Species' acclimation capacity and their ability to maintain molecular homeostasis outside ideal temperature ranges will partly predict their success following climate change-induced thermal regime shifts. Theory predicts that ectothermic organisms from thermally stable environments have muted plasticity, and that these species may be particularly vulnerable to temperature increases. Whether such species retained or lost acclimation capacity remains largely unknown. We studied proteome changes in the planarian Crenobia alpina, a prominent member of cold-stable alpine habitats that is considered to be a cold-adapted stenotherm. We found that the species' critical thermal maximum (CTmax) is above its experienced habitat temperatures and that different populations exhibit differential CTmax acclimation capacity, whereby an alpine population showed reduced plasticity. In a separate experiment, we acclimated C. alpina individuals from the alpine population to 8, 11, 14 or 17°C over the course of 168 h and compared their comprehensively annotated proteomes. Network analyses of 3399 proteins and protein set enrichment showed that while the species' proteome is overall stable across these temperatures, protein sets functioning in oxidative stress response, mitochondria, protein synthesis and turnover are lower in abundance following warm acclimation. Proteins associated with an unfolded protein response, ciliogenesis, tissue damage repair, development and the innate immune system were higher in abundance following warm acclimation. Our findings suggest that this species has not suffered DNA decay (e.g. loss of heat-shock proteins) during evolution in a cold-stable environment and has retained plasticity in response to elevated temperatures, challenging the notion that stable environments necessarily result in muted plasticity.

Species adapted to cold-stable climates will experience novel selection pressures in response to emerging thermal regimes (Atkins and Travis, 2010; Kenkel et al., 2013). Temperature directly affects the properties of macromolecules, and imposes strong constraints on organisms by affecting the kinetics of chemical reactions. For example, given the rapid decrease in the freedom of molecular motion that accompanies cold temperatures, proteins executing essential cellular processes such as protein synthesis are constrained in their ability to function appropriately (Somero, 1995; 1996). Despite these constraints, life has colonized a wide variety of thermal environments, ranging from volcanic hot vents to thermally stable freshwater habitats. Generating a mechanistic basis for how temperature affects physiological processes and whether organisms from such thermally stable habitats have lost the capacity to compensate for the effects of increasing temperature is becoming more critical (Franklin and Hoppeler, 2021). In concert with adaptive evolution, phenotypic and physiological plasticity is often posited as a central feature of the acute and evolutionary response to climate change (Chevin et al., 2013; Seebacher et al., 2015). The capacity for plasticity, however, depends on the existence of mechanisms to sense and respond to the thermal environment. Both of these mechanisms may either fail to evolve or secondarily be lost in species from thermally stable habitats (Janzen, 1967; Kelly, 2019). As a result, these species are more likely to have narrow ranges of thermal tolerance (i.e. decreased plasticity), but also to exhibit reduced genetic variation in thermal tolerance, making them less likely to evolve wider tolerance over time relative to their eurythermic counterparts (Kelly, 2019). Knowing how species inhabiting thermally stable habitats will respond to rising temperatures is therefore a prerequisite if we are to understand future population dynamics and shifts in community composition (Reusch, 2014; Lamprecht et al., 2018; Rogora et al., 2018). However, there is limited evidence for a relationship between local extinction and limited tolerance to higher temperature (Cahill et al., 2013). The absence of this relationship may be due to the fact that certain organisms from high-elevation habitats are able to maintain homeostasis and stable molecular systems even at higher temperatures than they experience natively (Bernabò et al., 2011, 2020; Hotaling et al., 2020). A variety of molecular responses stabilize and maintain core functional processes such as metabolism, translation, transcription, DNA repair, protein folding, clearance of damaged proteins, mitigation of oxidative stress and regulation of intracellular transport (Thieringer et al., 1998; Morimoto, 2011; De Maayer et al., 2014; Tomanek, 2014). The functional execution and maintenance of these processes results not only from the presence or absence of potential molecular players encoded in the genome, or from stochastic changes in gene expression (McAdams and Arkin, 1997; Elowitz et al., 2002; Raj and van Oudenaarden, 2008), but also from multiple layers beyond the genome (Bludau and Aebersold, 2020), such as the proteome, interactome and whole-organism thermal responses, warranting their study to gain insight into molecular changes close to the phenotype (Feder and Walser, 2005; Karr, 2008).

Remnant populations of certain cold-stenothermal species experienced postglacial retractions and left behind populations in enclaves under relatively benign conditions (Hampe and Jump, 2011; Woolbright et al., 2014). Many of these alpine species are uniquely adapted to cold-stable conditions (Füreder, 1999; Lencioni et al., 2009), including the alpine freshwater flatworm Crenobia alpina (Dana 1766), which supposedly has limited thermal tolerance (Voigt, 1895; Thienemann, 1950). The species occurs primarily where annual average water temperature is <15°C and tolerates temperatures as low as 0.7°C (Voigt, 1895; Reynoldson, 1981; Durance and Ormerod, 2007). Despite this preference for the cold habitat patches of rivers and springs, previous investigations also showed a tolerance to a maximum temperature of 25°C for 48 h without adverse effects (Schlieper and Halsband, 1952) and survival in mountain environments with temperature peaks of 25°C (Pattee, 1965), challenging the classification of this species as a cold-adapted stenotherm.

To arrive at a mechanistic understanding of this potential acclimation capacity of C. alpina, and to test whether the species exhibits decreased thermal plasticity, we provide a detailed analysis of how its proteome changes following acclimation to warmer temperatures. Using an alpine C. alpina population directly from their natal freshwater spring as a model system, we compared the abundance of 3399 proteins after 168 h of acclimation at 8, 11, 14 or 17°C and tested whether sets of temperature-responsive proteins have distinct functional profiles. In a separate experiment, we compared the critical thermal maxima (CT­max­) of individuals from three C. alpina populations originating from springs with distinct altitudes (526, 1080 and 1548 m.a.s.l.) and temperatures during sampling (11, 9 and 8°C), to test whether there are interpopulational differences in CTmax and CTmax acclimation capacity.

Determining CTmax

In all sampling instances, individuals were collected with a 2 or 5 ml pipette and transferred into a cooling box containing water from the respective spring. Animals were transported to the laboratory within 3 h in a mobile fridge. We compared CTmax between one alpine and two non-alpine C. alpina populations by collecting similarly sized individuals from three springs (Fig. S1). Although temperature loggers were placed to determine the thermal breadths of sampling sites, these were not retrievable the following sampling season. Individuals were acclimated to laboratory conditions in flow-through channels (168 h; 8, 9 or 11°C for springs S1, S2 and S3, respectively, based on the temperatures measured at each spring during sampling). Thirty-one individuals per population were then randomly divided in four temperature treatment groups (for 168 h; 8, 11, 14 or 17°C). After this period, individuals were placed in a custom-built oxygenated water bath equipped with a plastic plate containing 2 ml glass vials, with one animal per glass vial. Temperature and dissolved oxygen (always >90%) were monitored throughout each experiment (MultiLine Multi 3650 IDS). The water bath was started at 8°C and heated up continuously to 34°C (0.25°C min−1) using a flow-through water heater (Hydor T08100) controlled by a proportional integral derivative controller (SYL-2352P) via a solid state relay and a K-type thermocouple. We recorded the CTmax value for each individual at the point at which it first curled up and arched its body according to Claussen and Walters (1982).

Proteomics experiment

In a separate experiment, we sampled individuals of similar size (n=100) from the alpine population (August 2020; spring S3 in Fig. S1) and acclimated them for 168 h at 8°C to laboratory conditions. After 168 h, individuals were randomly placed in 8, 11, 14 and 17°C (±0.25°C s.d.) flow-through channels for an additional 168 h. We chose these temperatures as they reflect predicted temperature increases in the Swiss Alps, with 8°C being the control temperature that was measured at the time of sampling, 11°C the IPCC high emission scenario RCP8.5 until mid-21st century, 14°C the RCP8.5 scenario for the end of the 21st century in Switzerland and 17°C an outlier temperature thought to elicit a clearly detectable signal at the proteome level, to better characterize C. alpina's proteome responses to warming. For each treatment temperature, 5 biological replicates (3 animals of similar size pooled per replicate) were kept in separate oxygenated boxes. Temperature, pH, conductivity and oxygen were monitored throughout the experiments and animals were reared on a light (L):dark (D) cycle of L10 h:D14 h. Animals were starved throughout the experiment and each box contained filtered (0.45 µm pore size) water obtained from the spring. At the end of the experiment, individuals were collected with a 2 ml pipette, blotted dry on filter paper, washed with 1× PBS, transferred to 1.5 ml protein LoBind tubes (Eppendorf), shock-frozen in liquid nitrogen and stored at −80°C until further processing.

Proteome analysis

Samples were thawed on ice and whole organisms treated with 5% N-acetyl-l-cysteine (NAC; Sigma-Aldrich) in 1× PBS for 8 min on ice to remove their mucous coating (Pearson et al., 2009). Samples were vortexed and spun down at 5000 rpm for 2 min (at 4°C) and the NAC decanted. Lysates were generated from whole organisms by adding 150 µl lysis buffer [5% SDS, 10 mmol l−1 TCEP, 100 mmol l−1 ABC, supplemented with 1 mmol l−1 HaltTM protease inhibitor cocktail (Thermo Scientific)], homogenizing samples manually with a glass pestle and subjecting them to 10 cycles of ultrasonication (30 s on, 30 s off) using a Bioruptor Pico at 4°C (Diagenode, Inc.). To denature proteins, samples were heated for 10 min at 95°C and mixed at 300 rpm in a PCR 96 thermomixer. At this point, protein concentrations were determined via a bicinchoninic acid assay (BCA; Thermo Scientific Pierce) according to the manufacturer's instructions (Table S1). After letting samples cool down to room temperature (RT), they were spun down at 5000 rpm for 10 s, 0.5 µl iodoacetamide (1 mol l−1) was added and samples were kept in the dark at 25°C for 30 min at 500 rpm. Further analytical steps were performed with an aliquot of no more than 50 µg protein. Following a suspension trapping (S-Trap) protocol (HaileMariam et al., 2018), 2.5 µl 12% phosphoric acid and 165 µl of S-trap buffer (90% methanol, 100 mmol l−1 TEAB, pH 7.1) were added to each sample. Samples were transferred to S-trap micro-columns and washed 3 times by adding 150 µl S-trap buffer followed by centrifugation at 4000 g for 1 min at each wash. Columns were then placed in fresh 2 ml tubes and 20 µl digestion buffer (50 mmol l−1 TEAB, pH 8), supplemented with 1:25 trypsin (Sequencing Grade Modified Trypsin, Promega), was added to each column. Matrix-bound proteins were digested at 47°C for 1 h and the resulting peptides collected by adding 40 µl of digestion buffer to columns and spinning at 4000 g for 1 min. Next, 40 µl of 0.2% formic acid was added to each column, followed by centrifugation at 4000 g for 1 min, then 35 µl of 50% acetonitrile containing 0.2% formic acid was added, and spun down at 4000 g for 1 min. Eluted peptides were concentrated to dryness by applying a vacuum for 2 h. Peptides were subsequently dissolved in 20 µl 0.1% formic acid by 10×1 s ultrasonication and shaking at 1400 rpm at 25°C for 5 min. Peptide concentrations were determined based on absorbance values using a SPECTROstar Nano Absorbance Plate Reader (BMG Labtech). Peptides were diluted to a concentration of 0.5 µg µl−1 in LC-buffer A. IRT peptides (Biognosys AG, Schlieren, Switzerland) were added to control for LC-MS performance, and samples were stored at −20°C prior to LC-MS/MS analysis using an Orbitrap Fusion Lumos Tribrid Mass Spectrometer fitted with an EASY-nLC 1200 (both Thermo Fisher Scientific) and a custom-made column heater set to 60°C. Peptides were resolved using an RP-HPLC column (75 µm×36 cm) packed in-house with C18 resin (ReproSil-Pur C18-AQ, 1.9 µm resin; Dr Maisch GmbH) at a flow rate of 0.2 µl min−1. The following gradient was used for peptide separation: from 5% B to 12% B over 10 min to 35% B over 80 min to 50% B over 30 min to 95% B over 2 min followed by 18 min at 95% B. Buffer A was 0.1% formic acid in water, and buffer B was 80% acetonitrile, 0.1% formic acid in water. The mass spectrometer was operated in data-dependent acquisition (DDA) mode with a cycle time of 3 s between master scans. Each master scan was acquired in the Orbitrap at a resolution of 120.000 full width at half maximum (at 200m/z, MS1) and a scan range from 375 to 1.600m/z followed by MS/MS (MS2) scans of the most intense precursors in the linear ion trap at ‘Rapid’ scan rate with isolation of the quadrupole set to 1.4m/z. Maximum ion injection time was set to 50 ms (MS1) and 35 ms (MS2) with an AGC target of 1.0E6 and 1.0E4, respectively. Monoisotopic precursor selection (MIPS) was set to peptide, and the intensity threshold was set to 5.0E3. Peptides were fragmented by HCD (higher-energy collisional dissociation) with collision energy set to 35%, and one microscan was acquired for each spectrum. The dynamic exclusion duration was set to 30 s.

