 49 s at 55 s at 52 s at 52 s at 50 s at 50 s at 53

°C, 1 min at 72 °C for 35 cycles, 10 min at 72 °C °C, 30 s at 72 °C for 28 cycles, 5 min at 72 °C °C, 30 s at 72 °C for 35 cycles, 5 min at 72 °C °C, 40 s at 72 °C for 28 cycles, 5 min at 72 °C °C, 40 s at 72 °C for 35 cycles, 5 min at 72 °C °C, 1 min at 72 °C for 35 cycles, 10 min at 72 °C °C, 1.5 min at 72 °C for 35 cycles, 10 min at 72 °C °C, 2 min at 72 °C for 30 cycles, 5 min at 72 °C °C, 2 min at 72 °C for 30 cycles, 5 min at 72 °C °C, 1.5 min at 72 °C for 28 cycles, 10 min at 72 °C

Please cite this article as: Liu, Z., et al., Two small heat shock protein genes in Apis cerana cerana: characterization, regulation, and developmental expression, Gene (2014), http://dx.doi.org/10.1016/j.gene.2014.05.034

4

Z. Liu et al. / Gene xxx (2014) xxx–xxx

Purification System (Promega, Madison, WI, USA) according to the manufacturer's instructions, and the purified proteins were examined by 15% SDS-PAGE. 2.9. Ecdysone treatment in vitro New pupae during the Pw stage were dissected in sterilized Ringer's saline for the extraction of the thoracic integument. The adhering muscle and fat body were removed, and pools of five or six thoracic integuments were incubated in 5 ml of commercially available Grace's culture medium specifically for insect tissues (Invitrogen, Carlsbad, CA, USA). The integument was incubated in the presence of 1, 0.1, or 0.001 μg of 20E per ml of culture medium for a period of 3 h. After this process, the integuments were washed for 3 h in hormone-free medium (Sun et al., 2013). Total RNA was extracted from the incubated integuments and analyzed by RT-qPCR. 2.10. Larval rearing in the laboratory A. cerana cerana larvae were collectively reared in 24-well plates (Costar, Corning Incorporated, USA) and placed in a desiccator at a constant temperature (34 °C) and humidity (96%); a 10% glycerol solution was used inside the desiccator. The larvae were initially fed an artificial diet (Silva et al., 2009) and, once reaching the third instar, were randomly transferred to artificial diets containing 1, 0.1, and 0.01 μg/ml 20E. One 24-well plate was filled with 48 larvae (two larvae/well), and four plates were treated with one concentration of 20E. The larvae that were still alive after 72 h of feeding on the treatment diets were separated into groups for RNA extraction and analyzed by RT-PCR. 2.11. Molecular chaperone activity of the recombinant proteins The in vitro molecular chaperone activity of AccHsp24.2 and AccHsp23.0 was evaluated by measuring their capacity to suppress the thermal aggregation of mitochondrial malate dehydrogenase (MDH) from pig heart (EC 1.1.1.37; Amresco). Four samples (A, MDH; B, MDH + BSA; C, MDH + AccHsp24.2; and D, MDH + AccHsp23.0) were incubated at 43 °C. BSA (bovine serum albumin) was used as a control to exclude nonspecific chaperone activity. The absorbance was monitored at 360 nm using an Ultrospec 3000 UV/visible spectrophotometer (Pharmacia Biotech) at regular intervals (10 min) for 1 h (Pérez-Morales et al., 2009).

