Skip to main content
  • Research article
  • Open access
  • Published:

Basal hsp70 expression levels do not explain adaptive variation of the warm- and cold-climate O3 + 4 + 7 and OST gene arrangements of Drosophila subobscura

Abstract

Background

Drosophila subobscura exhibits a rich inversion polymorphism, with some adaptive inversions showing repeatable spatiotemporal patterns in frequencies related to temperature. Previous studies reported increased basal HSP70 protein levels in homokaryotypic strains for a warm-climate arrangement compared to a cold-climate one. These findings do not match the similar hsp70 genomic organization between arrangements, where gene expression levels are expected to be similar. In order to test this hypothesis and understand the molecular basis for hsp70 expression, we compared basal hsp70 mRNA levels in males and females, and analysed the 5′ and 3′ regulatory regions of hsp70 genes in warm- and cold-climate isochromosomal O3 + 4 + 7 and OST lines of D. subobscura.

Results

We observed comparable mRNA levels between the two arrangements and a sex-biased hsp70 gene expression. The number of heat-shock elements (HSEs) and GAGA sites on the promoters were identical amongst the OST and O3 + 4 + 7 lines analysed. This is also true for 3′ AU-rich elements where most A and B copies of hsp70 have, respectively, two and one element in both arrangements. Beyond the regulatory elements, the only notable difference between both arrangements is the presence in 3′ UTR of a 14 bp additional fragment after the stop codon in the hsp70A copy in five O3 + 4 + 7 lines, which was not found in any of the six OST lines.

Conclusions

The equivalent hsp70 mRNA amounts in OST and O3 + 4 + 7 arrangements provide the first evidence of a parallelism between gene expression and genetic organization in D. subobscura lines having these arrangements. This is reinforced by the lack of important differential features in the number and structure of regulatory elements between both arrangements, despite the genetic differentiation observed when the complete 5′ and 3′ regulatory regions were considered. Therefore, the basal levels of hsp70 mRNA cannot account, in principle, for the adaptive variation of the two arrangements studied. Consequently, further studies are necessary to understand the intricate molecular mechanisms of hsp70 gene regulation in D. subobscura.

Background

Inversions can foster local adaptation as they protect chromosomal inverted sequences from recombination in heterokaryotypes. Drosophila subobscura exhibits a rich inversion polymorphism [1], with the O chromosome (Muller element E that also corresponds to chromosome arm 3R in Drosophila melanogaster ([2], p., 307) having the highest number of inversions that form complex gene arrangements of overlapping and non-overlapping inversions [3]. The overlapping inversions cluster in the distal segment of the chromosome and form the \( {\mathrm{O}}_{\underset{\_}{3+4}} \) phylad, whose different gene arrangements have higher frequencies at lower latitudes [1, 4]. Thus, the gene arrangements on chromosome O can be divided into two groups: group 1 (cold-climate) includes OST that shows a negative correlation coefficient with maximum temperatures along the latitudinal clines, or a positive correlation coefficient with latitude; and group 2 (warm-climate) that includes gene arrangements showing the reverse pattern (e.g., \( {\mathrm{O}}_{\underset{\_}{3+4}} \), \( {\mathrm{O}}_{\underset{\_}{3+4}+7} \) and \( {\mathrm{O}}_{\underset{\_}{3+4+8}} \)). In agreement with the clinal patterns, warm-climate inversions increase in frequency during summer and decrease during winter [5, 6]. Furthermore, they show long-term trends of increasing frequency related to global warming [7, 8]. Overall, these spatiotemporal patterns strongly suggest that temperature is the main selective factor driving inversion frequencies, but the underlying genetic causes are still poorly understood.

The heat-shock HSP70 protein is involved in protein folding, assembly, and can be induced by several stressors such as a sudden increase in temperature [9, 10]. It appears to be the primary protein involved in thermotolerance in D. melanogaster [11, 12], and allele frequencies show latitudinal clines and change in response to thermal evolution in the laboratory [13]. In a previous work with D. subobscura, expression levels of HSP70 protein in flies carrying cold- and warm-climate gene arrangements were analysed [14]. It was found that homokaryotypic \( {\mathrm{O}}_{\underset{\_}{3+4}} \) flies had increased HSP70 basal protein levels that did not boost after a heat-shock, in contrast to what happened to their OST homokaryotypic counterparts. This result was somehow surprising because HSP70 protein expression entails fitness costs [15,16,17]. The interpretation was that low-latitude populations with relatively high \( {\mathrm{O}}_{\underset{\_}{3+4}} \) frequencies are more likely to be exposed to occasionally higher temperatures than populations located at higher latitudes, and the benefits of mounting the stress response might outweigh the costs. Actually, an extreme heat wave in Europe in spring 2011 caused frequencies of warm-climate \( {\mathrm{O}}_{\underset{\_}{3+4}+7} \) to shift +5.4σ from the expected average frequency (in close agreement with the temperature anomaly), and high selection coefficients are needed to explain this shift [18].

This interpretation is not without caveats. A trend to a decrease in basal HSP70 protein expression was observed in strains of D. melanogaster [19, 20] and Drosophila buzzatii [21] continually exposed to high temperatures, probably to avoid the deleterious effects of increased HSP70 levels. On the other hand, increased basal HSP70 protein and mRNA levels have been observed in certain strains of Drosophila virilis with increased thermotolerance [22]. Therefore, basal levels of HSP70 protein expression might be contingent on the thermal history and the species under study. A positive link between hsp70 copy number, expression levels, and thermal resistance in artificial [23,24,25] and natural [26] Drosophila strains has been reported, but this does not apply to D. subobscura. In this species, we have recently shown that the genomic sequence organization of the hsp70 gene family (located in the 94A subsection of the O chromosome, which is inside the region covered by O3 and O4 inversions [27];), is essentially similar in cold- (OST) and warm-climate (\( {\mathrm{O}}_{\underset{\_}{3+4}} \), \( {\mathrm{O}}_{\underset{\_}{3+4+8}} \) and \( {\mathrm{O}}_{\underset{\_}{3+4+16}} \)) gene arrangements, and consists of only two genes, each with four heat-shock elements and three GAGA sites on its promoter [27]. To date, no clear evidence has been found to link the genomic sequence organization of the hsp70 gene family with the different HSP70 protein expression patters previously detected [14].