Protein identification and data analysis

Raw MS/MS spectra were converted to mzML format using MSConvert (Adusumilli and Mallick, 2017) and subsequently submitted to a closed search with default parameters in the MSFragger (Kong et al., 2017) v.3.1.1 pipeline as implemented in FragPipe v.14.0 (https://fragpipe.nesvilab.org). Spectra were matched against a deduced C. alpina transcriptome with the addition of common laboratory contaminants. The transcriptome de novo assembly pipeline was identical to the one used for other planarian transcriptomes available from the PlanMine database (Rozanski et al., 2019). From the de novo assembled transcriptome, we extracted long open reading frames (ORFs) using the TransDecoder v.5.3.0 pipeline TransDecoder.LongOrfs followed by TransDecoder.Predict. We specifically identified ORFs with homology to known proteins in the Pfam database (El-Gebali et al., 2019) with HMMER (Durbin et al., 1998) v.3.2.1 and subsequently used TransDecoder.Predict with the --retain_pfam_hits option specified. Redundancy of sequences in the deduced proteome was reduced using CD-hit (Li et al., 2001) v.4.8.1 with an identity cut-off of 0.9 and a minimum word size of 5. In MSFragger, enzyme specificity was set as fully tryptic, with a maximum of two missed cleavages. The peptide spectrum-match false discovery rate (FDR) and the protein FDR were both set to 0.01 (based on the target-decoy approach) using the Philosopher toolkit v.3.3.12 (da Veiga Leprevost et al., 2020). Precursor mass tolerance was set to 50 ppm and fragment mass tolerance was set to 20 ppm, with mass calibration and parameter optimization enabled. Label-free quantification (LFQ) was performed using IonQuant (Yu et al., 2021) and the match between runs option (MaxLFQ algorithm; minimum ions: 2). Oxidation of methionine (M) and acetylation (Protein N-term) were specified as variable and carbamidomethylation of cysteines (C) as fixed modifications. Minimum peptide length was set to 7 amino acids with two allowed missed tryptic cleavages. From the MSFragger protein-level output ‘reprint.int.’, the razor intensity columns (‘unique+razor’) were used for downstream analyses in R v.4.0.2 (http://www.R-project.org/). LFQ intensities had ∼19.95% missing values which were imputed using DreamAI, an ensemble machine learning algorithm specifically designed for LFQ data (Ma et al., 2020 preprint). Imputed relative abundances were then subjected to variance stabilization normalization (VSN) using the function normalizeVSN in the Linear Models for Microarray Data (LIMMA) package v.3.46.0 (Smyth, 2005). This method has been shown to outperform other normalization strategies in label-free proteomics data (Välikangas et al., 2018). Unsupervised non-metric multidimensional scaling (nMDS) was performed to visualize proteome differences following acclimation. For this, a Bray–Curtis distance matrix was computed based on relative abundance for each protein and used as input to the metaMDS function in vegan v.2.5.7 (https://CRAN.R-project.org/package=vegan). To identify differentially abundant proteins (DAPs), we computed pairwise differences using the LIMMA package. A linear model was fitted for each protein via function lmFit and contrasts from the model fit and summary statistics were computed via the functions contrasts.fit and eBayes. A list of significant proteins was generated using the function topTable, and P-values were adjusted for multiple testing using the Benjamini–Hochberg (BH) procedure. After testing for normality and homogeneity of variance, one-way analyses of variance (ANOVA) followed by post hoc Tukey Tests were performed on selected proteins. Suites of co-abundant proteins were identified using two methods. (1) For a soft clustering approach, we averaged protein LFQs over the 5 biological replicates per treatment, standardized the data (mean=0, s.d.=1) and subjected them to the fuzzy c-means algorithm as implemented in the Mfuzz (Kumar and Futschik, 2007) package v.2.50.0 with a fuzzification parameter of 2.54 and 8 centres (Fig. S2B,H). (2) Weighted Correlation Network Analysis (WGCNA) v.1.69 was used to identify networks of functionally co-regulated protein groups (Langfelder and Horvath, 2008). First, a sample network was constructed to identify outlying samples with a standardized connectivity score of less than −2.5 (Horvath, 2011). A signed protein co-abundance network (minModuleSize=30) was constructed with a soft threshold power (β) of 12 as found appropriate by the function pickSoftThreshold to reach a scale-free topology index of at least 0.90. We used the Dynamic Tree Cut approach to merge highly correlated modules using a height-cut of 0.25 (Zhang and Horvath, 2005). Module eigengenes (the first principal component of the abundances of all proteins in a module across replicates) were correlated with acclimation temperature as a continuous variable (Fig. S2C–G).

Proteome annotation

All MS/MS-identified (n=3399) protein sequences were queried against the UniProtKB/SwissProt database using stand-alone blastp v.2.11.0 (max_target_seqs=10, minimum e-value=0.000001). Matches with the lowest e-value were extracted from the results and sequences with no hits to the SwissProt database (n=282) were queried against the TrEMBL database with parameters as above. Remaining, unmatched sequences (n=122) were queried against the NCBI non-redundant database. When a sequence had no hits to any of the three databases (n=86), we performed structure homology modelling using SWISS-MODEL (Waterhouse et al., 2018). Protein domains were determined for all sequences using stand-alone InterProScan v.5.39.77.0 with default parameters (Jones et al., 2014). All sequences were assigned to Clusters of Orthologous Groups (COGs) (Tatusov et al., 2000), Kyoto Encyclopedia of Genes and Genomes orthology terms (KO) (Kanehisa and Goto, 2000) and Gene Ontology (GO) terms in eggNOG (Huerta-Cepas et al., 2019) v.5.0 via eggNOG-Mapper v.2 (Taxonomic scope: automatic, Orthologs: all orthologs, GO evidence: non-electronic terms, E-value: 0.001, minimum hit bit-score: 60). We used SignalP (Almagro Armenteros et al., 2019) v.5.0 and TMHMM (Krogh et al., 2001) v.2.0 to identify excreted proteins and transmembrane proteins. Finally, all annotation results were merged into a consensus annotated proteome (available as SI2 in Dryad: doi:10.5061/dryad.dfn2z3541). To test whether any GO terms were over-represented in mFuzz clusters and WGCNA modules, we sorted proteins by their membership and gene significance values, and performed ranked-based tests for each assigned GO term by applying Kolmogorov–Smirnov tests via package topGO (https://bioconductor.org/packages/topGO/) v.2.42.0, with the C. alpina proteome as the GO background universe (method=‘weight01’). We then used the clusterProfiler (Yu et al., 2012) v.3.12 to perform KO enrichment analyses (proteins with membership value >0.5; Benjamini–Hochberg adjusted P<0.05) and visualized results using package enrichplot (Yu, 2020) v.11.2, again using the whole C. alpina proteome as the KO background universe. To corroborate and expand upon GO enrichment results, we performed domain-centric GO enrichment of clusters and modules using the dcGO (Fang and Gough, 2013) suite with a stringent FDR of <0.005 and the IPS-predicted Pfam terms as input. For semantic summarization, we visualized dcGO enrichment results of clusters and modules through treemaps using REVIGO (Supek et al., 2011) with the SimRel semantic similarity measure and a medium allowed similarity of 0.7. As functional enrichment analyses [i.e. (dc)GO and KO enrichment] are inherently biased (Timmons et al., 2015; Tomczak et al., 2018), we additionally scanned the literature individually for each DAP to deduce their functional roles.

CTmax was 1.46°C higher in the alpine C. alpina population (Fig. 1A), which showed reduced CTmax plasticity following thermal acclimation (Fig. 1B). Differences in organism size may have had an influence (Dallas and Rivers-Moore, 2012), but no significant effect was found (Table S2). The mean (±s.d.) CTmax of all individuals, independent of source population, was 27.58±1.34°C (n=96), which is within the range of other planarian species and outweighs the species' theorized thermal niche maximum of 15°C (Claussen and Walters, 1982; Everatt et al., 2014). LC-MS/MS of the alpine population following acclimation resulted in an average of 120.507±3264 acquired spectra per sample, mapping to 12.677±1063 in silico digested peptides. All samples collectively covered 10.74% (3399 out of 31,653 predicted ORFs) of the transcriptome de novo assembly. Protein abundance profiles separated between acclimation groups in the nMDS space, whereby the 8°C centroid was furthest from that for 17°C, followed by 14°C and 11°C, indicating unique proteome states following acclimation (Fig. 1C). However, the separation between treatments was marginal, indicating overall proteome stability within the studied temperature range. C-means fuzzy clustering identified eight unique clusters with varying abundance patterns, ranging from linearly decreasing to stepwise increasing (Fig. S2H). WGCNA assigned all proteins to 15 co-abundance modules and six module eigengenes significantly (P<0.05) correlated with acclimation temperature. Both the ‘green’ (r=0.77) and ‘blue’ (r=0.8) modules correlated negatively, and modules ‘tan’ (r=0.51), ‘purple’ (r=0.48) and ‘magenta’ (r=0.45) correlated positively with acclimation temperature (Fig. 2B,C). These protein sets were enriched for multiple GO and KO terms (Fig. 2) and contained numerous proteins higher and lower in abundance mediating various cellular and physiological processes that changed in response to prolonged thermal acclimation and which are described in detail below.

Fig. 1.

Critical thermal maximum (CTmax) and proteomics results. (A) Violin plots of CTmax values of three Crenobia alpina populations (n=31 per population; see Fig. S1 for information on sites) independent of acclimation temperature. P­-values are shown above the brackets (ANOVA). (B) Raincloud plot of CTmax values following thermal acclimation of the three C. alpina populations shown in A. (C) Non-metric multidimensional scaling (nMDS) ordination based on the relative abundance of 3399 proteins commonly identified across n=5 biological replicates per acclimation temperature in the alpine population (Alps) following 168 h acclimation to 8, 11, 14 or 17°C. (D) Heatmap of all heat shock proteins (HSPs) quantified in the alpine population's proteome following 168 h thermal acclimation to 8, 11, 14 or 17°C. Colours represent the log2 label-free quantification (LFQ) values for each HSP measured in the 20 biological replicates of the alpine C. alpina population (columns; n=5 replicates per acclimation temperature).

Fig. 1.

Critical thermal maximum (CTmax) and proteomics results. (A) Violin plots of CTmax values of three Crenobia alpina populations (n=31 per population; see Fig. S1 for information on sites) independent of acclimation temperature. P­-values are shown above the brackets (ANOVA). (B) Raincloud plot of CTmax values following thermal acclimation of the three C. alpina populations shown in A. (C) Non-metric multidimensional scaling (nMDS) ordination based on the relative abundance of 3399 proteins commonly identified across n=5 biological replicates per acclimation temperature in the alpine population (Alps) following 168 h acclimation to 8, 11, 14 or 17°C. (D) Heatmap of all heat shock proteins (HSPs) quantified in the alpine population's proteome following 168 h thermal acclimation to 8, 11, 14 or 17°C. Colours represent the log2 label-free quantification (LFQ) values for each HSP measured in the 20 biological replicates of the alpine C. alpina population (columns; n=5 replicates per acclimation temperature).

Fig. 2.

Proteome changes of C. alpina following 168 h acclimation to sublethal temperatures. (A) GO enrichment results of proteins of higher abundance following acclimation (WGCNA modules: Tan, Purple and Magenta) for the Cellular Compartment (CC; top) and Biological Process (BP; bottom) GO category. (B) GO enrichment results of proteins of lower abundance following acclimation (WGCNA modules: Green and Blue) for the Molecular Function (MF; top), Cellular Compartment (CC, middle) and Biological Process (BP, bottom) GO category. Dashed lines indicate an adjusted P-value cut-off of 0.05. Colours represent the adjusted P-value and dot size indicates the number of significant proteins associated with the respective GO term. (C) Names (top) and statistics of WGCNA modules and their association with acclimation temperature, represented by Pearson's correlation coefficients (r) between module eigengene (ME) and acclimation temperature, the P-value of the correlation in parentheses and the number of proteins assigned to each module (n; with membership value >0.5). The plot in the middle shows the correlation between the ME expression of the Blue module and acclimation temperature.

Fig. 2.

Proteome changes of C. alpina following 168 h acclimation to sublethal temperatures. (A) GO enrichment results of proteins of higher abundance following acclimation (WGCNA modules: Tan, Purple and Magenta) for the Cellular Compartment (CC; top) and Biological Process (BP; bottom) GO category. (B) GO enrichment results of proteins of lower abundance following acclimation (WGCNA modules: Green and Blue) for the Molecular Function (MF; top), Cellular Compartment (CC, middle) and Biological Process (BP, bottom) GO category. Dashed lines indicate an adjusted P-value cut-off of 0.05. Colours represent the adjusted P-value and dot size indicates the number of significant proteins associated with the respective GO term. (C) Names (top) and statistics of WGCNA modules and their association with acclimation temperature, represented by Pearson's correlation coefficients (r) between module eigengene (ME) and acclimation temperature, the P-value of the correlation in parentheses and the number of proteins assigned to each module (n; with membership value >0.5). The plot in the middle shows the correlation between the ME expression of the Blue module and acclimation temperature.

Decreased protein abundance

Oxidative stress

Acclimation to higher temperatures resulted in a decreased abundance of proteins regulating cellular responses to oxidative stress. The blue and green WGCNA modules were enriched for GO terms ‘cellular response to oxidative stress’ (GO:0034599), ‘oxidation-reduction process’ (GO:0055114), ‘regulation of reactive oxygen species metabolic process’ (GO:2000377) and ‘peroxisome’ (GO:0005777). In mFuzz cluster 3, ‘aldehyde dehydrogenase NAD+’ (K00128) was an enriched KO term, and ‘reactive oxygen species biosynthetic process’ (GO:1903409) and ‘superoxide metabolic process’ (GO:0006801) were enriched GO terms. Proteins associated with these functions were cat, prdx6, bco2, sod1, sod2, aldh1b1 and aldh3a2.