consistent with De Jong et al. (1998), who determined that the secondary structure of sHSPs is rich in β-sheets. In addition, there were no cysteine residues in the entire protein sequence of AccHsp24.2 and only two in AccHsp23.0. These findings are consistent with the research of Fu et al. (2003), who showed that cysteine residues were rarely found in the sequences of molecular chaperones compared to other protein families. Moreover, we also analyzed the tertiary structures of the two gene products, and both proteins shared the best template for 3D homology modeling of the deduced protein, a solid-state NMR structure of the α-crystallin domain in α-β-crystallin oligomers (PDB code, 2KLR; E-Value, 5.5127E-25 for AccHsp24.2 and 5.7653E-26 for AccHsp23.0) (Fig. 1). This model consisted of a single polypeptide chain of approximately 175 amino acids (Jehle et al., 2010). 3.2. Protein sequence alignment and phylogenetic analysis of AccHsp24.2 and AccHsp23.0 To further validate the conserved domain, a multiple-sequence alignment was performed using the ClustalX program (Fig. 2). The predicted amino acid sequence of AccHsp24.2 exhibited 43.64%, 42.00%, 52.63%, 48.33%, and 55.72% similarity to Hsp27.6 from A. cerana cerana (GenBank accession number: GQ254650), Hsp20.7 from Locusta migratoria (GenBank accession number: DQ355965), sHSP from Macrocentrus cingulum (GenBank accession number: EU624206), Hsp21.7 from Nasonia vitripennis (GenBank accession number: XP001604512), and AccHsp23.0, respectively. AccHsp23.0 was 44.07%, 49.74%, 62.62%, 64.43%, and 55.72% identical to AccHsp27.6, LmHsp20.7, McsHsp, NvHsp21.7, and AccHsp24.2, respectively. The protein sequence comparison also showed that two conserved regions are present in these sHSPs. The first 27 amino acids constitute the first conserved region, which is highly hydrophobic and might be involved in modulating oligomerization, subunit dynamics, and substrate binding (Sun and MacRae, 2005). The second region, which includes the α-crystallin domain, is prominent and determines the molecular chaperone function of the sHSPs. As shown in Fig. 3 the two sHSP genes cloned appeared in the same cluster and were orthologous to the sHSP genes from hymenopteran species (McsHsp, NvHsp21.7, and AccHsp27.6). In addition, there was an orthologous cluster containing several sHSPs from different insect orders, which is shaded in Fig. 3. This orthologous cluster suggests that these sHSPs may have evolved prior to the divergence of the species (Kokolakis et al., 2008).

3. Results 3.1. Cloning and sequence analysis of AccHsp24.2 and AccHsp23.0 A sequence analysis indicated that the cDNA sequence of AccHsp24.2 (GenBank accession no. JF411009) obtained from A. cerana cerana is 868 bp in length and contains a 146-bp 5′ untranslated region (UTR) and a 74-bp 3′ UTR. The 648-bp ORF encodes a protein of 216 amino acids with a predicted molecular weight of 24.2 kDa and an estimated pI of 4.93. The cDNA sequence of AccHsp23.0 (GenBank accession no. JF411010) is 780 bp in length and contains a 124-bp 5′ UTR and a 53-bp 3′ UTR. The 603-bp ORF encodes a protein of 201 amino acids with a predicted molecular weight of 23.0 kDa and an estimated pI of 5.21. The genomic sequences of these two genes were also obtained, and no introns were present. Using the NCBI Conserved Domain (CD) search, the α-crystallin domains (ACDs) of metazoan α-crystallin-type sHSPs were found in the deduced amino acid sequences of AccHsp24.2 and AccHsp23.0, suggesting that the two genes might belong to the α-crystallin/sHSP family. An analysis of the secondary structure indicated that the helical contents of AccHsp24.2 and AccHsp23.0 are predicted to be 3.26% and 8.5%, respectively, with strand contents of 20.93% and 24.5%, respectively. β-Sheets were found to occur frequently in the α-crystallin domain, which is

Fig. 1. The tertiary structure of α-crystallin domain residues.

Please cite this article as: Liu, Z., et al., Two small heat shock protein genes in Apis cerana cerana: characterization, regulation, and developmental expression, Gene (2014), http://dx.doi.org/10.1016/j.gene.2014.05.034

Z. Liu et al. / Gene xxx (2014) xxx–xxx

5

Fig. 2. Amino acid sequence comparison of sHSP homologs. The asterisks, double dots, and single dots denote fully, strongly, and weakly conserved residues, respectively. The gray sections show the two conserved regions, and the α-crystallin domain is underlined. The secondary structure assignments of AccHsp24.2 and AccHsp23.0 are shown above the sequences; the arrowheads, ribbons, and dotted lines denote helices, strands, and coils, respectively.

3.3. Characterization of the 5′-flanking region of AccHsp24.2 and AccHsp23.0 We identified the 5′-flanking regions to investigate the regulation of the transcription of AccHsp24.2 and AccHsp23.0. Using inverse PCR, we obtained a 727-bp and an 885-bp DNA fragment upstream of the translation start site of AccHsp24.2 and AccHsp23.0, respectively. As shown in Fig. 4, three HSEs were detected in AccHsp24.2 and eleven were detected in AccHsp23.0. In addition, a TATA-box, representing the putative core promoter element upstream of the transcription start site, was found in both genes. We also searched for transcription factor binding sites using Matinspector, and two Broad-Complex (BR-C) genes for ecdysone steroid response elements were found in each gene. BR-C is a key regulator of metamorphosis and primary ecdysone response genes (von Kalm et al., 1994), which play an important role in insect metamorphosis (Uhlirova et al., 2003). The specific expression of this transcription factor in epidermal tissue is related to the development of the pupal skin (Zhou and Riddiford, 2001). The presence of DBRC sites suggested that AccHsp24.2 and AccHsp23.0 are related to the regulation of the ecdysone cascade.