Knowing that proteins are produced from mRNA which, in turn, depends on gene regulatory regions, we extended the former approach to gauge if hsp70 mRNA levels are similar in cold- and warm-climate arrangements as expected by their similar genomic organization. We quantified basal hsp70 mRNA expression levels in males and females, and included a higher number of isogenic OST and \( {\mathrm{O}}_{\underset{\_}{3+4}+7} \) strains than in our previous study [27]. We expanded the study to the 5′ and 3′ hsp70 regulatory regions, which are known to be involved in transcriptome variation in D. melanogaster [28]. A clarification is in order. The arrangement \( {\mathrm{O}}_{\underset{\_}{3+4}+7} \) has the overlapping inversions \( {\mathrm{O}}_{\underset{\_}{3+4}} \) in the distal segment and the single O7 inversion in the proximal segment of chromosome O. Because our lines derived from a natural population where both gene arrangements (OST and \( {\mathrm{O}}_{\underset{\_}{3+4}+7} \)) have a relatively high frequency (Berbikiz, north of Spain [18]) the heterokaryotype \( {\mathrm{O}}_{\underset{\_}{3+4}}/{\mathrm{O}}_{\underset{\_}{3+4}+7} \) freely recombines in the distal segment and, therefore, the gene content in the \( {\mathrm{O}}_{\underset{\_}{3+4}} \) region is expected to be the same in both arrangements. The reason for using \( {\mathrm{O}}_{\underset{\_}{3+4}+7} \) was because the integrity of the wild chromosome is better preserved during the isogenization process with the available balancer stock in D. subobscura (see below).

We found equivalent mRNA amounts in OST and \( {\mathrm{O}}_{\underset{\_}{3+4}+7} \) arrangements, a sex-biased hsp70 mRNA gene expression compatible with the high degree of conservation of the cis-regulatory elements observed, and an identical genomic organization of the gene family previously reported in our study.

Results

High degree of conservation in the hsp70 proximal and distal CRE elements between OST and O3 + 4 + 7 arrangements

We have sequenced the proximal and distal regulatory regions of the hsp70 locus in 12 isogenic lines (6 lines for OST and 6 for \( {\mathrm{O}}_{\underset{\_}{3+4}+7} \) gene arrangements). The proximal 5′ region consists of a ~ 550 bp, of which ~ 290 bp correspond to the region upstream of the transcription start site (TSS) in the two hsp70 gene copies (hsp70A and hsp70B). Multiple sequence alignments indicated high CRE (Cis-Regulatory Element) conservation between OST and \( {\mathrm{O}}_{\underset{\_}{3+4}+7} \). As previously found [27], four highly conserved heat shock elements (HSEs) were identified (Fig. 1), and no fixed nucleotide differences between arrangements were observed (Additional files 1 and 2). The genomic organization of HSEs resembles that of other Drosophila species [29] and consists of at least three continuous alternate nGAAn/nTTCn pentanucleotide subunits representing binding sites for the heat shock factor (HSF) monomer. Although the number of repetitions is identical in the 12 isogenic lines, it varies between HSE sequences: three in the most distal and proximal (HSE1, HSE4) and four in HSE2 and HSE3 (Fig. 1). Within each HSE, polymorphisms appear to be overrepresented in the first and fifth position of nGAAn/nTTCn pentanucleotides, which are the least constrained [29]. Some insertion/deletion (INDEL) polymorphisms were also detected in the presumably unconstrained regions separating the more distal HSEs. These polymorphisms tend to be shared between the two hsp70 paralogs of the same line as a result of ectopic gene conversion [27]. We did not observe differences between isogenic lines either in the two core proximal promoter elements (TATA box and TSS), or in the total number of GAGA sites in the proximal promoters. The latter is known to be recognized by the transcription factor GAF and to participate, like other promoters, in the regulation of global gene expression [30].

Fig. 1
figure 1

Representation of hsp70 5′ cis-regulatory elements of OST and O3 + 4 + 7 lines. The two copies of the gene (A and B) are included. Green rectangles: HSEs 1–4, where numbers in brackets represent the number of nGAAn or nTTCn pentanucleotide units. Red ellipse: GAGAG (G+) repeat motifs. Blue ellipses: CTCTC (G-) repeat motifs. Orange triangle: TATA box. The scale represents nucleotides relative to the transcription start site (+ 1), indicated by a forward arrow

The distal regulatory region corresponds to the 3′ untranslated mRNA region (3’UTR), which includes a region spanning 198 ± 4 bp in all hsp70A and hsp70B copies, except in one hsp70B copy from OST (4th line) that is 381 bp long. The in silico screening let us identify conserved regulators consisting in AU-rich elements (AREs), known to be involved in the post-transcriptional regulation of gene expression in Drosophila [31] by recruiting ARE-binding proteins that trigger mRNA deadenylation and turnover. In most lines we identified two and one ARE motifs, consisting in a single AUUUA pentamer in hsp70A and hsp70B copies, respectively. There are a few exceptions; namely, the 1st and 6th OST lines that have an unique motif in hsp70A (Fig. 2 and Additional file 3) and the 4th OST line, where the 3’UTR region of the hsp70B copy is longer and has a second ARE motif (Fig. 2 and Additional file 4). We found that ARE motifs are, in general, highly conserved in both copies of OST and \( {\mathrm{O}}_{\underset{\_}{3+4}+7} \), but the hsp70A copy in five \( {\mathrm{O}}_{\underset{\_}{3+4}+7} \) lines has a 14 bp additional fragment after the stop codon, which was not found in OST lines (Additional file 3).