Protein synthesis

Proteins regulating translational control and ribosome biogenesis showed a lower abundance following warm acclimation. Enriched KO terms were ‘large subunit ribosomal protein L6e’ (K02934) and ‘endoplasmic reticulum chaperone BiP’ (K09490). Enriched GO terms included ‘translation’ (GO:0006412), ‘translational initiation’ (GO:0006413), ‘translational elongation’ (GO:0006414), ‘ribosome assembly’ (GO:0042255), ‘ribosomal small subunit biogenesis’ (GO:0042274), ‘endoplasmic reticulum (ER) stress-induced pre-emptive quality control’ (GO:0061857) and ‘protein localization to ER’ (GO:0070972). Proteins related to these terms were multiple 40S and 60S ribosomal subunits (rps3-a, rps8, rps13, rps14, rps23, rpl3, fpl4, rpl6, rpl21, rpl14, rpl18), 60S acidic ribosomal proteins (fplp0, rpp1b, rpp-2), the ubiquitin-60S ribosomal protein L40 (rpl40) and multiple eukaryotic translation initiation factors (eif3d, eif3m, eif3i, infB, eif1ax, eif-5A, eif4h, elf-4g).

Cellular transport

Proteins related to cellular transport were also over-represented in those with decreased abundance following warm acclimation. ‘Transporter activity’ (GO:0005215), ‘endosomal transport’ (GO:0016197), ‘ER to golgi transport vesicle membrane’ (GO:0012507) and ‘golgi-associated vesicle membrane’ (GO:0030660) were enriched terms in the blue and green WGCNA modules (Fig. 2B). Additionally, ‘small guanosine triphosphatase’ (GTPase) was the most frequent Pfam term in mFuzz cluster 3 and both the blue and green modules with associated enriched KO terms ‘RHOC’ (K07857) and ‘RHOA’ (K04513). These transport proteins facilitate endosome to Golgi transport (snx2, snx12), protein transport into the nucleus (ntf2), transport across the ER (sec24b, sec61b, tram1l1) and the retention of ER resident proteins (ssr1, ssr2).

Muscle contraction

Proteins regulating muscle contraction and constitution had a lower abundance following warm acclimation. GO terms ‘muscle system process’ (GO:0003012), ‘muscle attachment’ (GO:0016203), ‘muscle cell cellular homeostasis’ (GO:0046716), ‘striated muscle tissue development’ (GO:0014706) and ‘regulation of muscle hypertrophy’ (GO:0014743) were enriched in warm-responsive clusters and modules. Additionally, transgelin (K20526) was an enriched KO term in cluster 3, represented by cnn1, myophilin (N/A) and tagln3. Lower abundance proteins regulating muscle composition and contraction were unc-89, tni-4, tnt, tropomyosin A, two copies of d-titin (sls), ankyrins (ank2, ank3) and multiple collagens (col4a1, col27a1b, col1a2).

Regeneration

Two proteins with high sequence similarity to Schmidtea mediterranea smedwi-1 and smedwi-2 (72.3% and 75.7% respectively) were of lower abundance following warm acclimation. These proteins belong to the PIWI/Argonaute family and participate in the synthesis and function of small RNAs called PIWI-associated RNAs (piRNAs) which regulate neoblast function and, consequently, regeneration (Reddien et al., 2005a,b; Palakodeti et al., 2008). In addition, the heat shock protein (HSP) 40 family member dnaja1 was also of lower abundance, a chaperone required for the stability of smedwi-1 and smedwi-2 (Wang et al., 2019).

Mitochondrial dynamics

The GO terms ‘ATP-dependent activity’ (GO:0140657), ‘mitochondrion organization’ (GO:0007005) and ‘regulation of mitochondrial membrane potential’ (GO:0051881) were enriched in the green and blue modules. Proteins associated with these GO terms were components of vacuolar ATPase (vha26, vha14, vha68-2, atp6v1c1a, vhaac39-1, vha55, nkb-3), ATP synthase (atpsyncf6, atp5 mg, atpsynd) and ATPase (serca, nkb-1, nkb-3). Additionally, a mitochondrial 10 kDa HSP (cisd3) and a MICOS complex subunit (hnaj_locus369) were lower in abundance following warm acclimation.

Increased protein abundance

Wound healing

Proteins involved in haemostasis, the first stage of wound healing, were of higher abundance following warm acclimation as indicated by the enriched GO terms ‘platelet activation’ (GO:0030168), ‘regulation of response to wounding’ (GO:1903034) and ‘positive regulation of wound healing’ (GO:0090303). Proteins mediating these processes were pdcd10, septin5, vwf, hspg2 and gp5.

Cilia-related proteins

Cilia-related proteins had higher abundance following warm acclimation as indicated by enriched of the GO term ‘9+2 motile cilium’ (GO:0097729) in the CC category of warm-responsive WGCNA modules (Fig. 2A). The proteins dctn1, pdcl, ttc26, ift80, mks1, cfap70 and ttc21b were of higher abundance following warm acclimation, and so were multiple dynein proteins such as a dynein beta chain (ciliary N/A), dynlrb2, dnah2 and two copies of dnah5.

Developmental growth

The GO terms ‘anatomical structure homeostasis’ (GO:0060249), ‘positive regulation of developmental growth’ (GO:0048639) and ‘developmental growth involved in morphogenesis’ (GO:0060560) were enriched in the warm-responsive WGCNA modules (Fig. 2A). Higher-abundance proteins were involved in three associated processes: nervous system development (otk, alpha-Spec, manf, emb-9, prss12, futsch, hil), cell division and fate determination (myh9, cdtn1, ist1, snx33, par-5, flil, cdc73, def6), and photoreceptor function/eye development (rab18, neur, slc2a1).

Immune system

Immune system processes were over-represented in warm-responsive clusters and modules; for example, GO terms ‘immune system process’ (GO:0002376), ‘immune response’ (GO:0006955) and ‘myeloid leukocyte mediated immunity’ (GO:0002444) in the magenta module. Associated proteins mx2, dock2, mrc1, sftpd, ifih1, ikbke, atg101, dpp9, rbck1 and mpeg1 showed increased abundance compared with the control temperature proteome state.

Muscle contraction

In contrast to proteins that positively regulate muscle contraction, proteins that negatively regulate contraction and differentiation such as unc-22, unc-89, crvp, csrp2 and a ryanodine receptor (ryr) exhibited increased abundance following warm acclimation. Additionally, myoglobin was of higher abundance, a protein that facilitates the movement of oxygen within muscles (Wittenberg, 1970).

Mitochondrial dynamics

While proteins associated with ATPases and ATP synthases were lower in abundance following warm acclimation, higher-abundance proteins were enriched for mitochondria (GO:0005739; Fig. 2A). These include proteins forming complexes of the electron transport chain: subunits of NADH dehydrogenase [dea37_0010400 (beta subcomplex subunit 6), d2030.4 (beta subcomplex subunit 7), NDUFA13 (alpha subcomplex subunit 13), ND-42 (alpha subcomplex subunit 10)], subunits of succinate dehydrogenase [sdhb (iron-sulfur subunit), sdhc (b560 subunit)], and a subunit of cytochrome bc1 (cyc1). Additionally, two voltage-dependent anion-selective channel 2 proteins (cdac2), and mitochondrial transmembrane transporter proteins (2× sh3glb1, tim14) were of higher abundance following warm acclimation.

CTmax and HSPs

The ∼28°C CTmax of C. alpina indicates short-term tolerance to warming temperatures and the three populations exhibited both CTmax and CTmax plasticity differences, whereby the alpine population exhibited the highest and least plastic CTmax. While these results suggest acute higher thermal tolerance for this species, CTmax measurements are not necessarily good predictors of long-term acclimation capacity, and depend on both the ramping rate and the starting temperature of the assay (Terblanche et al., 2007; 2011). Following long-term acclimation, individuals of the alpine C. alpina initiated an unfolded protein response (UPR), characterized by a higher abundance of HSP90 co-chaperones cdc37 and unc45a, ube2j1, nap1l4, dnaj homolog dnajc3, three disulfide-isomerases (2×pdi-2, p4hb), sdf2, ppib and tmbim6. Interestingly, all identified HSPs were of comparable abundance following acclimation (Fig. 1D), a pattern that may be due to increasing levels of cold denaturation through the hydrophobic effect (Tsai et al., 2002). Alternatively, while HSPs are initially induced upon a temperature increase, their expression reduces if high temperatures persist, partly because chronic induction can lead to detrimental effects in cells (Gasch et al., 2000; Wang et al., 2006; Roth et al., 2014). The initiation of an UPR and a high number of identified HSPs suggests that C. alpina did not lose mechanisms to assist refolding or removing damaged proteins should temperatures increase, separating it from certain cold-adapted species from the frigid (e.g. Hofmann et al., 2000; Clark et al., 2008) or temperate zones (e.g. Bosch et al., 1988).

Mitochondrial respiration and oxidative stress

Adjustments in mitochondrial properties and capacity are essential features of temperature acclimation in a variety of species. In C. alpina, warm acclimation resulted in increased abundance of electron transport chain (ETC) complex subunits, but in lowered abundance of ATP synthase and ATPase subunits. These results suggest an increase in mitochondrial respiration at warmer temperatures, while simultaneously indicating decreased levels of ADP phosphorylation and ATP dephosphorylation. Alternatively, lowered amounts of ATP synthases and ATPases may reflect an adaptive mechanism that limits ATP use for non-essential pathways, or indicate that the activity of these proteins is increased at higher temperatures such that lower numbers are needed to compensate for reduced enzyme activity at lower temperatures. Additionally, elevated mitochondrial respiration commonly leads to an increased production of reactive oxygen species (ROS; Abele et al., 2002). However, for C. alpina, we observed a lowered abundance of proteins mitigating oxidative stress following warm acclimation. Previous studies identified an increase in oxidative stress proteins during cold acclimation (Mádi et al., 2003; Ibarz et al., 2010), in particular of superoxide dismutases sod1 and sod2 (Malek et al., 2004). Taken together, these findings suggest complex responses at the proteome level regarding mitochondrial responses and that oxidative stress for C. alpina may be higher at low compared with high temperatures.

Cellular transport

Exposure to cold leads to a depression in the rate of virtually all biological processes, including transport mechanisms which regulate cellular functioning (Kuismanen and Saraste, 1989; Malan and Canguilhem, 1989). Multiple Rab homologues, sorting nexins and nuclear transport factors were of lower abundance following thermal acclimation, proteins which function as regulators of transportation networks (Novick and Zerial, 1997; Cullen, 2008). Proteins from the Rab family, for example, regulate intracellular vesicular transport and trafficking of proteins between organelles (Zerial and McBride, 2001; Etienne-Manneville and Hall, 2002). An extensive study on the effects of prolonged thermal acclimation on the proteome of Saccharomyces cerevisiae showed that protein re-localization is a core cellular response (Domnauer et al., 2021) and transport proteins have commonly been found to be up-regulated following warm acclimation (Palumbi et al., 2014; Veilleux et al., 2015). As in the case of a lower abundance of proteins functioning in protein synthesis (see below), it is difficult to discern whether this response constitutes a disruption of cellular functioning or whether lesser amounts of these proteins are needed as a result of the absence of cold-induced repression of transportation. The latter scenario leads us to hypothesize that higher temperatures may reduce the need to compensate cellular transport functioning in C. alpina at cold temperatures, in turn leading to reduced demands on this core cellular process.

Protein synthesis

In C. alpina, a lower abundance of 15 identified ribosomal subunits and 8 identified translation initiation factors following warm acclimation suggests reduced protein synthesis at warmer temperatures. As repression of protein synthesis is a feature of cells exposed to elevated temperatures (Duncan and Hershey, 1989), attributed to inhibition of translation initiation (Shalgi et al., 2013), this down-regulation may ensure the conservation of resources that are needed to survive adverse conditions (Sheikh and Fornace, 1999; Larade and Storey, 2007). However, lower temperatures slow down RNA translational efficiency, and reduce in vivo protein synthesis (Friedman et al., 1971; Broeze et al., 1978; Fraser et al., 2002b). A lowered abundance of these proteins at higher temperatures may therefore reflect a reduced need to compensate for impaired protein synthesis at colder temperatures. These findings lead to an interesting question: in species from cold-stable environments, does a lowered abundance of proteins functioning in the protein synthesis machinery represent a ‘positive’ (i.e. reduced demand at higher temperatures) or ‘negative’ (i.e. impaired protein synthesis) response to elevated temperatures?

Reduced DNA repair and demands on molecule structural flexibility

Mechanisms such as alternative splicing and control of nucleic acid topology allow organisms to alter molecule thermostability (López-García and Forterre, 2000; D'Amico et al., 2002). At colder temperatures, RNAs fold into thermodynamically stable secondary structures which impede transcription and translation. In C. alpina, three Y-box homolog proteins containing a cold-shock domain (CSD) were of lower abundance following warm acclimation (Fig. S3). The accuracy and regulation of alternative splicing events is also temperature dependent (Hopkins et al., 2018; Healy and Schulte, 2019), and multiple splicing-related proteins were of lower abundance at higher acclimation temperatures. These include srsf1, a protein that prevents exon skipping (Kohtz et al., 1994), nova1, rbm22, khsrp, dhx8, two copies of khdrbs2, ddx21 and rbm22. Additionally, a UV excision repair protein (rad23b), cbx3, an endonuclease mediating miRNA decay (snd1), serpinb10 and two 5'–3′ exonucleases (pld3, xrn1) were lower in abundance following warm acclimation, indicating a reduced need to mitigate cold-induced DNA damage and topology restrictions at higher temperatures. Based on these findings, we hypothesize that C. alpina maintains cold-inducible mechanisms to relax DNA topology and to ensure accurate transcript splicing, and that these processes are down-regulated at elevated temperatures.