3.4. Developmental and tissue-specific expression patterns of AccHsp24.2 and AccHsp23.0 As shown in Fig. 5, the trends of the developmental expression patterns of the AccHsp24.2 and AccHsp23.0 genes were similar. The amount of AccHsp24.2 mRNA increased rapidly from the fourth-instar larval to the sixth-instar larval stages, decreased from the sixth-instar larval to the pink-eyed pupal stages, increased slightly at the dark-eyed pupal stage, and showed an increasing trend at the adult stage. The amount of AccHsp23.0 mRNA reached a peak at in fourth-instar larvae and decreased in pink-eyed pupae, and, similar to AccHsp24.2, then increased slightly in dark-eyed pupae, increasing at the adult stage. The tissuespecific expression analysis indicated high transcript levels of the two genes predominantly in the head and muscle. 3.5. Heat stress and cold stress regulation of AccHsp24.2 and AccHsp23.0 The expression levels of the two genes under different temperature conditions were evaluated using RT-qPCR, and the results indicated that

Please cite this article as: Liu, Z., et al., Two small heat shock protein genes in Apis cerana cerana: characterization, regulation, and developmental expression, Gene (2014), http://dx.doi.org/10.1016/j.gene.2014.05.034

6

Z. Liu et al. / Gene xxx (2014) xxx–xxx

Fig. 3. Phylogenetic analysis of sHSP amino acid sequences from different insect species (Diptera, Hymenoptera, and Lepidoptera). The full names of the species and the accession numbers of the genes are designated with the following abbreviations: DmHsp26 (Drosophila melanogaster, NM079273), DmHsp27 (D. melanogaster, NM079276), DmCG14207 (D. melanogaster, NM134482), LhHsp20 (Liriomyza huidobrensis, DQ452370), LsHsp21.7 (Liriomyza sativae, DQ452372), McsHsp (Macrocentrus cingulum, EU624206), NvHsp21.7 (Nasonia vitripennis, XP001604512), NvHsp20.6 (N. vitripennis, XM001607625), AccHsp27.6 (Apis cerana cerana, GQ254650), AccHsp24.2 (A. cerana cerana, JF411009), AccHsp23.0 (A. cerana cerana, JF411010), BmHsp20.1 (Bombyx mori, AB195971), BmHsp23.7 (B. mori, AB195973), BmHsp20.4 (B. mori, AF315318), BmHsp22.6 (B. mori, FJ602788), BmHsp21.4 (B. mori, AB195972), AgHsp20.9 (Anopheles gambiae, XM560153), TcHsp21.8 (Tribolium castaneum, XM968592), AmHsp25.6 (Apis mellifera, XM392405), and LmHsp20.6 (Locusta migratoria, DQ355964).

the two genes could be induced by both heat shock and cold shock. The expression patterns of the two genes were similar, but the level of AccHsp24.2 was notably higher than AccHsp23.0. As shown in Fig. 6, after incubation at 37 °C, the amount of mRNA of both genes showed

no significant change but increased rapidly during the recovery period at 34 °C. In response to 43 °C, the expression of the two genes increased rapidly during the first 30 min and then decreased to a low level; during the recovery period at 34 °C, the expression remained high and showed

Fig. 4. The nucleotide sequence and putative transcription factor binding sites of the 5′-flanking regions of AccHsp24.2 and AccHsp23.0. The transcription and translation start sites are indicated with arrowheads. The underlined sections indicate the DBRC binding sites. The HSEs and the TATA-box are highlighted in light gray.

Please cite this article as: Liu, Z., et al., Two small heat shock protein genes in Apis cerana cerana: characterization, regulation, and developmental expression, Gene (2014), http://dx.doi.org/10.1016/j.gene.2014.05.034

Z. Liu et al. / Gene xxx (2014) xxx–xxx

7

Fig. 5. Expression patterns of AccHsp24.2 and AccHsp23.0 at different developmental stages and in various tissues. (A) Expression patterns of AccHsp24.2 and AccHsp23.0 at different developmental stages. Total RNA was extracted from whole honeybees at fourth (L4) and sixth larval instars (L6), the successive pupal stages, including the white-eyed pupal phase (Pw), pink-eyed pupal phase (Pp), and dark-eyed pupal phase (Pd), and 2- and 10-day-old adult workers (A2 and A10). (B) Expression patterns of AccHsp24.2 and AccHsp23.0 in various tissues. Total RNA was extracted from the dissected head (HE), thorax (TH), midgut (MG), muscle (MS), epidermis (EP), and fat body (FA) of newly emerged bees. The A. cerana cerana βactin gene was used as a reference housekeeping gene to normalize the expression level of the investigated gene in the RT-qPCR experiments. The histograms indicate the relative expression levels. The data represent the means ± SE of three independent experiments. The different letters above the columns indicate significant differences (A, P b 0.05; B, P b 0.01).