Fig. 2
figure 2

Schematic representation hsp70 3′ UTR region of OST and O3 + 4 + 7 lines. The two copies of the gene (A and B) are included. The region represented spans from the stop codon to the polyadenylation signal. Green rectangles: ARE motifs. In red: polyadenylation signal. Square brackets: number of nucleotides from the stop codon and between ARE motifs

Table 1 summarizes the level of genetic differentiation between chromosomal rearrangements for the complete nucleotide sequence of 5′ and 3′ regulatory regions of A and B copies. Regulatory regions in both copies, except 3’B, exhibited significant genetic differentiation between the two arrangements studied. However, whereas differentiation levels were similar in 5′ regulatory regions for A and B copies, they were significantly higher in 3′ regulatory regions from A copy, where the average number of differences per site (Dxy) was much higher than in the B copy.

Table 1 Genetic differentiation between O3 + 4 + 7 and OST arrangements for 5′ and 3′ regulatory regions

Basal hsp70 mRNA expression levels

hsp70 mRNA levels have been quantified in 67 (out of 72) D. subobscura samples consisting of 3 biological replicates per isogenic line (6 OST and 6 \( {\mathrm{O}}_{\underset{\_}{3+4}+7} \)) and sex, except five lines where only two replicates for females and/or males could be analysed. The ANOVA is shown in Table 2, and results are plotted in Fig. 3. Global hsp70 mRNA levels were about the same between arrangements (P = 0.473) but were higher in females than in males (P = 0.005), particularly for the O3 + 4 + 7 gene arrangement; but the arrangement × sex interaction was not statistically significant (P = 0.296).

Table 2 ANOVA for the effects of chromosome gene arrangements (O3 + 4 + 7 and OST), isochromosomal line, and sex on mRNA relative expression (\( {2}^{-\varDelta {C}_{\mathrm{T}}} \)). Data plotted in Fig. 3
Fig. 3
figure 3

Hsp70 expression in females (black) and males (grey) carrying O3 + 4 + 7 and OST arrangements. Error bars are 95% of confidence intervals

Discussion

Here we have sequenced the proximal and distal regulatory regions of the hsp70 loci and have analysed mRNA expression levels in males and females from 6 isochromosomal lines of warm- (\( {\mathrm{O}}_{\underset{\_}{3+4}+7} \)) and 6 isochromosomal lines of cold-climate (OST) gene arrangements in D. subobscura. Our aim was to shed some light on the basal amounts of hsp70 mRNA in these arrangements that show an almost identical genetic organization in the hsp70 locus [27]. We have now performed for the first time a thorough analysis of 5′ and 3′ hsp70 regulatory regions together with the quantification of mRNA levels in OST and O3 + 4 + 7 arrangements. We found similar hsp70 mRNA levels in both arrangements, in accordance with the identical gene configuration found in our previous work [27], but also found differences due to sex (Table 2). Since transcript levels are under the influence of 5′ and 3′ regulatory elements, we analysed these regions and found that the number of 5′ HSE elements in both \( {\mathrm{O}}_{\underset{\_}{3+4}+7} \) and OST arrangements was also identical, as previously observed in other chromosomal arrangements as O3 + 4 + 8 and O3 + 4 + 16, where these elements evolve in concert through gene conversion [27]. Analysis of 3′ UTR regions also showed no noticeable differences between chromosomal arrangements in the number of ARE elements, considered to be critical regulators of mRNA turnover [31]. However, differences were observed between paralogous genes: in most lines, two ARE motifs were found in hsp70A but only one in hsp70B, which could point to post-transcriptional regulation events responsible for RNA shortening and degradation. Interestingly, the 3’UTRs were already considered as more divergent between paralogs than the coding and proximal promoter regions [27].

Our main aim here was to correlate the genome organization of hsp70 loci between cold- (OST) and a warm-climate (\( {\mathrm{O}}_{\underset{\_}{3+4}+7} \)) gene arrangements with mRNA basal expression patterns. The use of isochromosomal lines precludes any extrapolation to wild, outbred flies. mRNA and protein levels are not always correlated [32], and we did not subject our flies to a thermal stress because inbreeding can interfere with the heat-shock response [33]. Besides, inbred individuals might express conditional lethality under heat stress [34]. Therefore, the present results cannot be directly comparable to those obtained by Calabria et al. [14] because these authors measured both basal and heat-induced HSP70 protein levels in outbred strains (obtained from cyclically permuted reciprocal crosses between isochromosomal lines). The differences observed here in the number of ARE motifs among the two hsp70 copies could contribute to ensure basal amounts of HSP70 protein, basically from one of the gene copies. Moreover, the existence in D. melanogaster of 3 and five ARE motifs, respectively, in the two hsp70 gene clusters compared to D. subobscura could contribute to explain why in the previous work [14] certain basal protein amounts were detected. In this regard, a recent study suggested that differences in acclimation capacity between D. melanogaster and D. subobscura correlate with structural differences in the hsp70 regulatory regions [35]. However, in view of the highly conserved regulatory elements between chromosomal arrangements found here, we cannot easily explain the differences observed in protein amounts between arrangements [14]. These regulatory elements seem to be particularly highly constrained even though significant genetic differentiation between arrangements was detected when regulatory 5′ and 3′ were considered as a whole (Table 1). This is in concordance with a previous study where differences between arrangements at non-coding regions were also observed [36].