Proteome homeostasis

If temperature-induced damaged or misfolded proteins accumulate in the cell, they may become cytotoxic and interfere with cellular function (Sherman and Goldberg, 2001). If proteins are beyond repair, they are degraded by proteases, primarily by the ubiquitin–proteasome pathway (Wickner et al., 1999). In C. alpina, ubiquitin-related proteins (ubq11, rpL40, ubcD1, ube2l3, uba1 and ube2z) and 26S proteasome-related proteins (psmd1, psmd2, psmc2, psmb2, psmb6, psmb7, psmf1, rpt6 and rpn11) were of lower abundance following warm acclimation. These findings indicate that changes in the ubiquitin–proteasome pathway and the degradation of ubiquitinated proteins play key roles in the maintenance of protein homeostasis at colder temperatures, but that, at higher temperatures, these processes are down-regulated in C. alpina. Interestingly, compared with cold-adapted Mytilus galloprovincialis, in warm-adapted M. galloprovincialis, proteasome isoforms were also shown to be down-regulated (Tomanek, 2012), potentially prolonging the lifespan of proteins implicated in coping with elevated temperatures, as has been shown for oxidative stress proteins (Bieler et al., 2009).

Muscle contraction

The maintenance of mechanical muscle performance at low temperatures can be found in a variety of species (Bennett, 1985; Goldspink, 1995; Gillis and Tibbits, 2002; Tiitu and Vornanen, 2002; Coughlin et al., 2016) and locomotion speed commonly changes in an acclimation temperature-dependent manner (Watabe, 2002). A variety of muscle-related proteins were of lower abundance following warm acclimation of C. alpina. These include troponins (tnt, tni-4), which confer calcium sensitivity to striated muscle (Gomes et al., 2002), tropomyosins (tm1, smtmi and egtrpa), which additionally regulate calcium-dependent muscle contraction (Galińska-Rakoczy et al., 2008), calponins (box15_mlig001580g2, and cnn1), which modulate muscle contraction (Winder et al., 1998), and ankyrins (ank2 and ank3), which regulate muscle gene expression (Kojic et al., 2010; Laure et al., 2010). Conversely, muscle-related proteins such as unc-22, unc-45a, unc-89 and a cysteine- and glycine-rich protein 2 (csrp2) were of higher abundance following warm acclimation. These proteins facilitate sarcomere organization and extensibility, and thereby regulation of muscle composition, contraction and relaxation (Tskhovrebova and Trinick, 2003; Nyegaard et al., 2012; Koh et al., 2013). Locomotion speed of planarians is reduced at lower temperatures, and their muscle-powered movements are susceptible to temperature changes (Inoue et al., 2014). Our findings pinpoint to the proteins that are probably regulating muscle contraction and sarcomere organization during muscle adaptation to higher and lower temperatures in planarians.

Neurotransmitter regulation and wound healing

Both dopamine and γ-aminobutyric acid (GABA) modulate temperature sensitivity (Fraser et al., 2002a; Ueno et al., 2012; Jakšić et al., 2020), and the higher abundance of proteins gch1 (Tegeder et al., 2006), manf (Palgi et al., 2009), aldh5a1 (Kim et al., 2009) and ddc (Rorsman et al., 1995) suggests regulation of the levels of these neurotransmitters following warm acclimation. Interestingly, planarian locomotion such as phototaxis is coordinated by dopaminergic and GABAergic neurons (Nishimura et al., 2007; 2008), suggesting a link between the abundance of the above-mentioned proteins and a response to warm acclimation. A higher abundance of vwf, which promotes the adhesion of platelets (Kroll et al., 1991), and gp5, which mediates vwf-dependent platelet adhesion to blood vessels (Hickey et al., 1993), suggests a possible regulation of haemostasis following warm acclimation.

Maintenance of ciliated structures

The ventral epidermis of planarians is composed of multiciliated cells (MCCs), allowing them to glide along substrates. Specifically, cilia that populate MCCs are highly conserved, with a 9+2 axoneme and a full complement of inner and outer arm dynein motors (Rompolas et al., 2013; Azimzadeh and Basquin, 2016). Here, we found that a variety of proteins that maintain ciliated structures increased in abundance following warm acclimation. After exposure to elevated temperatures, ciliated structures were absent from the dorsal epidermis of the planarian Girardia tigrina (De Oliveira et al., 2018), and individuals of the planarian Dugesia dorotocephala exhibit reduced movement speed at higher temperatures (Claussen et al., 2003). Accordingly, we hypothesize that cilia-related proteins may increase in C. alpina following warm acclimation either to compensate for temperature-induced damage to MCCs or to facilitate increased movement speed at higher temperatures.

Immune system processes

Innate immune system-related proteins showed both increased and decreased abundance following warm acclimation. Two general trends could be discerned: proteins functioning in leukotriene biosynthesis (two copies of cotl1, and lta4 h) showed reduced abundance, and proteins related to autophagy, viral infection and macrophage-like cell activity such as atg101 (Mercer et al., 2009), ikbke (Hemmi et al., 2004), ifih1 (Sarmiento et al., 2017), mx2 (Kane et al., 2013), mrc1 (Harris et al., 1992) and sftpd (Holmskov et al., 1999) showed increased abundance. Additionally, mrc1, a C-type lectin that recognizes different types of pathogens such as bacteria and fungi (Gao et al., 2017; Brown et al., 2018) was of higher abundance following warm acclimation. Together, these data indicate regulation of the innate immune system to respond to changes in the pathogenic microbial community composition following a thermal regime change. The spring water used in the acclimation experiment was filtered but microbial communities were not monitored throughout the experiment, so it cannot be ruled out that pathogens were present in higher acclimation treatments. However, temperature increase has previously been shown to increase the capacity of planarians to cope with pathogens (Hammoudi et al., 2018; Kangale et al., 2021) and cold-stable habitats such as freshwater springs harbour a unique microbial community that has only recently begun to be characterized (Savio et al., 2018; Moldovan et al., 2020). The identified changes in immune system proteins hint at unknown interactions between freshwater planarian immunity, pathogen presence and environmental temperature, providing hypotheses for future studies. For example, planarians may reduce leukotriene synthesis at warmer temperatures and increase the de novo synthesis of pathogen receptors and the number of macrophage-like cells. However, whether the regulation of immune system processes is related to mitigation of temperature-induced damage or is an adaptive feature of environmentally dictated pathogen resistance mechanisms cannot be gleaned from our comparative analysis alone.

Study limitations

With regards to limitations of the proteomics approach, the de novo transcriptome may not contain specific transcripts that play a role in thermal acclimation as the RNA-seq data did not originate from the same individuals used for proteome analysis. Furthermore, proteins could serve purposes that have not yet been functionally annotated (e.g. moonlighting proteins), which could lead to inaccurate conclusions (Jeffery, 2014). Additionally, starvation may have had antagonistic or synergistic effects on C. alpina's responses to a temperature increase compared with conditions that are not nutrient limiting. The absence of classically observed stress responses following exposure to elevated temperatures may therefore have been due to a lack of energy resulting from starvation. Future experiments should aim at comparing nutrient-limiting versus non-limiting conditions on thermal tolerance. However, the identified proteome differences between acclimation treatments should be independent of starvation-induced mechanisms as all individuals were equally starved. Finally, molecular responses to constant versus varying thermal regimes differ (Salachan and Sørensen, 2022), suggesting that molecular responses of C. alpina under natural conditions of warming may differ from the ones identified in this study.

Implications for the ecology of C. alpina

Based on our findings, we hypothesize that C. alpina is able to tolerate acute and longer-term temperature increases better than previously assumed. Given that the distribution of C. alpina is generally restricted (Brändle et al., 2007), yet it can tolerate the typical temperatures where other planarian species occur, factors other than temperature may dictate the species' distribution (see also Vila-Farré and Rink, 2018). Interestingly, in the central Pyrenees, a single spring did not harbour more than one planarian species (Roca et al., 1992) and in Welsh streams, an exploitation competition between C. alpina, Phagocata vitta and Polycelis felina was observed, whereby co-existence was confined to conditions with high prey abundance (Lock and Reynoldson, 1976; Durance and Ormerod, 2010). Such competitive exclusion is frequently observed in planarians and may be the main reason for C. alpina's restricted distribution. Temperature may therefore only seemingly be a limiting factor for this species, and inter-specific competition following warming of groundwater temperatures may be a stronger factor influencing future population dynamics. Finally, behavioural aspects of thermal tolerance such as phototaxis and thermotaxis, as indicated by the higher abundance of proteins related to photoreceptor development, are expected to play an important role under natural conditions (Dillon et al., 2009; Inoue et al., 2014; Paskin et al., 2014). In this context, we provide anecdotal evidence that individuals of the alpine C. alpina population were preferentially located on the surface of floating plants, arching their bodies towards the sun, suggesting positive phototaxis under cold-stable conditions to regulate body temperature.

Warm tolerance of alpine aquatic species: a contentious issue?

Alpine aquatic species are considered particularly vulnerable to environmental warming because of their inability to tolerate warming temperatures (Füreder, 1999; Bernabò et al., 2011), and because of the combined effects of habitat insularity and upward distributional shifts into increasingly smaller areas of suitable habitat (Jacobsen et al., 2012; Rubidge et al., 2012; Hotaling et al., 2017). However, evidence increasingly suggests that certain high-elevation aquatic organisms may be able to tolerate warming temperatures. For example, the meltwater stonefly Lednia tumana and the cold-stenothermal chironomid Pseudodiamesa branickii exhibit high thermal tolerance, and adaptive gene expression responses to warming temperatures (e.g. expression of stable hsps and hsp70; Bernabò et al., 2011; Treanor et al., 2013; Hotaling et al., 2020), the caddisfly Crunoecia irrorata from cold-stable freshwater springs exhibits plastic proteome responses following thermal acclimation (Ebner et al., 2019), the cold-stenothermal chironomid Diamesa tonsa expresses a heat-inducible pseudo-hsp70 gene encoding a putative long non-coding RNA (lncRNA) (Bernabò et al., 2020), and, at the community level, high-elevation aquatic invertebrate communities can persist despite deglaciation (Muhlfeld et al., 2020). Despite these findings, alpine communities have also been shown to homogenize as a delayed response to increased thermal conditions (Lencioni et al., 2022). Should alpine aquatic biota indeed have a higher physiological thermal tolerance, it is reasonable to assume that factors aside from limited physiological tolerance dictate future population and community processes (see also Shah et al., 2020). While there certainly are cold-stenothermal species that are not able to tolerate warming temperatures (e.g. Schoville et al., 2015; Lencioni et al., 2022), the influence of biotic interactions at the community level such as microbial community changes (Broadbent et al., 2021), competitive exclusion (Węsławski et al., 2021), invasion potential (Holzapfel and Vinebrooke, 2005) and the reduction of available habitat (Fenoglio et al., 2010) may be dictating alpine aquatic biota persistence and restructuring under climate change scenarios.

Conclusion

By studying how prolonged thermal acclimation affects the proteome of C. alpina, we have identified a variety of responses to warming temperatures in this alpine planarian. We found signs of evolved molecular cold tolerance such as CSD-containing proteins and increased abundance of proteins functioning in core cellular processes at lower temperatures (i.e. transcription, translation and cellular transport). However, an acclimation temperature-independent abundance of HSPs, an initiated UPR and a comparatively high CTmax lead us to suggest that C. alpina's acute tolerance to warming temperatures and its acclimation capacity may historically have been underestimated. Our study indicates that this species may be a cold-adapted eurytherm rather than a cold-adapted stenotherm. CTmax and CTmax plasticity differed between populations, further suggesting that thermal history affects the species' resilience to acute temperature stress. While alpine aquatic biodiversity is adversely affected by global climate change, there is mounting evidence from molecular and community-level studies indicating that certain high-elevation, cold-stenothermal and alpine organisms may have the physiological capacity to adequately respond to warming temperatures. We suggest that other factors such as competitive exclusion may better explain their current restricted distributions. In addition, observed immune-system modulation at higher temperatures indicates largely unstudied interactions between thermal regime change, microbial composition and immune capacity of invertebrates, warranting further investigation.

The Crenobia alpina transcriptome data were generously provided by the Rink group at the Department of Tissue Dynamics and Regeneration at the Max Planck Institute for Biophysical Chemistry, Germany. We thank Miquel Vila Farré and colleagues for granting us early access to the unpublished data. We want to thank Lucas Blattner, Colin Courtney-Mustaphi and Gabriel Erni Cassola for comments and corrections on earlier versions of the manuscript. Part of the calculations was performed at sciCORE (http://scicore.unibas.ch/) at the University of Basel. The proteomics data were generated at the Proteomics Core Facility (PCF) at the Biozentrum of the University of Basel.

Author contributions

Conceptualization: J.N.E.; Methodology: J.N.E., M.K.W., D.R.; Software: J.N.E., D.R.; Validation: J.N.E.; Formal analysis: J.N.E.; Investigation: J.N.E.; Data curation: J.N.E., D.R.; Writing - original draft: J.N.E.; Writing - review & editing: J.N.E., D.R., S.v.F.; Visualization: J.N.E.; Supervision: S.v.F.; Project administration: S.v.F.; Funding acquisition: S.v.F.

Funding

This work was supported by the Swiss National Science Foundation (Schweizerischer Nationalfonds zur Förderung der Wissenschaftlichen Forschung grant number 31003A_176234). The funding body had no role in the design of the study and collection, analysis and interpretation of data, and in writing the manuscript. Open Access funding provided by Universität Basel. Deposited in PMC for immediate release.