increasing trends. In addition to heat stress, we also examined the expression patterns of the two genes under cold shock. As shown in Fig. 6, the expression patterns of the two genes rapidly reached their peak levels in the first 30 min and then decreased to their lowest levels, even during recovery at 34 °C. 3.6. Ecdysone regulation of the AccHsp24.2 and AccHsp23.0 genes According to Zhou and Riddiford (2001), the expression of BR-C mRNA under the control of ecdysone is one of the first molecular events underlying the pupal commitment of the epidermis. As we detected two BR-C sites in the 5′-flanking regions of both AccHsp24.2 and AccHsp23.0, we speculated that the expression of the two genes might be regulated by ecdysone. To verify this hypothesis, we studied the expression pattern of the two genes in the pupal epidermis following induction by ecdysone. As shown in Fig. 7A, the results of real-time PCR indicated that the two genes exhibited similar expression patterns. The expression levels of the two genes decreased significantly after the removal of 0.1 μg/ml and 0.01 μg/ml ecdysone; however the removal of 0.001 μg/ml ecdysone did not cause a notable change in expression. We also detected the ecdysone regulation of AccHsp24.2 and AccHsp23.0 in vivo. We added three different titers of ecdysone to the diet of L2 to L6 larvae and examined the expression of the two genes using real-time PCR. As shown in Fig. 7B, all three titers of ecdysone could upregulate the two genes, though different expression patterns were observed. The expression of AccHsp24.2 only slightly responded to 0.1 μg/ml ecdysone in the ration, which was the lowest of the three titers, whereas the expression of AccHsp23.0 reached extremely high levels in response to this dose. 3.7. Molecular chaperone activity of recombinant AccHsp24.2 and AccHsp23.0 proteins The two genes were expressed in E. coli BL21 cells as 6× His-tagged fusion proteins. After induction with IPTG at 37 °C, an SDS-PAGE analysis detected bands corresponding to the two 6 × His fusion proteins (Fig. 8A, lanes 1 and 4); no protein induction was observed in the non-induced controls (Fig. 8A, lanes 2 and 3). The results indicated that the two genes were expressed, and the fusion proteins were then purified (Fig. 8B, lanes 1 and 2). The α-crystallin domain is involved in the structure of sHsp and not in the determination of the chaperone function. Indeed, some sHsps

that carry the α-crystallin domain have no chaperone activity (Kokke et al., 1998). Therefore, the molecular chaperone activities of the purified recombinant AccHsp24.2 and AccHsp23.0 proteins were investigated in vitro using the MDH thermal aggregation assay. As shown in Fig. 9, MDH aggregated when the solution was incubated at 43 °C, as indicated by an increase in dispersed light at 320 nm. In addition, MDH aggregated faster when mixed with BSA, which has no chaperone activity. However, the addition of an equivalent amount of AccHsp24.2 or AccHsp23.0 to MDH almost completely suppressed MDH aggregation (Fig. 9). Therefore, the two recombinant proteins can act as efficient molecular chaperones in vitro.

4. Discussion In a previous study, we cloned and characterized the first sHSP gene (AccHsp27.6) in the Chinese honeybee (A. cerana cerana). In the present study, we isolated two other sHSP genes, and the amino acid sequence, structure, presence of the α-crystallin domain, and in vitro molecular chaperone activity of the recombination proteins indicated that the two genes belong to the α-crystallin/sHSP family. A homologous sequence alignment showed that these two sHSP genes share a high sequence homology to AccHsp27.6 (Fig. 2). Accordingly, the genes were named AccHsp24.2 and AccHsp23.0. The protein sequence comparison also showed that two conserved regions are present in the sHSPs. The first 27 amino acids constitute the first conserved region, which is highly hydrophobic and might be involved in modulating oligomerization, subunit dynamics, and substrate binding (Sun and MacRae, 2005). The second region, which includes the α-crystallin domain, is prominent and determines the molecular chaperone function of the sHSPs. While in the phylogenetic analysis of the two genes, the orthologous cluster suggests that these sHSPs may have evolved prior to the divergence of the species (Kokolakis et al., 2008). These two genes also had similar developmental and heat shock expression profiles. In addition, both were induced by ecdysone, suggesting that the two genes might be regulated by ecdysone. Furthermore, we also expressed the recombinant AccHsp24.2 and AccHsp23.0 proteins and detected a remarkable in vitro molecular chaperone activity. Heat shock elements (HSEs), which are recognized by heat shock factors (HSFs), are the most critical elements in HSP gene transcription. We detected three putative HSEs in the 5′-flanking region of AccHsp24.2 and eleven in AccHsp23.0. According to a report on Drosophila Hsp22,