Previous findings pointed out that changes in the expression of particular genes adjacent to inversion breakpoints could contribute to an inversion evolutionary success [37, 38]. However, studies using genome wide transcriptome analyses in D. melanogaster [39] and D. pseudoobscura [40] concluded that, whereas inversions affect the expression of numerous genes, little evidence of direct position effects was globally detected at the chromosomal level. In D. subobscura the hsp70 loci are located in subsection 94A of the hypothetical O3 ancestral arrangement. This subsection is inside the derived O4 and OST inversions that gave rise to present day \( {\mathrm{O}}_{\underset{\_}{3+4}} \) and OST gene arrangements, respectively [27]. Using the recently assembled genome of D. subobscura [41] we have estimated that the hsp70 loci are around 0.8 Mb from the closest O4 distal breakpoint and likely too far away as to be directly affected by this breakpoint, an inference that is in agreement with the similar RNA amounts found in the two chromosomal arrangements analysed here. Four genes (Pxd, Set 8, CG5225 and Acf) are located near O4 inversion breakpoints [42], but at present we do not know whether potential position effects affecting these genes play any role in the in the adaptive value of this inversion. In any case, fitness interactions among loci (epistasis) seem to be important because a recombination load (i.e., the loss of fitness caused by recombination between beneficial combinations of interacting alleles) was detected in \( {\mathrm{O}}_{\mathrm{ST}}/{\mathrm{O}}_{\underset{\_}{3+4}+7} \) heterokaryotypes [43].

A potential limitation of our study is that we cannot extrapolate the results to mRNA expression levels during the larval stage. Insects with complex life cycles as D. subobscura might evolve different thermal sensitivities across life stages, particularly between mobile adults (with a higher capacity for thermoregulatory behaviour) and more sessile (eggs, larvae, pupae) life stages. A previous study assessing the thermal evolution of total mRNA expression levels in D. subobscura larvae did not detect differential gene expression for hsp70 among thermal environments, though the highest temperature used in the experiments was likely not high enough as to induce a heat-shock response [44]. We cannot unequivocally generalize our present results to other life stages, but we think that there are not compelling reasons to believe that the basal mRNA expression of hsp70 would have been different between \( {\mathrm{O}}_{\underset{\_}{3+4}} \) and OST gene arrangements had we used larvae instead of adults.

An important finding is the trend towards higher mRNA amounts in females compared to males. Differences between sexes are, in principle, not surprising because in Drosophila a substantial fraction of the transcriptome shows sex-dependent regulation [45]. In our case, the differences could be somehow understandable if the HSP70 protein played sex-specific functions. In D. melanogaster, this protein is expressed during spermatogenesis under heat-shock conditions in adults [46], and at normal rearing temperatures in the larval stage and onwards [16, 47]. In the case of females, a novel role for the molecular chaperone HSP70 in the regulation of border cell migration during Drosophila oogenesis has also been reported [48].

Conclusions

The mRNA levels from isogenic lines for OST and O3 + 4 + 7 gene arrangements in D. subobscura are quantitatively equivalent. This provides some evidence of a parallelism between hsp70 mRNA expression and its genetic organization. Therefore, the basal levels of hsp70 mRNA cannot, in principle, explain the spatiotemporal patterns observed for these two gene arrangements. A potential drawback of our study is that we cannot extrapolate the results to wild, outbred flies. Further experiments are necessary to understand the actual role (if any) played by hsp70 genes in the adaptive value of the rich chromosomal polymorphism on the distal segment of chromosome O, where the hsp70 genes are located.

Methods

Base stocks and fly handling

Wild D. subobscura males were collected by net-sweeping over banana baits fermented with baker’s yeast in Berbikiz (north of Spain; 43°11′20.31″ N, 3°5′23.74″ W) in 2012. They were individually crossed to three or four virgin females from the ch-cu marker strain, which carries the recessive markers on the O chromosome cherry eyes (ch) and curled wings (cu) and has a highly homogeneous genetic background, to identify the gene arrangements of one set of the five major wild-type chromosomes (the other set of homologous chromosomes coming from the ch-cu strain). Thereafter, offspring males from each line were crossed to females from the Va/Ba (Varicose/Bare) balancer stock [49], which has the same genetic background than the ch-cu strain, to obtain isochromosomal lines for the O chromosome in an otherwise highly homogeneous genetic background.

A total of 6 OST and 6 \( {\mathrm{O}}_{\underset{\_}{3+4}+7} \) isogenic lines were used in the experiments. All 12 lines were reared at 18 °C (the optimum temperature [50]) with a 12:12 light/dark cycle on David’s killed-yeast Drosophila medium [51]. To obtain the experimental flies, eggs from the different lines were placed in 2 × 8 cm vials (45 eggs/vial) with 6 mL of food, which guaranteed low or negligible larval crowding that could have triggered the hsp70 stress response [20]. For each isogenic line tested, two independent sets of an average of three biological replicates, each containing three 7-day old adult males and females, were used for mRNA analyses.

Characterization and sequence analysis of the hsp70 5′ and 3′ regulatory regions

For each isogenic line, two genomic regions expanding ~ 300 bp upstream the transcription starting site and ~ 205 bp downstream the stop codon were sequenced for the two hsp70 gene copies (hsp70B and hsp70A; see [27]). The primers used for amplification of 3′ hsp70B and hsp70A regulatory regions were, respectively: P154-proximal (5′- ACC CGC AAA ATT GAA CCC AA-3′), CD_1_F (5′-ACC ATA CAG AAC GAC AAG GGT- 3′) and DMT_proximal (5′-AGT CGG AAT TGT GAA GCC TT-3′). Due to the SNPs (Single nucleotide polymorphisms) found in 5’regulatory regions between lines, different primer sets were used: a) hsp70B: ARN_1R (5′-TGT GAT GCT TTG GCC CAG AT-3′) and D5_NP (5′-GAT CCT CAA GGA ATG CGA AA-3′)/D7_NP(5′-CTG TAG ATC TTC AGG CAT AAG CTG-3′)/ NP-1-FW (5′ GTA TTG TGG CTT CTT AAC GAG GTTC C-3′); b) hsp70A: ARN_1R and and D1_NP (5′-CTT CGG TGG GTT GGA TTG TAG ATT C-3′)/ D3_NP(5′-CTA CTT TCT CTC CTG TGT ATG TCT GG-3′)/D4_NP(5′-CTC TCC TGT GTA CAT ATG TCT TGC G-3′). To identify several CREs (Cis-Regulatory Elements) that are typically found in the proximal 5′ flanking region, the sequences of the two hsp70 genes were first separately aligned using the MAFFT E-INS-i algorithm [52] available at: https://mafft.cbrc.jp/alignment/server/. Next, we used the information available on the hsp70 proximal promoters of D. subobscura [27] and other Drosophila species [29] to identify the conserved heat shock elements in the sequences of the 12 isogenic lines (HSEs 1–4), and the Eukaryotic Neural Network Promoter Prediction software (http://www.fruitfly.org/seq_tools/promoter.html) to find other CREs such as GAGA binding sites and some core promoter elements, including the transcription start site (TSS) and the TATA box. 3′ UTR sequences were aligned using the MAFFT E-INS-i algorithm [52] and manually refined using the Genedoc alignment editor (version 2.7) [53]. The web-based ARE score application (http://arescore.dkfz.de/arescore.pl) was used to identify post-transcriptional regulators of mRNA levels such as AU-rich elements [54]. The DNASp version v6.12 program [55] was used for genetic differentiation analyses between arrangements, assessed using Dxy [56] and Fst [57], and the significance for Snn [58] was obtained with 1000 replicates. Analyses were performed excluding gaps.