Data availability

Datasets (including raw, transformed and filtered data) supporting the conclusions of this article are available via the Dryad data repository (Ebner, 2022): doi:10.5061/dryad.dfn2z3541. Files on Dryad include the fully annotated C. alpina proteome (SI2), results from dcGO enrichment analyses (SI3 and SI4), Pfam frequencies of mFuzz clusters and WGCNA modules (SI5), and KO enrichment results (SI6). Mass spectrometry raw data have been deposited to the ProteomeXchange Consortium via the PRIDE (Perez-Riverol et al., 2019) partner repository with the dataset identifier PXD024592 (https://www.ebi.ac.uk/pride/archive/projects/PXD024592).

Abele
,
D.
,
Heise
,
K.
,
Pörtner
,
H. O.
and
Puntarulo
,
S.
(
2002
).
Temperature-dependence of mitochondrial function and production of reactive oxygen species in the intertidal mud clam Mya arenaria
.
J. Exp. Biol.
205
,
1831
-
1841
.
Adusumilli
,
R.
and
Mallick
,
P.
(
2017
).
Data conversion with ProteoWizard msConvert
.
Methods Mol. Biol.
1550
,
339
-
368
.
Almagro Armenteros
,
J. J.
,
Tsirigos
,
K. D.
,
Sønderby
,
C. K.
,
Petersen
,
T. N.
,
Winther
,
O.
,
Brunak
,
S.
,
von Heijne
,
G.
and
Nielsen
,
H.
(
2019
).
SignalP 5.0 improves signal peptide predictions using deep neural networks
.
Nat. Biotechnol.
37
,
420
-
423
.
Atkins
,
K. E.
and
Travis
,
J. M. J.
(
2010
).
Local adaptation and the evolution of species’ ranges under climate change
.
J. Theor. Biol.
266
,
449
-
457
.
Azimzadeh
,
J.
and
Basquin
,
C.
(
2016
).
Basal bodies across eukaryotes series: basal bodies in the freshwater planarian Schmidtea mediterranea
.
Cilia
5
,
15
.
Bennett
,
A. F.
(
1985
).
Temperature and muscle
.
J. Exp. Biol.
115
,
333
-
344
.
Bernabò
,
P.
,
Rebecchi
,
L.
,
Jousson
,
O.
,
Martínez-Guitarte
,
J. L.
and
Lencioni
,
V.
(
2011
).
Thermotolerance and hsp70 heat shock response in the cold-stenothermal chironomid Pseudodiamesa branickii (NE Italy)
.
Cell Stress Chaperones
16
,
403
-
410
.
Bernabò
,
P.
,
Viero
,
G.
and
Lencioni
,
V.
(
2020
).
A long noncoding RNA acts as a post-transcriptional regulator of heat shock protein (HSP70) synthesis in the cold hardy Diamesa tonsa under heat shock
.
PLoS One
15
,
e0227172
.
Bieler
,
S.
,
Meiners
,
S.
,
Stangl
,
V.
,
Pohl
,
T.
and
Stangl
,
K.
(
2009
).
Comprehensive proteomic and transcriptomic analysis reveals early induction of a protective anti-oxidative stress response by low-dose proteasome inhibition
.
Proteomics
9
,
3257
-
3267
.
Bludau
,
I.
and
Aebersold
,
R.
(
2020
).
Proteomic and interactomic insights into the molecular basis of cell functional diversity
.
Nat. Rev. Mol. Cell Biol.
21
,
327
-
340
.
Bosch
,
T. C.
,
Krylow
,
S. M.
,
Bode
,
H. R.
and
Steele
,
R. E.
(
1988
).
Thermotolerance and synthesis of heat shock proteins: these responses are present in Hydra attenuata but absent in Hydra oligactis
.
Proc. Natl. Acad. Sci. USA
85
,
7927
-
7931
.
Brändle
,
M.
,
Heuser
,
R.
,
Marten
,
A.
and
Brandl
,
R.
(
2007
).
Population structure of the freshwater flatworm Crenobia alpina (Dana): old lineages and low gene flow
.
J. Biogeogr.
34
,
1183
-
1192
.
Broadbent
,
A. A. D.
,
Snell
,
H. S. K.
,
Michas
,
A.
,
Pritchard
,
W. J.
,
Newbold
,
L.
,
Cordero
,
I.
,
Goodall
,
T.
,
Schallhart
,
N.
,
Kaufmann
,
R.
,
Griffiths
,
R. I.
et al. 
(
2021
).
Climate change alters temporal dynamics of alpine soil microbial functioning and biogeochemical cycling via earlier snowmelt
.
ISME J.
15
,
2264
-
2275
.
Broeze
,
R. J.
,
Solomon
,
C. J.
,
Pope
,
D. H.
(
1978
).
Effects of low temperature on in vivo and in vitro protein synthesis in Escherichia coli and Pseudomonas fluorescens
.
J. Bacteriol.
134
,
861
-
874
.
Brown
,
G. D.
,
Willment
,
J. A.
and
Whitehead
,
L.
(
2018
).
C-type lectins in immunity and homeostasis
.
Nat. Rev. Immunol.
18
,
374
-
389
.
Cahill
,
A. E.
,
Aiello-Lammens
,
M. E.
,
Fisher-Reid
,
M. C.
,
Hua
,
X.
,
Karanewsky
,
C. J.
,
Yeong Ryu
,
H.
,
Sbeglia
,
G. C.
,
Spagnolo
,
F.
,
Waldron
,
J. B.
,
Warsi
,
O.
et al. 
(
2013
).
How does climate change cause extinction?
Proc. R. Soc. B
280
,
20121890
.
Chevin
,
L.-M.
,
Gallet
,
R.
,
Gomulkiewicz
,
R.
,
Holt
,
R. D.
and
Fellous
,
S.
(
2013
).
Phenotypic plasticity in evolutionary rescue experiments
.
Philos. Trans. R. Soc. B Biol. Sci.
368
,
20120089
.
Clark
,
M. S.
,
Fraser
,
K. P. P.
,
Burns
,
G.
and
Peck
,
L. S.
(
2008
).
The HSP70 heat shock response in the Antarctic fish Harpagifer antarcticus
.
Polar Biol.
31
,
171
-
180
.
Claussen
,
D. L.
and
Walters
,
L. M.
(
1982
).
Thermal acclimation in the fresh water planarians, Dugesia tigrina and D. dorotocephala
.
Hydrobiologia
94
,
231
-
236
.
Claussen
,
D. L.
,
Grisak
,
A. G.
and
Brown
,
P. F.
(
2003
).
The thermal relations of the freshwater triclad flatworm, Dugesia dorotocephala (Turbellaria: Tricladida)
.
J. Therm. Biol.
28
,
457
-
464
.
Coughlin
,
D. J.
,
Shiels
,
L. P.
,
Nuthakki
,
S.
and
Shuman
,
J. L.
(
2016
).
Thermal acclimation to cold alters myosin content and contractile properties of rainbow smelt, Osmerus mordax, red muscle
.
Comp. Biochem. Physiol. A Mol. Integr. Physiol.
196
,
46
-
53
.
Cullen
,
P. J.
(
2008
).
Endosomal sorting and signalling: an emerging role for sorting nexins
.
Nat. Rev. Mol. Cell Biol.
9
,
574
-
582
.
da Veiga Leprevost
,
F.
,
Haynes
,
S. E.
,
Avtonomov
,
D. M.
,
Chang
,
H.-Y.
,
Shanmugam
,
A. K.
,
Mellacheruvu
,
D.
,
Kong
,
A. T.
and
Nesvizhskii
,
A. I.
(
2020
).
Philosopher: a versatile toolkit for shotgun proteomics data analysis
.
Nat. Methods
17
,
869
-
870
.
Dallas
,
H. F.
and
Rivers-Moore
,
N. A.
(
2012
).
Critical Thermal Maxima of aquatic macroinvertebrates: towards identifying bioindicators of thermal alteration
.
Hydrobiologia
679
,
61
-
76
.
D'Amico
,
S.
,
Claverie
,
P.
,
Collins
,
T.
,
Georlette
,
D.
,
Gratia
,
E.
,
Hoyoux
,
A.
,
Meuwis
,
M.-A.
,
Feller
,
G.
and
Gerday
,
C.
(
2002
).
Molecular basis of cold adaptation
.
Philos. Trans. R. Soc. Lond. B Biol. Sci.
357
,
917
-
925
.
De Maayer
,
P.
,
Anderson
,
D.
,
Cary
,
C.
and
Cowan
,
D. A.
(
2014
).
Some like it cold: understanding the survival strategies of psychrophiles
.
EMBO Rep.
15
,
508
-
517
.
de Oliveira
,
M. S.
,
Lopes
,
K. A. R.
,
Leite
,
P. M. S. C. M.
,
Morais
,
F. V.
and
de Campos Velho
,
N. M. R.
(
2018
).
Physiological evaluation of the behavior and epidermis of freshwater planarians (Girardia tigrina and Girardia sp.) exposed to stressors
.
Biol. Open
7
,
bio029595
.
Dillon
,
M. E.
,
Wang
,
G.
,
Garrity
,
P. A.
and
Huey
,
R. B.
(
2009
).
Thermal preference in Drosophila
.
J. Therm. Biol.
34
,
109
-
119
.
Domnauer
,
M.
,
Zheng
,
F.
,
Li
,
L.
,
Zhang
,
Y.
,
Chang
,
C. E.
,
Unruh
,
J. R.
,
Conkright-Fincham
,
J.
,
Mccroskey
,
S.
,
Florens
,
L.
,
Zhang
,
Y.
et al. 
(
2021
).
Proteome plasticity in response to persistent environmental change
.
Mol. Cell
81
,
3294
-
3309.e12
.
Duncan
,
R. F.
and
Hershey
,
J. W.
(
1989
).
Protein synthesis and protein phosphorylation during heat stress, recovery, and adaptation
.
J. Cell Biol.
109
,
1467
-
1481
.
Durance
,
I.
and
Ormerod
,
S. J.
(
2007
).
Climate change effects on upland stream macroinvertebrates over a 25-year period
.
Glob. Change Biol.
13
,
942
-
957
.
Durance
,
I.
and
Ormerod
,
S. J.
(
2010
).
Evidence for the role of climate in the local extinction of a cool-water triclad
.
J. North Am. Benthol. Soc.
29
,
1367
-
1378
.
Durbin
,
R.
,
Eddy
,
S. R.
,
Krogh
,
A.
and
Mitchison
,
G.
(
1998
).
Biological Sequence Analysis: Probabilistic Models of Proteins and Nucleic Acids
.
Cambridge University Press
.
Ebner
,
J. N.
,
Ritz
,
D.
and
Von Fumetti
,
S.
(
2019
).
Comparative proteomics of stenotopic caddisfly Crunoecia irrorata identifies acclimation strategies to warming
.
Mol. Ecol.
28
,
4453
-
4469
.
Ebner
,
J.
(
2022
).
Supplementary data for: Effects of thermal acclimation on the proteome of the planarian Crenobia alpina from an alpine freshwater spring
.
Dryad, Dataset
.
El-Gebali
,
S.
,
Mistry
,
J.
,
Bateman
,
A.
,
Eddy
,
S. R.
,
Luciani
,
A.
,
Potter
,
S. C.
,
Qureshi
,
M.
,
Richardson
,
L. J.
,
Salazar
,
G. A.
,
Smart
,
A.
et al. 
(
2019
).
The Pfam protein families database in 2019
.
Nucleic Acids Res.
47
,
D427
-
D432
.
Elowitz
,
M. B.
,
Levine
,
A. J.
,
Siggia
,
E. D.
and
Swain
,
P. S.
(
2002
).
Stochastic gene expression in a single cell
.
Science
297
,
1183
-
1186
.
Etienne-Manneville
,
S.
and
Hall
,
A.
(
2002
).
Rho GTPases in cell biology
.
Nature
420
,
629
-
635
.
Everatt
,
M. J.
,
Convey
,
P.
,
Worland
,
M. R.
,
Bale
,
J. S.
and
Hayward
,
S. A. L.
(
2014
).
Are the Antarctic dipteran, Eretmoptera murphyi, and Arctic collembolan, Megaphorura arctica, vulnerable to rising temperatures?
Bull. Entomol. Res.
104
,
494
-
503
.
Fang
,
H.
and
Gough
,
J.
(
2013
).
dcGO: database of domain-centric ontologies on functions, phenotypes, diseases and more
.
Nucleic Acids Res.
41
,
D536
-
D544
.
Feder
,
M. E.
and
Walser
,
J.-C.
(
2005
).
The biological limitations of transcriptomics in elucidating stress and stress responses
.
J. Evol. Biol.
18
,
901
-
910
.
Fenoglio
,
S.
,
Bo
,
T.
,
Cucco
,
M.
,
Mercalli
,
L.
and
Malacarne
,
G.
(
2010
).
Effects of global climate change on freshwater biota: a review with special emphasis on the Italian situation
.
Ital. J. Zool.
77
,
374
-
383
.
Franklin
,
C. E.
and
Hoppeler
,
H. H.
(
2021
).
Elucidating mechanism is important in forecasting the impact of a changing world on species survival
.
J. Exp. Biol.
224
,
jeb242284
.
Fraser
,
E. J.
,
Bosma
,
P. T.
,
Trudeau
,
V. L.
and
Docherty
,
K.
(
2002a
).