Please cite this article as: Liu, Z., et al., Two small heat shock protein genes in Apis cerana cerana: characterization, regulation, and developmental expression, Gene (2014), http://dx.doi.org/10.1016/j.gene.2014.05.034

8

Z. Liu et al. / Gene xxx (2014) xxx–xxx

Fig. 7. Expression patterns of AccHsp24.2 and AccHsp23.0 in response to different titers of ecdysone in vitro and in vivo. (A) Expression patterns of AccHsp24.2 and AccHsp23.0 in response to different titers of ecdysone in vitro. Total RNA was extracted from the epidermis of new pupae that were incubated in vitro in an incubation medium containing different titers of ecdysone (0.1, 0.01, and 0.001 μg/ml) for 3 h and were then washed for 3 h. (B) Expression patterns of AccHsp24.2 and AccHsp23.0 in response to different titers of ecdysone in vivo. Total RNA was extracted from L6 bees fed an artificial diet containing different titers of ecdysone (0, 0.01, 0.1, and 1 μg/ml) for 96 h. The A. cerana cerana β-actin gene was used as a reference housekeeping gene to normalize the expression level of the investigated gene in the RT-qPCR experiments. The histograms indicate the relative expression levels. The data represent the means ± SE of three independent experiments. The different letters above the columns indicate significant differences (P b 0.01).

Fig. 6. Expression patterns of AccHsp24.2 and AccHsp23.0 in response to different temperatures. Total RNA was extracted from adult bees (12 day) treated for the indicated times at 37 °C (A), 43 °C (B) and 4 °C (C). The A. cerana cerana β-actin gene was used as a reference housekeeping gene to normalize the expression level of the investigated gene in the RTqPCR experiments. The histograms indicate the relative expression levels. The data represent the means ± SE of three independent experiments. The different letters above the columns indicate significant differences (P b 0.01).

three functional HSEs in the 5′-flanking region are required for the proper expression of Hsp22 following stress (Morrow et al., 2004b). However, the HSEs responsible for the heat induction of AccHsp24.2 and AccHsp23.0 have not yet been determined. According to previous reports, sHSPs can prevent the caspasedependent apoptosis that occurred during heat and cold stress in Drosophila cells (Concannon et al., 2003; Yi et al., 2007). Some studies have analyzed the expression of the sHSPs in response to heat shock in insects. For example, the expression of the Drosophila Hsp23 gene could be induced several times by heat, over a broad temperature range (30–37 °C), reaching its maximum level at 35 °C (Kokolakis et al., 2009). However, only a few studies have investigated the expression of sHSPs in response to cold stress. In non-diapausing flesh flies, the expression of Hsp23 was induced by both heat and cold shock (43 and − 10 °C) (Yocum et al., 1998). In the present study, we observed the

notable upregulation of AccHsp24.2 and AccHsp23.0 in response to 37 °C during recovery for 60 min. Similarly, we also found a remarkable upregulation during incubation at 43 °C and throughout the recovery process. These results suggested that sHSPs could markedly improve the insect's thermal tolerance. The upregulation of AccHsp24.2 and AccHsp23.0 was only observed at 30 min in response to 4 °C, in agreement with a previous report that showed that there was no modulation of Hsp23 transcription during the recovery from a short cold stress at 0 °C for 3 h in Drosophila (Sinclair et al., 2007). However, Qin et al. (2005) showed that a cold stress of only 2 h at 0 °C resulted in the upregulation of Hsp23 and Hsp26 in D. melanogaster during a 30-min recovery phase; according to Colinet et al. (2010), the four sHSP genes of D. melanogaster were all upregulated after cold stress at 0 °C for 9 h. In addition to their molecular chaperone functions, sHSPs are involved in various physiological processes, some of which are important for insect cold tolerance. For example, sHSPs help protect the integrity of the actin cytoskeleton and microfilaments, which is important for insect cold tolerance because increasing evidence has shown that cytoskeletal components are involved in insect cold tolerance (Colinet et al., 2007; Kim and Denlinger., 2009). In addition to heat induction, we also observed the expression of AccHsp24.2 and AccHsp23.0 during normal development in this study. The two genes exhibited similar expression patterns during development, suggesting that they are controlled by the same regulatory mechanisms, except in the fourth larval instar. AccHsp23.0 RNA reached its peak level at the fourth larval instar, whereas the RNA level of AccHsp24.2 was very low at the fourth larval instar and peaked at the sixth larval instar. After the larval stage, the two genes had different maximum RNA levels,

Please cite this article as: Liu, Z., et al., Two small heat shock protein genes in Apis cerana cerana: characterization, regulation, and developmental expression, Gene (2014), http://dx.doi.org/10.1016/j.gene.2014.05.034