Quantification of hsp70 transcripts by qRTPCR

Total RNA extraction was performed separately in adult males and females using the Trizol Lysis reagent (Thermo Fisher Scientific Inc.) following manufacturer’s instructions, and then treated with 1 μl (2 U/μl) of Ambion DNase I (Thermo Fisher Scientific Inc.) 4 h at 37 °C. Concentration of DNA-free samples was checked in a Nanodrop-2000 spectrophotometer (Thermo Scientific Nanodrop) and adjusted to ~ 1 μg/μL and then cDNA synthesis was carried out with anchored-Oligo (dT)18 primers using the Transcriptor First Strand cDNA Synthesis kit (Roche Diagnostics SL, SPAIN). Gene expression was quantified by measuring fluorescence intensity using iQ SYBR Green Supermix (BioRad, Hercules CA, USA) on a CFX96 BioRad Real-Time lightcycler at 59 °C using hsp70 specific primers. All assays were performed in three technical replicates, and the rp49 housekeeping gene was used as an endogenous control because this gene showed to be expressed equally in all samples analysed (P = 0.772, see Additional file 5). Relative expression of hsp70 was then calculated with the comparative CT method [59] and ΔCT values for the set of samples are provided in Additional file 6.

For each isogenic line, we amplified a conserved hsp70 coding fragment of 226 bp, common to both hsp70 copies described in D. subobscura [27]. The primers used for amplification were: HSP70_1L (5′-CAC AGT CTT TGA CGC CAA GC-3′) and HSP70_1R (5′-TGT GAT GCT TTG GCC CAG AT-3′). For the rp49 housekeeping gene, the primers RP49F (5′-ACA TCG GTT ATG GCT CCA C-3′) and RP49R (5′-GAT TTC CTT GCG CTT CTT TG-3′) were designed from a D. subobscura GenBank sequence (Accession: AJ228921), amplifying a 164 bp segment from the gene’s second exon. Primer efficiencies of hsp70 and rp49, calculated on a standard curve slope using serially diluted templates, were 98 and 99.6%, respectively.

Statistical methods

Hsp70 mRNA expression levels were analysed using the linear model.

yijkl = μ + Ai + Lj(i) + Sk + ASik + εijkl, (1)

where μ is the overall grand mean, Ai is the fixed effect of gene arrangement (OST, \( {\mathrm{O}}_{\underset{\_}{3+4}+7} \)), Lj(i) is the random effect of the jth isochromosomal line (j = 1, 2, , 6) within arrangement i, Sk is the fixed effect of sex, and εijkl is the residual error associated with the mRNA expression levels of the lth replicate from the kth sex with the ith gene arrangement from the jth isochromosomal line. Following Schmittgen and Livak [59], statistical tests for mRNA relative expression were performed after the \( {2}^{-\varDelta {C}_{\mathrm{T}}} \) transformation.

To avoid potential problems with the standard assumption that errors are independent, and identically distributed as normal random variables with common variance and zero expectation, we also performed permutation tests [60, 61] to check the validity of parametric P-values. In model (1) the permutation test is a four-stage process. First, random permutation among lines, within arrangements, for the among-line Fline(arrangement); second, random permutation among lines and arrangements for the between-arrangements Farrangement; third, random permutation between sexes for the between-sex Fsex; and fourth, random permutation between arrangements and sexes for the arrangement × sex interaction Farrangement × sex. In each case, 9999 random allocations plus the observed F-values were obtained, which gave us 10,000 values of each F-statistics for each dependent variable. We then calculate the percentage of replications under the null hypothesis that provide as large a value of each F as the obtained F.

The equal expression of rp49 gene was tested on Ct values using one-way ANOVA with four sample groups: females OST, males OST, females \( {\mathrm{O}}_{\underset{\_}{3+4}+7} \) and males \( {\mathrm{O}}_{\underset{\_}{3+4}+7} \). The Levene’s test was used to test the homogeneity of variances between groups.

Analyses were performed using the statistical software STATISTICA 8.0 [62]. Permutation tests were implemented in the algebra environment MATLAB using tools supplied by the Statistics Toolbox [63].

Availability of data and materials

The data analysed in this study are included in this article and its additional files. All newly reported sequences in this study were deposited in GenBank/EMBL/DDBJ database under accession numbers: MK862365- MK862404.

Abbreviations

ARE:

AU-rich element

CREs:

Cis-regulatory elements

HSEs:

Heat shock elements

HSF:

Heat shock factor

SNPs:

Single nucleotide polymorphisms

UTR:

Untranslated region

References

  1. Menozzi P, Krimbas CB. The inversion polymorphism of D. subobscura revisited: Synthetic maps of gene arrangement frequencies and their interpretation. J Evol Biol. 1992;5:625–41.