The effect of water temperature on the GABAergic and reproductive systems in female and male goldfish (Carassius auratus)
.
Gen. Comp. Endocrinol.
125
,
163
-
175
.
Fraser
,
K. P. P.
,
Clarke
,
A.
and
Peck
,
L. S.
(
2002b
).
Low-temperature protein metabolism: seasonal changes in protein synthesis and RNA dynamics in the Antarctic limpet Nacella concinna Strebel 1908
.
J. Exp. Biol.
205
,
3077
-
3086
.
Friedman
,
H.
,
Lu
,
P.
and
Rich
,
A.
(
1971
).
Temperature control of initiation of protein synthesis in Escherichia coli
.
J. Mol. Biol.
61
,
105
-
121
.
Füreder
,
L.
(
1999
).
High alpine streams: cold habitats for insect larvae
. In
Cold-Adapted Organisms: Ecology, Physiology, Enzymology and Molecular Biology
(ed.
R.
Margesin
and
F.
Schinner
), pp.
181
-
196
.
Berlin, Heidelberg
:
Springer
.
Galińska-Rakoczy
,
A.
,
Engel
,
P.
,
Xu
,
C.
,
Jung
,
H. S.
,
Craig
,
R.
,
Tobacman
,
L. S.
and
Lehman
,
W.
(
2008
).
Structural basis for the regulation of muscle contraction by troponin and tropomyosin
.
J. Mol. Biol.
379
,
929
-
935
.
Gao
,
L.
,
Han
,
Y.
,
Deng
,
H.
,
Hu
,
W.
,
Zhen
,
H.
,
Li
,
N.
,
Qin
,
N.
,
Yan
,
M.
,
Wu
,
W.
,
Liu
,
B.
et al. 
(
2017
).
The role of a novel C-type lectin-like protein from planarian in innate immunity and regeneration
.
Dev. Comp. Immunol.
67
,
413
-
426
.
Gasch
,
A. P.
,
Spellman
,
P. T.
,
Kao
,
C. M.
,
Carmel-Harel
,
O.
,
Eisen
,
M. B.
,
Storz
,
G.
,
Botstein
,
D.
and
Silver
,
P. A.
(
2000
).
Genomic expression programs in the response of yeast cells to environmental changes
.
Mol. Biol. Cell
11
,
4241
-
4257
.
Gillis
,
T. E.
and
Tibbits
,
G. F.
(
2002
).
Beating the cold: the functional evolution of troponin C in teleost fish
.
Comp. Biochem. Physiol. A Mol. Integr. Physiol.
132
,
763
-
772
.
Goldspink
,
G.
(
1995
).
Adaptation of fish to different environmental temperature by qualitative and quantitative changes in gene expression
.
J. Therm. Biol.
20
,
167
-
174
.
Gomes
,
A. V.
,
Potter
,
J. D.
and
Szczesna-Cordary
,
D.
(
2002
).
The role of troponins in muscle contraction
.
IUBMB Life
54
,
323
-
333
.
Hailemariam
,
M.
,
Eguez
,
R. V.
,
Singh
,
H.
,
Bekele
,
S.
,
Ameni
,
G.
,
Pieper
,
R.
and
Yu
,
Y.
(
2018
).
S-Trap, an ultrafast sample-preparation approach for shotgun proteomics
.
J. Proteome Res.
17
,
2917
-
2924
.
Hammoudi
,
N.
,
Torre
,
C.
,
Ghigo
,
E.
and
Drancourt
,
M.
(
2018
).
Temperature affects the biology of Schmidtea mediterranea
.
Sci. Rep.
8
,
14934
.
Hampe
,
A.
and
Jump
,
A. S.
(
2011
).
Climate relicts: past, present, future
.
Annu. Rev. Ecol. Evol. Syst.
42
,
313
-
333
.
Harris
,
N.
,
Super
,
M.
,
Rits
,
M.
,
Chang
,
G.
and
Ezekowitz
,
R. A.
(
1992
).
Characterization of the murine macrophage mannose receptor: demonstration that the downregulation of receptor expression mediated by interferon-gamma occurs at the level of transcription
.
Blood
80
,
2363
-
2373
.
Healy
,
T. M.
and
Schulte
,
P. M.
(
2019
).
Patterns of alternative splicing in response to cold acclimation in fish
.
J. Exp. Biol.
222
,
jeb193516
.
Hemmi
,
H.
,
Takeuchi
,
O.
,
Sato
,
S.
,
Yamamoto
,
M.
,
Kaisho
,
T.
,
Sanjo
,
H.
,
Kawai
,
T.
,
Hoshino
,
K.
,
Takeda
,
K.
and
Akira
,
S.
(
2004
).
The roles of two IκB kinase-related kinases in lipopolysaccharide and double stranded RNA signaling and viral infection
.
J. Exp. Med.
199
,
1641
-
1650
.
Hickey
,
M. J.
,
Hagen
,
F. S.
,
Yagi
,
M.
and
Roth
,
G. J.
(
1993
).
Human platelet glycoprotein V: characterization of the polypeptide and the related Ib-V-IX receptor system of adhesive, leucine-rich glycoproteins
.
Proc. Natl. Acad. Sci. USA
90
,
8327
-
8331
.
Hofmann
,
G. E.
,
Buckley
,
B. A.
,
Airaksinen
,
S.
,
Keen
,
J. E.
and
Somero
,
G. N.
(
2000
).
Heat-shock protein expression is absent in the antarctic fish Trematomus bernacchii (family Nototheniidae)
.
J. Exp. Biol.
203
,
2331
-
2339
.
Holmskov
,
U.
,
Mollenhauer
,
J.
,
Madsen
,
J.
,
Vitved
,
L.
,
Grønlund
,
J.
,
Tornøe
,
I.
,
Kliem
,
A.
,
Reid
,
K. B. M.
,
Poustka
,
A.
and
Skjødt
,
K.
(
1999
).
Cloning of gp-340, a putative opsonin receptor for lung surfactant protein D
.
Proc. Natl. Acad. Sci. USA
96
,
10794
-
10799
.
Holzapfel
,
A. M.
and
Vinebrooke
,
R. D.
(
2005
).
Environmental warming increases invasion potential of alpine lake communities by imported species
.
Glob. Change Biol.
11
,
2009
-
2015
.
Hopkins
,
D.
,
Envall
,
T.
,
Poikela
,
N.
,
Pentikäinen
,
O. T.
and
Kankare
,
M.
(
2018
).
Effects of cold acclimation and dsRNA injections on Gs1l gene splicing in Drosophila montana
.
Sci. Rep.
8
,
7577
.
Horvath
,
S
. (
2011
).
Weighted Network Analysis: Applications in Genomics and Systems Biology
.
Springer Science & Business Media.
Hotaling
,
S.
,
Finn
,
D. S.
,
Joseph Giersch
,
J.
,
Weisrock
,
D. W.
and
Jacobsen
,
D.
(
2017
).
Climate change and alpine stream biology: progress, challenges, and opportunities for the future
.
Biol. Rev.
92
,
2024
-
2045
.
Hotaling
,
S.
,
Shah
,
A. A.
,
Mcgowan
,
K. L.
,
Tronstad
,
L. M.
,
Giersch
,
J. J.
,
Finn
,
D. S.
,
Woods
,
H. A.
,
Dillon
,
M. E.
and
Kelley
,
J. L.
(
2020
).
Mountain stoneflies may tolerate warming streams: evidence from organismal physiology and gene expression
.
Glob. Change Biol.
26
,
5524
-
5538
.
Huerta-Cepas
,
J.
,
Szklarczyk
,
D.
,
Heller
,
D.
,
Hernã¡Ndez-Plaza
,
A.
,
Forslund
,
S. K.
,
Cook
,
H.
,
Mende
,
D. R.
,
Letunic
,
I.
,
Rattei
,
T.
,
Jensen
,
L. J.
et al. 
(
2019
).
EggNOG 5.0: a hierarchical, functionally and phylogenetically annotated orthology resource based on 5090 organisms and 2502 viruses
.
Nucleic Acids Res.
47
,
D309
-
D314
.
Ibarz
,
A.
,
Martín-Pérez
,
M.
,
Blasco
,
J.
,
Bellido
,
D.
,
de Oliveira
,
E.
and
Fernández-Borràs
,
J.
(
2010
).
Gilthead sea bream liver proteome altered at low temperatures by oxidative stress
.
Proteomics
10
,
963
-
975
.
Inoue
,
T.
,
Yamashita
,
T.
and
Agata
,
K.
(
2014
).
Thermosensory signaling by TRPM is processed by brain serotonergic neurons to produce planarian thermotaxis
.
J. Neurosci.
34
,
15701
-
15714
.
Jacobsen
,
D.
,
Milner
,
A. M.
,
Brown
,
L. E.
and
Dangles
,
O.
(
2012
).
Biodiversity under threat in glacier-fed river systems
.
Nat. Clim. Change
2
,
361
-
364
.
Jakšić
,
A. M.
,
Karner
,
J.
,
Nolte
,
V.
,
Hsu
,
S.-K.
,
Barghi
,
N.
,
Mallard
,
F.
,
Otte
,
K. A.
,
Svečnjak
,
L.
,
Senti
,
K.-A.
and
Schlötterer
,
C.
(
2020
).
Neuronal function and dopamine signaling evolve at high temperature in drosophila
.
Mol. Biol. Evol.
37
,
2630
-
2640
.
Janzen
,
D. H.
(
1967
).
Why mountain passes are higher in the tropics
.
Am. Nat.
101
,
233
-
249
.
Jeffery
,
C. J.
(
2014
).
An introduction to protein moonlighting
.
Biochem. Soc. Trans.
42
,
1679
-
1683
.
Jones
,
P.
,
Binns
,
D.
,
Chang
,
H.-Y.
,
Fraser
,
M.
,
Li
,
W.
,
Mcanulla
,
C.
,
Mcwilliam
,
H.
,
Maslen
,
J.
,
Mitchell
,
A.
,
Nuka
,
G.
et al. 
(
2014
).
InterProScan 5: genome-scale protein function classification
.
Bioinformatics
30
,
1236
-
1240
.
Kane
,
M.
,
Yadav
,
S. S.
,
Bitzegeio
,
J.
,
Kutluay
,
S. B.
,
Zang
,
T.
,
Wilson
,
S. J.
,
Schoggins
,
J. W.
,
Rice
,
C. M.
,
Yamashita
,
M.
,
Hatziioannou
,
T.
et al. 
(
2013
).
MX2 is an interferon-induced inhibitor of HIV-1 infection
.
Nature
502
,
563
-
566
.
Kanehisa
,
M.
and
Goto
,
S.
(
2000
).
KEGG: Kyoto encyclopedia of genes and genomes
.
Nucleic Acids Res.
28
,
27
-
30
.
Kangale
,
L. J.
,
Raoult
,
D.
,
Fournier
,
P.-E.
,
Abnave
,
P.
and
Ghigo
,
E.
(
2021
).
Planarians (Platyhelminthes)—an emerging model organism for investigating innate immune mechanisms
.
Front. Cell Infect. Microbiol.
11
,
619081
.
Karr
,
T. L.
(
2008
).
Application of proteomics to ecology and population biology
.
Heredity
100
,
200
-
206
.
Kelly
,
M.
(
2019
).
Adaptation to climate change through genetic accommodation and assimilation of plastic phenotypes
.
Philos. Trans. R. Soc. B Biol. Sci.
374
,
20180176
.
Kenkel
,
C. D.
,
Meyer
,
E.
and
Matz
,
M. V.
(
2013
).
Gene expression under chronic heat stress in populations of the mustard hill coral (Porites astreoides) from different thermal environments
.
Mol. Ecol.
22
,
4322
-
4334
.
Kim
,
Y.-G.
,
Lee
,
S.
,
Kwon
,
O.-S.
,
Park
,
S.-Y.
,
Lee
,
S.-J.
,
Park
,
B.-J.
and
Kim
,
K.-J.
(
2009
).
Redox-switch modulation of human SSADH by dynamic catalytic loop
.
EMBO J.
28
,
959
-
968
.
Koh
,
A.
,
Lee
,
M. N.
,
Yang
,
Y. R.
,
Jeong
,
H.
,
Ghim
,
J.
,
Noh
,
J.
,
Kim
,
J.
,
Ryu
,
D.
,
Park
,
S.
,
Song
,
P.
et al. 
(
2013
).
C1-Ten is a protein tyrosine phosphatase of insulin receptor substrate 1 (IRS-1), regulating IRS-1 stability and muscle atrophy
.
Mol. Cell. Biol.
33
,
1608
-
1620
.
Kohtz
,
J. D.
,
Jamison
,
S. F.
,
Will
,
C. L.
,
Zuo
,
P.
,
Lührmann
,
R.
,
Garcia-Blanco
,
M. A.
and
Manley
,
J. L.
(
1994
).
Protein-protein interactions and 5′-splice-site recognition in mammalian mRNA precursors
.
Nature
368
,
119
-
124
.
Kojic
,
S.
,
Nestorovic
,
A.
,
Rakicevic
,
L.
,
Belgrano
,
A.
,
Stankovic
,
M.
,
Divac
,
A.
and
Faulkner
,
G.
(
2010
).
A novel role for cardiac ankyrin repeat protein Ankrd1/CARP as a co-activator of the p53 tumor suppressor protein
.
Arch. Biochem. Biophys.
502
,
60
-
67
.
Kong
,
A. T.
,
Leprevost
,
F. V.
,
Avtonomov
,
D. M.
,
Mellacheruvu
,
D.
and
Nesvizhskii
,
A. I.
(
2017
).
MSFragger: ultrafast and comprehensive peptide identification in mass spectrometry-based proteomics
.
Nat. Methods
14
,
513
-
520
.
Krogh
,
A.
,
Larsson
,
B. Ã.¶
,
Von Heijne
,
G.
and
Sonnhammer
,
E. L. L.
(
2001
).
Predicting transmembrane protein topology with a hidden Markov model: application to complete genomes
.
J. Mol. Biol.