Z. Liu et al. / Gene xxx (2014) xxx–xxx

Fig. 8. Expression in E. coli and the purification of recombinant AccHsp24.2 and AccHsp23.0 proteins as analyzed by SDS-PAGE. Lane A1, induced AccHsp24.2 recombinant protein; lanes A2 and A3, non-induced; lane A4, induced AccHsp23.0 recombinant protein; lane B1, purified AccHsp24.2 recombinant protein; lane B2, purified AccHsp23.0 recombinant protein; lane M, molecular mass standard.

one of which was at the dark-eyed pupal stage and the other in 10-day post-emergent adults. According to Kokolakis et al. (2009), medfly Hsp23 RNA levels peaked at the third larval instar and remained very high during most of larval development, decreased significantly at the end of the larval stage, increased again at puparium formation and the prepupal stages, and increased during the aging of adults. The expression patterns of AccHsp24.2 and AccHsp23.0 were similar to the two medfly Hsp23 genes, suggesting analogous regulation of these genes in the two species during development. In Drosophila, the expression of the Hsp27 gene during embryonic, late larval, and prepupal development is correlated with the three major peaks in the titer of the molting hormone, ecdysone. This finding suggests that the hormone may be responsible for most of the developmental expression of this gene (Kokolakis et al., 2008). In our study, we detected two BR-C sites in the 5′-flanking regions of AccHsp24.2 and AccHsp23.0. Broad-Complex is an early gene induced by ecdysone and is the key transcription factor in ecdysone cascade regulation. The specific expression of Broad-Complex in the epidermal tissue is related to the formation of the pupal epidermis. It has been reported that mutated Drosophila larvae without all BR-C isomers could only develop to 3 days old and could not form chrysalises (Zheng et al., 2010). According to Ireland et al. (1982), the activation of Drosophila small hsp genes in the late larval and early pupal stages may be regulated by the steroid hormone ecdysterone. Therefore, we examined how ecdysone affected the transcription of the AccHsp24.2 and AccHsp23.0 genes in vitro and in vivo and found that a high titer of ecdysone could significantly upregulate the expression of the two small HSP genes in the pupal epidermis in vitro. This result was consistent with Mestril et al. (1986), who showed that all four Drosophila small hsps (Hsp22, Hsp23, Hsp26, and Hsp27) were synthesized in elevated quantities during puparium formation, a stage when the ecdysterone titer is high. However, treatment with 0.001 μg/ml ecdysone did not improve the expression of AccHsp24.2 and AccHsp23.0, suggesting that the residual ecdysone levels in the epidermis were approximately 0.001 μg/ml. In the larva-rearing experiment, both genes were upregulated by all the titers of ecdysone tested, though different patterns of expression were observed, suggesting the existence of different regulatory mechanisms in the larval stage. In this study, we isolated and characterized two small heat shock protein genes, AccHsp24.2 and AccHsp23.0, from A. cerana cerana and focused on their regulation by heat shock and ecdysone. However, it

9

Fig. 9. Molecular chaperone activity of AccHsp24.2 and AccHsp23.0 demonstrated using the malate dehydrogenase thermal aggregation assay. MDH incubated at 43 °C is represented by filled diamonds. MDH incubated with AccHsp24.2 in a molar ratio 1:1 is represented by filled squares. MDH incubated with AccHsp23.0 in a molar ratio 1:1 is represented by filled triangles, and MDH incubated with BSA in a molar ratio 1:1 is represented by filled circles. The absorbance was monitored at 360 nm at regular intervals (10 min) for 1 h.