    Article  Google Scholar 

  2. . Powell JR: Progress and Prospects in Evolutionary Biology. The Drosophila Model, New York: Oxford Univ Press; 1997.

  3. Krimbas CB. Drosophila subobscura, Biology, Genetics and Inversion Polymorphism. Hamburg: Verlag Dr. Kovač; 1993.

    Google Scholar 

  4. Balanyà J, Serra L, Gilchrist GW, Huey RB, Pascual M, Mestres F, Solé E. evolutionary pace of chromosomal polymorphism in colonizing populations of Drosophila subobscura: an evolutionary time series. Evol. 2003;57:1837–45.

    Article  Google Scholar 

  5. Fontdevila A, Zapata C, Alvarez C, Sánchez L, Méndez J, Enriquez I. Genetic coadaptation in the chromosomal polymorphism of Drosophila subobscura. I. Seasonal changes of gametic disequilibrium in a natural population. Genetics. 1983;105:935–55.

    CAS  PubMed  PubMed Central  Google Scholar 

  6. Rodríguez-Trelles F, Alvarez G, Zapata C. time-series analysis of seasonal changes of the O inversion polymorphism of Drosophila subobscura. Genetics. 1996;142:179–87.

    PubMed  PubMed Central  Google Scholar 

  7. Rodríguez-Trelles F, Rodríguez MA. rapid micro-evolution and loss of chromosomal diversity in Drosophila in response to climate warming. Evol Ecol. 1998;12:829–38.

    Article  Google Scholar 

  8. Balanyá J, Oller JM, Huey RB, Gilchrist GW, Serra L. global genetic change tracks global climate warming in Drosophila subobscura. Sci. 2006;313:1773–5.

    Article  CAS  Google Scholar 

  9. Lindquist S, Craig EA. the heat-shock proteins. Ann Rev Genet. 1988;22:631–77.

    Article  CAS  PubMed  Google Scholar 

  10. Lindquist S. the heat-shock response. Annu Rev Biochem. 1986;55:1151–91.

    Article  CAS  PubMed  Google Scholar 

  11. Parsell DA, Taulien J, Lindquist S. the role of heat shock proteins in thermotolerance. Phil Trans Royal Soc London Series B. 1993;339:279–86.

    Article  CAS  Google Scholar 

  12. Feder ME, Blair N, Figueras H. natural thermal stress and heat-shock protein expression in Drosophila larvae and pupae. Funct Ecol. 1997;11:90–100.

    Article  Google Scholar 

  13. Bettencourt BR, Kim I, Hoffmann AA, Feder ME. response to natural and laboratory selection at the Drosophila hsp70 genes. Evol. 2002;56:1796–801.

    Article  CAS  Google Scholar 

  14. Calabria G, Dolgova O, Rego C, Castañeda LE, Rezende EL, Balanyà J, Pascual M, Sørensen JG, Loeschcke V, Santos M. Hsp70 protein levels and thermotolerance in Drosophila subobscura: a reassessment of the thermal co-adaptation hypothesis. J Evol Biol. 2012;25:691–700.

    Article  CAS  PubMed  Google Scholar 

  15. Krebs RA, Loeschcke V. acclimation and selection for increased resistance to thermal stress in Drosophila buzzatii. Genetics. 1996;142:471–9.

    CAS  PubMed  PubMed Central  Google Scholar 

  16. Krebs RA, Feder ME. tissue-specific variation in Hsp70 expression and thermal damage in Drosophila melanogaster larvae. J Exp Biol. 1997;200:2007–15.

    CAS  PubMed  Google Scholar 

  17. Feder JH, Rossi JM, Solomon J, Solomon N, Lindquist S. the consequences of expressing hsp70 in Drosophila cells at normal temperatures. Genes Dev. 1992;6:1402–13.

    Article  CAS  PubMed  Google Scholar 

  18. Rodriguez-Trelles F, Tarrio R, Santos M. Genome-wide evolutionary response to a heat wave in Drosophila. Biol Lett. 2013;9:20130228.

    Article  PubMed  PubMed Central  Google Scholar 

  19. Bettencourt BR, Feder ME, Cavicchi S. experimental evolution of Hsp70 expression and thermotolerance in Drosophila melanogaster. Evol. 1999;53:484.

    Article  CAS  Google Scholar 

  20. Sørensen JG, Loeschcke V. larval crowding in Drosophila melanogaster induces Hsp70 expression, and leads to increased adult longevity and adult thermal stress resistance. J Insect Physiol. 2001;47:1301–7.

    Article  PubMed  Google Scholar 

  21. Sørensen JG, Michalak P, Justesen J, Loeschcke V. expression of the heat-shock protein HSP70 in Drosophila buzzatii lines selected for thermal resistance. Hereditas. 1999;131:155–64.

    Article  PubMed  Google Scholar 

  22. Garbuz D, Evgenev MB, Feder ME, Zatsepina OG. Evolution of thermotolerance and the heat-shock response: evidence from inter/intraspecific comparison and interspecific hybridization in the virilis species group of Drosophila. I. Thermal phenotype. J Exp Biol. 2003;206:2399–408.

    Article  CAS  PubMed  Google Scholar 

  23. Gong WJ, Golic KG. loss of Hsp70 in Drosophila is pleiotropic, with effects on thermotolerance, recovery from heat shock and neurodegeneration. Genetics. 2006;172:275–86.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  24. Feder ME, Cartaño NV, Milos L, Krebs RA, Lindquist SL. effect of engineering Hsp70 copy number on Hsp70 expression and tolerance of ecologically relevant heat shock in larvae and pupae of Drosophila melanogaster. J Exp Biol. 1996;199:1837–44.