305
,
567
-
580
.
Kroll
,
M. H.
,
Harris
,
T. S.
,
Moake
,
J. L.
,
Handin
,
R. I.
and
Schafer
,
A. I.
(
1991
).
von Willebrand factor binding to platelet GpIb initiates signals for platelet activation
.
J. Clin. Invest.
88
,
1568
-
1573
.
Kuismanen
,
E.
and
Saraste
,
J.
(
1989
).
Low temperature-induced transport blocks as tools to manipulate membrane traffic.
In
Methods in Cell Biology, Chapter 10, Vesicular Transport Part B
(ed.
A. M.
Tartakoff
), pp.
257
-
274
.
Academic Press
.
Kumar
,
L.
and
E Futschik
,
M.
(
2007
).
Mfuzz: a software package for soft clustering of microarray data
.
Bioinformation
2
,
5
-
7
.
Lamprecht
,
A.
,
Semenchuk
,
P. R.
,
Steinbauer
,
K.
,
Winkler
,
M.
and
Pauli
,
H.
(
2018
).
Climate change leads to accelerated transformation of high-elevation vegetation in the central Alps
.
New Phytol.
220
,
447
-
459
.
Langfelder
,
P.
and
Horvath
,
S.
(
2008
).
WGCNA: an R package for weighted correlation network analysis
.
BMC Bioinform.
9
,
559
.
Larade
,
K.
and
Storey
,
K. B.
(
2007
).
Arrest of transcription following anoxic exposure in a marine mollusc
.
Mol. Cell. Biochem.
303
,
243
-
249
.
Laure
,
L.
,
Danièle
,
N.
,
Suel
,
L.
,
Marchand
,
S.
,
Aubert
,
S.
,
Bourg
,
N.
,
Roudaut
,
C.
,
Duguez
,
S.
,
Bartoli
,
M.
and
Richard
,
I.
(
2010
).
A new pathway encompassing calpain 3 and its newly identified substrate cardiac ankyrin repeat protein is involved in the regulation of the nuclear factor-κB pathway in skeletal muscle
.
FEBS J.
277
,
4322
-
4337
.
Lencioni
,
V.
,
Boschini
,
D.
and
Rebecchi
,
L.
(
2009
).
Expression of the 70 kDa Heat shock protein family in Alpine freshwater chironomids (Diptera, Chironomidae) under natural conditions
.
J. Limnol.
68
,
251
-
256
.
Lencioni
,
V.
,
Stella
,
E.
,
Zanoni
,
M. G.
and
Bellin
,
A.
(
2022
).
On the delay between water temperature and invertebrate community response to warming climate
.
Sci. Total Environ.
837
,
155759
.
Li
,
W.
,
Jaroszewski
,
L.
and
Godzik
,
A.
(
2001
).
Clustering of highly homologous sequences to reduce the size of large protein databases
.
Bioinformatics
17
,
282
-
283
.
Lock
,
M. A.
and
Reynoldson
,
T. B.
(
1976
).
The role of interspecific competition in the distribution of two stream Dwelling Triclads, Crenobia alpina (Dana) and Polycelis felina (Dalyell), in North Wales
.
J. Anim. Ecol.
45
,
581
-
592
.
López-García
,
P.
and
Forterre
,
P.
(
2000
).
DNA topology and the thermal stress response, a tale from mesophiles and hyperthermophiles
.
BioEssays
22
,
738
-
746
.
Ma
,
W.
,
Kim
,
S.
,
Chowdhury
,
S.
,
Li
,
Z.
,
Yang
,
M.
,
Yoo
,
S.
,
Petralia
,
F.
,
Jacobsen
,
J.
,
Li
,
J. J.
,
Ge
,
X.
et al. 
(
2020
).
DreamAI: algorithm for the imputation of proteomics data
.
bioRxiv, 2020.07.21.214205
.
Mádi
,
A.
,
Mikkat
,
S.
,
Ringel
,
B.
,
Ulbrich
,
M.
,
Thiesen
,
H.-J. Ã.¼
and
Glocker
,
M. O.
(
2003
).
Mass spectrometric proteome analysis for profiling temperature-dependent changes of protein expression in wild-type Caenorhabditis elegans
.
Proteomics
3
,
1526
-
1534
.
Malan
,
A.
and
Canguilhem
,
B.
(
1989
).
Living in the Cold
.
John Libbey Eurotext
.
Malek
,
R. L.
,
Sajadi
,
H.
,
Abraham
,
J.
,
Grundy
,
M. A.
and
Gerhard
,
G. S.
(
2004
).
The effects of temperature reduction on gene expression and oxidative stress in skeletal muscle from adult zebrafish
.
Comp. Biochem. Physiol. C Toxicol. Pharmacol.
138
,
363
-
373
.
Mcadams
,
H. H.
and
Arkin
,
A.
(
1997
).
Stochastic mechanisms in gene expression
.
Proc. Natl. Acad. Sci. USA
94
,
814
-
819
.
Mercer
,
C. A.
,
Kaliappan
,
A.
and
Dennis
,
P. B.
(
2009
).
A novel, human Atg13 binding protein, Atg101, interacts with ULK1 and is essential for macroautophagy
.
Autophagy
5
,
649
-
662
.
Moldovan
,
O. T.
,
Baricz
,
A.
,
Szekeres
,
E.
,
Kenesz
,
M.
,
Hoaghia
,
M. A.
,
Levei
,
E. A.
,
Mirea
,
I.
,
Năstase-Bucur
,
R.
,
Brad
,
T.
,
Chiciudean
,
I.
et al. 
(
2020
).
Testing different membrane filters for 16S rRNA gene-based metabarcoding in karstic springs
.
Water
12
,
3400
.
Morimoto
,
R. I.
(
2011
).
The heat shock response: systems biology of proteotoxic stress in aging and disease
.
Cold Spring Harbor Symp. Quant. Biol.
76
,
91
-
99
.
Muhlfeld
,
C. C.
,
Cline
,
T. J.
,
Giersch
,
J. J.
,
Peitzsch
,
E.
,
Florentine
,
C.
,
Jacobsen
,
D.
and
Hotaling
,
S.
(
2020
).
Specialized meltwater biodiversity persists despite widespread deglaciation
.
Proc. Natl. Acad. Sci. USA
117
,
12208
-
12214
.
Nishimura
,
K.
,
Kitamura
,
Y.
,
Inoue
,
T.
,
Umesono
,
Y.
,
Sano
,
S.
,
Yoshimoto
,
K.
,
Inden
,
M.
,
Takata
,
K.
,
Taniguchi
,
T.
,
Shimohama
,
S.
et al. 
(
2007
).
Reconstruction of dopaminergic neural network and locomotion function in planarian regenerates
.
Dev. Neurobiol.
67
,
1059
-
1078
.
Nishimura
,
K.
,
Kitamura
,
Y.
,
Umesono
,
Y.
,
Takeuchi
,
K.
,
Takata
,
K.
,
Taniguchi
,
T.
and
Agata
,
K.
(
2008
).
Identification of glutamic acid decarboxylase gene and distribution of GABAergic nervous system in the planarian Dugesia japonica
.
Neuroscience
153
,
1103
-
1114
.
Novick
,
P.
and
Zerial
,
M.
(
1997
).
The diversity of Rab proteins in vesicle transport
.
Curr. Opin. Cell Biol.
9
,
496
-
504
.
Nyegaard
,
M.
,
Overgaard
,
M. T.
,
Søndergaard
,
M. T.
,
Vranas
,
M.
,
Behr
,
E. R.
,
Hildebrandt
,
L. L.
,
Lund
,
J.
,
Hedley
,
P. L.
,
Camm
,
A. J.
,
Wettrell
,
G.
et al. 
(
2012
).
Mutations in calmodulin cause ventricular tachycardia and sudden cardiac death
.
Am. J. Hum. Genet.
91
,
703
-
712
.
Palakodeti
,
D.
,
Smielewska
,
M.
,
Lu
,
Y.-C.
,
Yeo
,
G. W.
and
Graveley
,
B. R.
(
2008
).
The PIWI proteins SMEDWI-2 and SMEDWI-3 are required for stem cell function and piRNA expression in planarians
.
RNA
14
,
1174
-
1186
.
Palgi
,
M.
,
Lindström
,
R.
,
Peränen
,
J.
,
Piepponen
,
T. P.
,
Saarma
,
M.
and
Heino
,
T. I.
(
2009
).
Evidence that DmMANF is an invertebrate neurotrophic factor supporting dopaminergic neurons
.
Proc. Natl. Acad. Sci. USA
106
,
2429
-
2434
.
Palumbi
,
S. R.
,
Barshis
,
D. J.
,
Traylor-Knowles
,
N.
and
Bay
,
R. A.
(
2014
).
Mechanisms of reef coral resistance to future climate change
.
Science
344
,
895
-
898
.
Paskin
,
T. R.
,
Jellies
,
J.
,
Bacher
,
J.
and
Beane
,
W. S.
(
2014
).
Planarian Phototactic Assay Reveals Differential Behavioral Responses Based on Wavelength
.
PLoS One
9
,
e114708
.
Pattee
,
E.
(
1965
).
Sténothermie et eurythermie les invertébrés d'eau douce et la variation journalière de température
.
Ann. Limnol.
1
,
281
-
434
.
Pearson
,
B. J.
,
Eisenhoffer
,
G. T.
,
Gurley
,
K. A.
,
Rink
,
J. C.
,
Miller
,
D. E.
and
Sánchez Alvarado
,
A.
(
2009
).
Formaldehyde-based whole-mount in situ hybridization method for planarians
.
Dev. Dyn.
238
,
443
-
450
.
Perez-Riverol
,
Y.
,
Csordas
,
A.
,
Bai
,
J.
,
Bernal-Llinares
,
M.
,
Hewapathirana
,
S.
,
Kundu
,
D. J.
,
Inuganti
,
A.
,
Griss
,
J.
,
Mayer
,
G.
,
Eisenacher
,
M.
et al. 
(
2019
).
The PRIDE database and related tools and resources in 2019: improving support for quantification data
.
Nucleic Acids Res.
47
,
D442
-
D450
.
Raj
,
A.
and
Van Oudenaarden
,
A.
(
2008
).
Nature, nurture, or chance: stochastic gene expression and its consequences
.
Cell
135
,
216
-
226
.
Reddien
,
P. W.
,
Bermange
,
A. L.
,
Murfitt
,
K. J.
,
Jennings
,
J. R.
and
Sánchez Alvarado
,
A.
(
2005a
).
Identification of genes needed for regeneration, stem cell function, and tissue homeostasis by systematic gene perturbation in planaria
.
Dev. Cell
8
,
635
-
649
.
Reddien
,
P. W.
,
Oviedo
,
N. J.
,
Jennings
,
J. R.
,
Jenkin
,
J. C.
and
Sánchez Alvarado
,
A
. (
2005b
).
SMEDWI-2 Is a PIWI-like protein that regulates planarian stem cells
.
Science
310
,
1327
-
1330
.
Available at: https://www.science.org/doi/abs/10.1126/science.1116110 (Accessed: 2 September 2021)
.
Reusch
,
T. B. H.
(
2014
).
Climate change in the oceans: evolutionary versus phenotypically plastic responses of marine animals and plants
.
Evol. Appl.
7
,
104
-
122
.
Reynoldson
,
T. B.
(
1981
).
The ecology of the Turbellaria with special reference to the freshwater triclads
.
Hydrobiologia
84
,
87
-
90
.
Roca
,
J. R.
,
Ribas
,
M.
and
Baguñà
,
J.
(
1992
).
Distribution, ecology, mode of reproduction and karyology of freshwater planarians (Platyhelminthes; Turbellaria; Tricladida) in the springs of the central Pyrenees
.
Ecography
15
,
373
-
384
.
Rogora
,
M.
,
Frate
,
L.
,
Carranza
,
M. L.
,
Freppaz
,
M.
,
Stanisci
,
A.
,
Bertani
,
I.
,
Bottarin
,
R.
,
Brambilla
,
A.
,
Canullo
,
R.
,
Carbognani
,
M.
et al. 
(
2018
).
Assessment of climate change effects on mountain ecosystems through a cross-site analysis in the Alps and Apennines
.
Sci. Total Environ.
624
,
1429
-
1442
.
Rompolas
,
P.
,
Azimzadeh
,
J.
,
Marshall
,
W. F.
and
King
,
S. M.
(
2013
).
Analysis of ciliary assembly and function in planaria
.
Methods Enzymol.
525
,
245
-
264
.
Rorsman
,
F.
,
Husebye
,
E. S.
,
Winqvist
,
O.
,
Björk
,
E.
,
Karlsson
,
F. A.
and
Kämpe
,
O.
(
1995
).
Aromatic-L-amino-acid decarboxylase, a pyridoxal phosphate-dependent enzyme, is a beta-cell autoantigen
.
Proc. Natl. Acad. Sci. USA
92
,
8626
-
8629
.
Roth
,
D. M.
,
Hutt
,
D. M.
,
Tong
,
J.
,
Bouchecareilh
,
M.
,
Wang
,
N.
,
Seeley
,
T.
,
Dekkers
,
J. F.
,
Beekman
,
J. M.
,
Garza
,
D.
,
Drew
,
L.
et al. 
(
2014
).
Modulation of the maladaptive stress response to manage diseases of protein folding
.
PLoS Biol.
12
,
e1001998
.
Rozanski
,
A.
,
Moon
,
H. K.
,
Brandl
,
H.
,
Martín -Durán
,
J. M.
,
Grohme
,
M. A.
,
Hüttner
,
K.
,
Bartscherer
,
K.
,
Henry
,
I.
and
Rink
,
J. C.
(
2019
).