remains to be determined whether there is a correlation between the heat shock regulation and ecdysone regulation. Conflict of interest The authors declare that they have no conflict of interest. Acknowledgments This work was financially supported through earmarked funds from the China's Agricultural Research System (CARS-45) and special funds for Agro-Scientific Research in the Public Interest (No. 200903006). References Arrigo, A.P., 1998. Small stress proteins: chaperones that act as regulators of intracellular redox state and programmed cell death. Biological Chemistry 379, 19–26. Brandt, G.E., Blagg, B.S., 2009. Alternate strategies of Hsp90 modulation for the treatment of cancer and other diseases. Current Topics in Medicinal Chemistry 9, 1447–1461. Colinet, H., Nguyen, T.T., Cloutier, C., Michaud, D., Hance, T., 2007. Proteomic profiling of a parasitic wasp exposed to constant and fluctuating cold exposure. Insect Biochemistry and Molecular Biology 37, 1177–1188. Colinet, H., Lee, S.F., Hoffmann, A., 2010. Temporal expression of heat shock genes during cold stress and recovery from chill coma in adult Drosophila melanogaster. The FEBS Journal 277, 174–185. Concannon, C.G., Gorman, A.M., Samali, A., 2003. On the role of Hsp27 in regulating apoptosis. Apoptosis 8, 61–70. De Jong, W.W., Caspers, G.J., Leunissen, J.A., 1998. Genealogy of the alpha-crystallin — small heat-shock protein superfamily. International Journal of Biological Macromolecules 22, 151–162. Fu, X., Li, W., Mao, Q., Chang, Z., 2003. Disulfide bonds convert small heat shock protein Hsp16.3 from a chaperone to a non-chaperone: implications for the evolution of cysteine in molecular chaperones. Biochemical and Biophysical Research Communications 308, 627–635. Gkouvitsas, T., Kontogiannatos, D., Kourti, A., 2008. Differential expression of two small Hsps during diapause in the corn stalk borer Sesamia nonagrioides (Lef.). Journal of Insect Physiology 54, 1503–1510. Hendrick, J.P., Hartl, F.U., 1993. Molecular chaperone functions of heat-shock proteins. Annual Review of Biochemistry 62, 349–384. Ireland, R.C., Berger, E., Sirotkin, K., Yund, M.A., Osterbur, D., Fristrom, J., 1982. Ecdysterone induces the transcription of four heat-shock genes in Drosophila S3 cells and imaginal discs. Developmental Biology 93, 498–507. Jehle, S., Rajagopal, P., Bardiaux, B., Markovic, S., Kuhne, R., Stout, J.R., Higman, V.A., Klevit, R.E., van Rossum, B.J., Oschkinat, H., 2010. Solid-state NMR and SAXS studies provide a structural basis for the activation of alphaB-crystallin oligomers. Nature Structural & Molecular Biology 17, 1037–1042.

Please cite this article as: Liu, Z., et al., Two small heat shock protein genes in Apis cerana cerana: characterization, regulation, and developmental expression, Gene (2014), http://dx.doi.org/10.1016/j.gene.2014.05.034

10

Z. Liu et al. / Gene xxx (2014) xxx–xxx

Kim, M., Denlinger, D.L., 2009. Decrease in expression of beta-tubulin and microtubule abundance in flight muscles during diapause in adults of Culex pipiens. Insect Molecular Biology 18, 295–302. Kokke, B.P., Leroux, M.R., Candido, E.P., Boelens, W.C., de Jong, W.W., 1998. Caenorhabditis elegans small heat-shock proteins Hsp12.2 and Hsp12.3 form tetramers and have no chaperone-like activity. FEBS Letters 433, 228–232. Kokolakis, G., Tatari, M., Zacharopoulou, A., Mintzas, A.C., 2008. The hsp27 gene of the Mediterranean fruit fly, Ceratitis capitata: structural characterization, regulation and developmental expression. Insect Molecular Biology 17, 699–710. Kokolakis, G., Kritsidima, M., Tkachenko, T., Mintzas, A.C., 2009. Two hsp23 genes in the Mediterranean fruit fly, Ceratitis capitata: structural characterization, heat shock regulation and developmental expression. Insect Molecular Biology 18, 171–181. Liu, Z., Xi, D., Kang, M., Guo, X., Xu, B., 2012. Molecular cloning and characterization of Hsp27.6: the first reported small heat shock protein from Apis cerana cerana. Cell Stress & Chaperones 17, 539–551. Mestril, R., Schiller, P., Amin, J., Klapper, H., Ananthan, J., Voellmy, R., 1986. Heat shock and ecdysterone activation of the Drosophila melanogaster hsp23 gene; a sequence element implied in developmental regulation. The EMBO Journal 5, 1667–1673. Michaud, S., Morrow, G., Marchand, J., Tanguay, R.M., 2002. Drosophila small heat shock proteins: cell and organelle-specific chaperones? Progress in Molecular and Subcellular Biology 28, 79–101. Mizrahi, T., Heller, J., Goldenberg, S., Arad, Z., 2010. Heat shock proteins and resistance to desiccation in congeneric land snails. Cell Stress & Chaperones 15, 351–363. Morrow, G., Battistini, S., Zhang, P., Tanguay, R.M., 2004a. Decreased lifespan in the absence of expression of the mitochondrial small heat shock protein Hsp22 in Drosophila. The Journal of Biological Chemistry 279, 43382–43385. Morrow, G., Samson, M., Michaud, S., Tanguay, R.M., 2004b. Overexpression of the small mitochondrial Hsp22 extends Drosophila life span and increases resistance to oxidative stress. FASEB Journal 18, 598–599. Pérez-Morales, D., Ostoa-Saloma, P., Espinoza, B., 2009. Trypanosoma cruzi SHSP16: characterization of an α-crystallin small heat shock protein. Experimental Parasitology 123, 182–189. Qin, W., Neal, S.J., Robertson, R.M., Westwood, J.T., Walker, V.K., 2005. Cold hardening and transcriptional change in Drosophila melanogaster. Insect Molecular Biology 14, 607–613. Reineke, A., 2005. Identification and expression of a small heat shock protein in two lines of the endoparasitic wasp Venturia canescens. Comparative Biochemistry and Physiology. Part A, Molecular & Integrative Physiology 141, 60–69. Silva, I.C., Message, D., Cruz, C.D., Campos, L.A., Sousa-Majer, M.J., 2009. Rearing Africanized honey bee (Apis mellifera L.) brood under laboratory conditions. Genetics and Molecular Research 8, 623–629.