    CAS  PubMed  Google Scholar 

  25. Bettencourt BR, Hogan CC, Nimali M, Drohan BW. Inducible and constitutive heat shock gene expression responds to modification of Hsp70 copy number in Drosophila melanogaster but does not compensate for loss of thermotolerance in Hsp70 null flies. BMC Biol. 2007;6:5.

    Article  CAS  Google Scholar 

  26. Evgen'ev MB, Zatsepina OG, Garbuz D, Lerman DN, Velikodvorskaya V, Zelentsova E, Feder ME. evolution and arrangement of the hsp70 gene cluster in two closely related species of the virilis group of Drosophila. Chromosoma. 2004;113:223–32.

    Article  CAS  PubMed  Google Scholar 

  27. Puig Giribets M, García Guerreiro MP, Santos M, Ayala FJ, Tarrío R, Rodríguez-Trelles F. chromosomal inversions promote genomic islands of concerted evolution of Hsp70 genes in the Drosophila subobscura species subgroup. Mol Ecol. 2019;28:1316–32.

    Article  CAS  PubMed  Google Scholar 

  28. Genissel A, LM MI, Wayne ML, Nuzhdin SV. Cis and trans regulatory effects contribute to natural variation in transcriptome of Drosophila melanogaster. Mol Biol Evol. 2008;25:101–10.

    Article  CAS  PubMed  Google Scholar 

  29. Tian S, Haney RA, Feder ME. Phylogeny disambiguates the evolution of heat-shock cis-regulatory elements in Drosophila. PLoS One. 2010;5:e10669.

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  30. Omichinski JG, Pedone PV, Felsenfeld G, Gronenborn AM, Clore GM. the solution structure of a specific GAGA factor-DNA complex reveals a modular binding mode. Nat Struct Biol. 1997;4:122–32.

    Article  CAS  PubMed  Google Scholar 

  31. Cairrao F, Halees AS, KSA K, Morello D, Vanzo N. AU-rich elements regulate Drosophila gene expression. Mol Cell Biol. 2009;29:2636–43.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  32. Becker K, Bluhm A, Casas-Vila N, Dinges N, Dejung M, Sayols S, Kreutz C, Roignant JY, Butter F, Legewie S. Quantifying post-transcriptional regulation in the development of Drosophila melanogaster. Nat Commun. 2018;9:4970.

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  33. Franke K, Fischer K. inbreeding interferes with the heat-shock response. Heredity. 2015;114:80–4.

    Article  CAS  PubMed  Google Scholar 

  34. Vermeulen CJ, Bijlsma R. characterization of conditionally expressed mutants affecting age specific survival in inbred lines of Drosophila melanogaster lethal conditions and temperature-sensitive periods. Genetics. 2004;167:1241–8.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  35. Sørensen JG, Giribets MP, Tarrío R, Rodríguez-Trelles F, Schou MF, Loeschcke V. expression of thermal tolerance genes in two Drosophila species with different acclimation capacities. J Therm Biol. 2019;84:200–7.

    Article  PubMed  CAS  Google Scholar 

  36. Simões P, Pascual M. Patterns of geographic variation of thermal adapted candidate genes in Drosophila subobscura sex chromosome arrangements. BMC Evol Biol. 2018;18:60.

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  37. Puig M, Cáceres M, Ruiz A. silencing of a gene adjacent to the breakpoint of a widespread Drosophila inversion by a transposon-induced antisense RNA. Proc Natl Acad Sci. 2004;101:9013–8.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  38. Huang YC, Dang VD, ChangNC WJ. Multiple large inversions and breakpoint rewiring of gene expression in the evolution of the fire ant social supergene. Proc R Soc B. 2018;285:20180221.

    Article  PubMed  PubMed Central  Google Scholar 

  39. Lavington E, Kern AD. The effect of common inversion polymorphisms In (2L) t and In (3R) Mo on patterns of transcriptional variation in Drosophila melanogaster. G3 (Bethesda). 2017;7:3659–68.

    Article  CAS  Google Scholar 

  40. Fuller ZL, Haynes GD, Richards S, Schaeffer SW. genomics of natural populations: how differentially expressed genes shape the evolution of chromosomal inversions in Drosophila pseudoobscura. Genetics. 2016;204:287–301.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  41. Karageorgiou C, Gámez-Visairas V, Tarrío R, Rodríguez-Trelles F. Long-read based assembly and synteny analysis of a reference Drosophila subobscura genome reveals signatures of structural evolution driven by inversions recombination-suppression effects. BMC Genomics. 2019;20:223.

    Article  PubMed  PubMed Central  Google Scholar 

  42. Puerma E, Orengo DJ, Aguadé M. Multiple and diverse structural changes affect the breakpoint regions of polymorphic inversions across the Drosophila genus. Sci Rep. 2016;6:36248.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  43. Santos M. recombination load in a chromosomal inversion polymorphism of Drosophila subobscura. Genetics. 2009;181:803–9.

    Article  PubMed  PubMed Central  Google Scholar 

  44. Laayouni H, García-Franco F, Chávez-Sandoval BE, Trotta V, Beltran S, Corominas M, Santos M. Thermal evolution of gene expression profiles in Drosophila subobscura. BMC Evol Biol. 2007;10:363.

    Google Scholar 

  45. Ranz JM, Castillo-Davis CI, Meiklejohn CD, Hartl DL. sex-dependent gene expression and evolution of the Drosophila transcriptome. Sci. 2003;300:1742–5.

    Article  CAS  Google Scholar 

  46. Michaud S, Marin R, Westwood JT, Tanguay RM. cell-specific expression and heat-shock induction of Hsps during spermatogenesis in Drosophila melanogaster. J Cell Sci. 1997;110:1989–97.

    CAS  PubMed  Google Scholar 

  47. Lakhotia SC, Prasanth KV. tissue- and development-specific induction and turnover of hsp70 transcripts from loci 87A and 87C after heat shock and during recovery in Drosophila melanogaster. J Exp Biol. 2002;205:345–58.

    CAS  PubMed  Google Scholar 

  48. Cobreros L, Fernández-Miñán A, Luque CM, González-Reyes A, Martín-Bermudo MD. a role for the chaperone Hsp70 in the regulation of border cell migration in the Drosophila ovary. Mech Dev. 2008;125:1048–58.

    Article  CAS  PubMed  Google Scholar 

  49. Sperlich D, Feuerbach-Mravlag H, Lange P. genetic load and viability distribution in central and marginal populations of Drosophila subobscura. Genetics. 1977;86:835–48.

    CAS  PubMed  PubMed Central  Google Scholar 

  50. Rego C, Balanyà J, Fragata I, Matos M, Rezende EL, Santos M. Clinal patterns of chromosomal inversion polymorphisms in Drosophila subobscura are partly associated with thermal preferences and heat stress resistance. Evol. 2010;64:385–97.

    Article  Google Scholar 

  51. David J. A new medium for rearing Drosophila in axenic conditions. Dros Inf Serv. 1962;36:128.

    Google Scholar 

  52. Katoh K, Standley DM. MAFFT multiple sequence alignment software version 7: improvements in performance and usability. Mol Biol Evol. 2013;30:772–80.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  53. Nicholas KB, Nicholas HBJ, DWI D. GeneDoc: Analysis and visualization of genetic variation. EMBNEW NEWS. 1997;4:14.

    Google Scholar 

  54. Spasic M, Friedel CC, Schott J, Kreth J, Leppek K, Hofmann S, Ozgur S, Stoecklin G. Genome-wide assessment of AU-rich elements by the AREScore algorithm. PLoS Genet. 2012;8:e1002433.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  55. Rozas J, Ferrer-Mata A, JC S-DB, Guirao-Rico S, Librado P, Ramos-Onsins SE, Sánchez-Gracia A. DnaSP 6: DNA sequence polymorphism analysis of large data sets. Mol Biol Evol. 2017;34:3299–302.

    Article  CAS  PubMed  Google Scholar 

  56. Nei M. Molecular Evolutionary Genetics. New York: Univ Press; 1987.

    Book  Google Scholar 

  57. Hudson RR, Slatkin M, Maddison WP. estimation of levels of gene flow from DNA sequence data. Genetics. 1992;132:583–9.

    CAS  PubMed  PubMed Central  Google Scholar 

  58. Hudson RR. a new statistic for detecting genetic differentiation. Genetics. 2000;155:2011–4.

    CAS  PubMed  PubMed Central  Google Scholar 

  59. Schmittgen TD, Livak KJ. analyzing real-time PCR data by the comparative CT method. Nat Protoc. 2008;3:1101–8.

    Article  CAS  PubMed  Google Scholar 

  60. Edgington ES. Randomization Tests. 3rd ed. New York: Marcel Dekker; 1995.

    Google Scholar 

  61. BFJ M. Randomization, Bootstrap and Monte Carlo Methods in Biology. 3rd ed. New York: Chapman & Hall/CRC; 2007.

    Google Scholar 

  62. StatSoft, Inc. 2007. STATISTICA (data analysis software system), version 8.0. www.statsoft.com.

  63. MATLAB and Statistics Toolbox Release. 2016. The MathWorks, Inc., Natick.

Download references

Acknowledgements

We thank two anonymous reviewers for comments on the manuscript. Rosa Tarrío and Francisco Rodríguez-Trelles supplied the isogenic D. subobscura lines used in this work.

Funding

This work was supported by the research grants CGL2013-42432P and CGL2017-89160P from the Spanish Ministerio de Ciencia e Innovación, and grants 2014SGR 1346 and 2017SGR 1379 from Generalitat de Catalunya to the Grup de Genòmica, Bioinformàtica i Biologia Evolutiva (GGBE). MPG was supported by a PIF predoctoral fellowship from the Universitat Autònoma de Barcelona (Spain). The founding bodies played no role in the design of the study and collection, analysis, and interpretation of data and in writing the manuscript.

Author information

Authors and Affiliations

Authors

Contributions

MPGG designed the research; MPG and MPGG performed the research and sequence analysis, MS performed statistical analysis. MPG, MPGG and MS wrote the article. All authors have read and approved the manuscript.

Corresponding author

Correspondence to María Pilar García Guerreiro.

Ethics declarations

Ethics approval and consent to participate

Not applicable.

Consent for publication

Not applicable.

Competing interests

The authors declare no competing interests.

Additional information

Publisher’s Note

Springer Nature remains neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Supplementary information

Additional file 1.

Multiple sequence alignment of 5’proximal promoters of hsp70A.

Additional file 2.

Multiple sequence alignment of 5’proximal promoters of hsp70B.

Additional file 3.

Multiple sequence alignment of hsp70A 3’UTR.

Additional file 4.

Multiple sequence alignment of hsp70B 3’UTR.

Additional file 5.

Statistical results for the homogeneous expression of rp49 housekeeping gene.

Additional file 6.

Raw gene expression data.

Rights and permissions

Open Access This article is distributed under the terms of the Creative Commons Attribution 4.0 International License (http://creativecommons.org/licenses/by/4.0/), which permits unrestricted use, distribution, and reproduction in any medium, provided you give appropriate credit to the original author(s) and the source, provide a link to the Creative Commons license, and indicate if changes were made. The Creative Commons Public Domain Dedication waiver (http://creativecommons.org/publicdomain/zero/1.0/) applies to the data made available in this article, unless otherwise stated.

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Puig Giribets, M., Santos, M. & García Guerreiro, M.P. Basal hsp70 expression levels do not explain adaptive variation of the warm- and cold-climate O3 + 4 + 7 and OST gene arrangements of Drosophila subobscura. BMC Evol Biol 20, 17 (2020). https://doi.org/10.1186/s12862-020-1584-z

Download citation

  • Received:

  • Accepted:

  • Published:

  • DOI: https://doi.org/10.1186/s12862-020-1584-z

Keywords