Sinclair, B.J., Gibbs, A.G., Roberts, S.P., 2007. Gene transcription during exposure to, and recovery from, cold and desiccation stress in Drosophila melanogaster. Insect Molecular Biology 16, 435–443. Sun, Y., MacRae, T.H., 2005. Small heat shock proteins: molecular structure and chaperone function. Cellular and Molecular Life Sciences 62, 2460–2476. Sun, R., Zhang, Y., Xu, B., 2013. Characterization of the response to ecdysteroid of a novel cuticle protein R&R gene in the honey bee, Apis cerana cerana. Comparative Biochemistry and Physiology. Part B, Biochemistry & Molecular Biology. http://dx.doi.org/10. 1016/j.cbpb.2013.07.002. Tsunoda, T., Takagi, T., 1999. Estimating transcription factor bindability on DNA. Bioinformatics 15, 622–630. Tsvetkova, N.M., Horvath, I., Torok, Z., Wolkers, W.F., Balogi, Z., Shigapova, N., Crowe, L.M., Tablin, F., Vierling, E., Crowe, J.H., Vigh, L., 2002. Small heat-shock proteins regulate membrane lipid polymorphism. Proceedings of the National Academy of Sciences of the United States of America 99, 13504–13509. Uhlirova, M., Foy, B.D., Beaty, B.J., Olson, K.E., Riddiford, L.M., Jindra, M., 2003. Use of Sindbis virus-mediated RNA interference to demonstrate a conserved role of BroadComplex in insect metamorphosis. Proceedings of the National Academy of Sciences of the United States of America 100, 15607–15612. von Kalm, L., Crossgrove, K., Von Seggern, D., Guild, G.M., Beckendorf, S.K., 1994. The Broad-Complex directly controls a tissue-specific response to the steroid hormone ecdysone at the onset of Drosophila metamorphosis. The EMBO Journal 13, 3505–3516. Wang, J., Wang, X., Liu, C., Zhang, J., Zhu, C., Guo, X., 2008. The NgAOX1a gene cloned from Nicotiana glutinosa is implicated in the response to abiotic and biotic stresses. Bioscience Reports 28, 259–266. Yang, G., 2005. Harm of introducing the western honeybee Apis mellifera L. to the Chinese honeybee Apis cerana F. and its ecological impact. Acta Entomologica Sinica 3, 401–406. Yi, S.X., Moore, C.W., Lee Jr., R.E., 2007. Rapid cold-hardening protects Drosophila melanogaster from cold-induced apoptosis. Apoptosis 12, 1183–1193. Yocum, G.D., Joplin, K.H., Denlinger, D.L., 1998. Upregulation of a 23 kDa small heat shock protein transcript during pupal diapause in the flesh fly, Sarcophaga crassipalpis. Insect Biochemistry and Molecular Biology 28, 677–682. Zheng, W.W., Yang, D.T., Wang, J.X., Song, Q.S., Gilbert, L.I., Zhao, X.F., 2010. Hsc70 binds to ultraspiracle resulting in the upregulation of 20-hydroxyecdsone-responsive genes in Helicoverpa armigera. Molecular and Cellular Endocrinology 315, 282–291. Zhou, B., Riddiford, L.M., 2001. Hormonal regulation and patterning of the Broad-Complex in the epidermis and wing discs of the tobacco hornworm, Manduca sexta. Developmental Biology 231, 125–137.

Please cite this article as: Liu, Z., et al., Two small heat shock protein genes in Apis cerana cerana: characterization, regulation, and developmental expression, Gene (2014), http://dx.doi.org/10.1016/j.gene.2014.05.034

Two small heat shock protein genes in Apis cerana cerana: characterization, regulation, and developmental expression.

In the present study, we identified and characterized two small heat shock protein genes from Apis cerana cerana, named AccHsp24.2 and AccHsp23.0. An ...
2MB Sizes 2 Downloads 4 Views