PlanMine 3.0-improvements to a mineable resource of flatworm biology and biodiversity
.
Nucleic Acids Res.
47
,
D812
-
D820
.
Rubidge
,
E. M.
,
Patton
,
J. L.
,
Lim
,
M.
,
Burton
,
A. C.
,
Brashares
,
J. S.
and
Moritz
,
C.
(
2012
).
Climate-induced range contraction drives genetic erosion in an alpine mammal
.
Nat. Clim. Change
2
,
285
-
288
.
Salachan
,
P. V.
and
Sørensen
,
J. G.
(
2022
).
Molecular mechanisms underlying plasticity in a thermally varying environment
.
Mol. Ecol.
31
,
3174
-
3191
.
Sarmiento
,
L.
,
Frisk
,
G.
,
Anagandula
,
M.
,
Hodik
,
M.
,
Barchetta
,
I.
,
Netanyah
,
E.
,
Cabrera-Rode
,
E.
and
Cilio
,
C.
(
2017
).
Echovirus 6 infects human exocrine and endocrine pancreatic cells and induces pro-inflammatory innate immune response
.
Viruses
9
,
25
.
Savio
,
D.
,
Stadler
,
P.
,
Reischer
,
G. H.
,
Kirschner
,
A. K. T.
,
Demeter
,
K.
,
Linke
,
R.
,
Blaschke
,
A. P.
,
Sommer
,
R.
,
Szewzyk
,
U.
,
Wilhartitz
,
I. C.
et al. 
(
2018
).
Opening the black box of spring water microbiology from alpine karst aquifers to support proactive drinking water resource management
.
WIREs Water
5
,
e1282
.
Schlieper
,
C. B. J.
and
Halsband
,
E.
(
1952
).
Experimentelle Veränderungen der Temperaturtoleranz bei stenothermen und eurythermen Wassertieren
.
Zool. Anz.
149
,
163
-
169
.
Schoville
,
S. D.
,
Slatyer
,
R. A.
,
Bergdahl
,
J. C.
and
Valdez
,
G. A.
(
2015
).
Conserved and narrow temperature limits in alpine insects: thermal tolerance and supercooling points of the ice-crawlers, Grylloblatta (Insecta: Grylloblattodea: Grylloblattidae)
.
J. Insect Physiol.
78
,
55
-
61
.
Seebacher
,
F.
,
White
,
C. R.
and
Franklin
,
C. E.
(
2015
).
Physiological plasticity increases resilience of ectothermic animals to climate change
.
Nat. Clim. Change
5
,
61
-
66
.
Shah
,
A. A.
,
Dillon
,
M. E.
,
Hotaling
,
S.
and
Woods
,
H. A.
(
2020
).
High elevation insect communities face shifting ecological and evolutionary landscapes
.
Curr. Opin. Insect Sci.
41
,
1
-
6
.
Shalgi
,
R.
,
Hurt
,
J. Â. A.
,
Krykbaeva
,
I.
,
Taipale
,
M.
,
Lindquist
,
S.
and
Burge
,
C. Â. B.
(
2013
).
Widespread regulation of translation by elongation pausing in heat shock
.
Mol. Cell
49
,
439
-
452
.
Sheikh
,
M. S.
and
Fornace
,
A. J.
(
1999
).
Regulation of translation initiation following stress
.
Oncogene
18
,
6121
-
6128
.
Sherman
,
M. Y.
and
Goldberg
,
A. L.
(
2001
).
Cellular defenses against unfolded proteins: a cell biologist thinks about neurodegenerative diseases
.
Neuron
29
,
15
-
32
.
Smyth
,
G. K.
(
2005
).
limma: linear models for microarray data
. In
Bioinformatics and Computational Biology Solutions Using R and Bioconductor
(ed.
R.
Gentleman
et al. 
), pp.
397
-
420
.
New York, NY
:
Springer (Statistics for Biology and Health)
.
Somero
,
G. N.
(
1995
).
Proteins and temperature
.
Annu. Rev. Physiol.
57
,
43
-
68
.
Somero
,
G.
(
1996
).
Temperature and proteins: little things can mean a lot
.
Physiology
11
,
72
-
77
.
Supek
,
F.
,
Bošnjak
,
M.
, Š
kunca
,
N.
and
Šmuc
,
T.
(
2011
).
REVIGO summarizes and visualizes long lists of gene ontology terms
.
PLoS One
6
,
e21800
.
Tatusov
,
R. L.
,
Galperin
,
M. Y.
,
Natale
,
D. A.
and
Koonin
,
E. V.
(
2000
).
The COG database: a tool for genome-scale analysis of protein functions and evolution
.
Nucleic Acids Res.
28
,
33
-
36
.
Tegeder
,
I.
,
Costigan
,
M.
,
Griffin
,
R. S.
,
Abele
,
A.
,
Belfer
,
I.
,
Schmidt
,
H.
,
Ehnert
,
C.
,
Nejim
,
J.
,
Marian
,
C.
,
Scholz
,
J.
et al. 
(
2006
).
GTP cyclohydrolase and tetrahydrobiopterin regulate pain sensitivity and persistence
.
Nat. Med.
12
,
1269
-
1277
.
Terblanche
,
J. S.
,
Deere
,
J. A.
,
Clusella-Trullas
,
S.
,
Janion
,
C.
and
Chown
,
S. L.
(
2007
).
Critical thermal limits depend on methodological context
.
Proc. R. Soc. B
274
,
2935
-
2943
.
Terblanche
,
J. S.
,
Hoffmann
,
A. A.
,
Mitchell
,
K. A.
,
Rako
,
L.
,
Le Roux
,
P. C.
and
Chown
,
S. L.
(
2011
).
Ecologically relevant measures of tolerance to potentially lethal temperatures
.
J. Exp. Biol.
214
,
3713
-
3725
.
Thienemann
,
A.
(
1950
). ‘
Verbreitungsgeschichte der Süßwassertierwelt Europas’. Available at: https://www.schweizerbart.de/publications/detail/isbn/9783510407279 (Accessed: 30 June 2020). ISBN 978-3-510-40727-9
Thieringer
,
H. A.
,
Jones
,
P. G.
and
Inouye
,
M.
(
1998
).
Cold shock and adaptation
.
BioEssays
20
,
49
-
57
.
Tiitu
,
V.
and
Vornanen
,
M.
(
2002
).
Regulation of cardiac contractility in a cold stenothermal fish, the burbot Lota lota L
.
J. Exp. Biol.
205
,
1597
-
1606
.
Timmons
,
J. A.
,
Szkop
,
K. J.
and
Gallagher
,
I. J.
(
2015
).
Multiple sources of bias confound functional enrichment analysis of global -omics data
.
Genome Biol.
16
,
186
.
Tomanek
,
L.
(
2012
).
Environmental proteomics of the mussel mytilus: implications for tolerance to stress and change in limits of biogeographic ranges in response to climate change
.
Integr. Comp. Biol.
52
,
648
-
664
.
Tomanek
,
L.
(
2014
).
Proteomics to study adaptations in marine organisms to environmental stress
.
J. Proteomics
105
,
92
-
106
.
Tomczak
,
A.
,
Mortensen
,
J. M.
,
Winnenburg
,
R.
,
Liu
,
C.
,
Alessi
,
D. T.
,
Swamy
,
V.
,
Vallania
,
F.
,
Lofgren
,
S.
,
Haynes
,
W.
,
Shah
,
N. H.
et al. 
(
2018
).
Interpretation of biological experiments changes with evolution of the Gene Ontology and its annotations
.
Sci. Rep.
8
,
5115
.
Treanor
,
H. B.
,
Giersch
,
J. J.
,
Kappenman
,
K. M.
,
Muhlfeld
,
C. C.
and
Webb
,
M. A. H.
(
2013
).
Thermal tolerance of meltwater stonefly Lednia tumana nymphs from an alpine stream in Waterton–Glacier International Peace Park, Montana, USA
.
Freshw. Sci.
32
,
597
-
605
.
Tsai
,
C.-J.
,
Maizel
,
J. V.
and
Nussinov
,
R.
(
2002
).
The hydrophobic effect: a new insight from cold denaturation and a two-state water structure
.
Crit. Rev. Biochem. Mol. Biol.
37
,
55
-
69
.
Tskhovrebova
,
L.
and
Trinick
,
J.
(
2003
).
Titin: properties and family relationships
.
Nat. Rev. Mol. Cell Biol.
4
,
679
-
689
.
Ueno
,
T.
,
Tomita
,
J.
,
Kume
,
S.
and
Kume
,
K.
(
2012
).
Dopamine modulates metabolic rate and temperature sensitivity in drosophila melanogaster
.
PLoS One
7
,
e31513
.
Välikangas
,
T.
,
Suomi
,
T.
and
Elo
,
L. L.
(
2018
).
A systematic evaluation of normalization methods in quantitative label-free proteomics
.
Brief. Bioinform.
19
,
1
-
11
.
Veilleux
,
H. D.
,
Ryu
,
T.
,
Donelson
,
J. M.
,
Van Herwerden
,
L.
,
Seridi
,
L.
,
Ghosheh
,
Y.
,
Berumen
,
M. L.
,
Leggat
,
W.
,
Ravasi
,
T.
and
Munday
,
P. L.
(
2015
).
Molecular processes of transgenerational acclimation to a warming ocean
.
Nat. Clim. Change
5
,
1074
-
1078
.
Vila-Farré
,
M.
and
C. Rink
,
J.
(
2018
).
The ecology of freshwater planarians
.
Methods Mol. Biol.
1774
,
173
-
205
.
Voigt
,
W.
(
1895
).
Planaria gonocephala als Eindringling in das Verbreitungsgebiet von Planaria alpina und Polycelis cornuta
.
Zool. Jahrb.
8
,
131
-
176
.
Wang
,
X.
,
Venable
,
J.
,
Lapointe
,
P.
,
Hutt
,
D. M.
,
Koulov
,
A. V.
,
Coppinger
,
J.
,
Gurkan
,
C.
,
Kellner
,
W.
,
Matteson
,
J.
,
Plutner
,
H.
et al. 
(
2006
).
Hsp90 cochaperone Aha1 downregulation rescues misfolding of CFTR in cystic fibrosis
.
Cell
127
,
803
-
815
.
Wang
,
C.
,
Yang
,
Z.-Z.
,
Guo
,
F.-H.
,
Shi
,
S.
,
Han
,
X.-S.
,
Zeng
,
A.
,
Lin
,
H.
and
Jing
,
Q.
(
2019
).
Heat shock protein DNAJA1 stabilizes PIWI proteins to support regeneration and homeostasis of planarian Schmidtea mediterranea
.
J. Biol. Chem.
294
,
9873
-
9887
.
Watabe
,
S.
(
2002
).
Temperature plasticity of contractile proteins in fish muscle
.
J. Exp. Biol.
205
,
2231
-
2236
.
Waterhouse
,
A.
,
Bertoni
,
M.
,
Bienert
,
S.
,
Studer
,
G.
,
Tauriello
,
G.
,
Gumienny
,
R.
,
Heer
,
F. T.
,
Deâ Beer
,
T. AÂ. P.
,
Rempfer
,
C.
,
Bordoli
,
L.
et al. 
(
2018
).
SWISS-MODEL: homology modelling of protein structures and complexes
.
Nucleic Acids Res.
46
,
W296
-
W303
.
Węsławski
,
J. M.
,
Maurizi
,
M. R.
and
Gottesman
,
S.
(
2021
).
Gammarus (Amphipoda) species competitive exclusion or coexistence as a result of climate change in the Arctic?
Pol. Polar Res.
42
,
287
-
302
.
Wickner
,
S.
,
Maurizi
,
M. R.
and
Gottesman
,
S.
(
1999
).
Posttranslational quality control: folding, refolding, and degrading proteins
.
Science
286
,
1888
-
1893
.
Winder
,
S. J.
,
Allen
,
B. G.
,
Clément-Chomienne
,
O.
and
Walsh
,
M. P.
(
1998
).
Regulation of smooth muscle actin—myosin interaction and force by calponin
.
Acta Physiol. Scand.
164
,
415
-
426
.
Wittenberg
,
J. B.
(
1970
).
Myoglobin-facilitated oxygen diffusion: role of myoglobin in oxygen entry into muscle
.
Physiol. Rev.
50
,
559
-
636
.
Woolbright
,
S. A.
,
Whitham
,
T. G.
,
Gehring
,
C. A.
,
Allan
,
G. J.
and
Bailey
,
J. K.
(
2014
).
Climate relicts and their associated communities as natural ecology and evolution laboratories
.
Trends Ecol. Evol.
29
,
406
-
416
.
Yu
,
G.
(
2020
).
enrichplot: Visualization of Functional Enrichment Results. R package version 1.10.1. https://yulab-smu.top/biomedical-knowledge-mining-book/
Yu
,
G.
,
Wang
,
L.-G.
,
Han
,
Y.
and
He
,
Q.-Y.
(
2012
).
clusterProfiler: an R package for comparing biological themes among gene clusters
.
OMICS
16
,
284
-
287
.
Yu
,
F.
,
Haynes
,
S. E.
and
Nesvizhskii
,
A. I.
(
2021
).
IonQuant enables accurate and sensitive label-free quantification with FDR-controlled match-between-runs
.
Mol. Cell. Proteomics
20
,
100077
.
Zerial
,
M.
and
Mcbride
,
H.
(
2001
).
Rab proteins as membrane organizers
.
Nat. Rev. Mol. Cell Biol.
2
,
107
-
117
.
Zhang
,
B.
and
Horvath
,
S.
(
2005
).
A general framework for weighted gene co-expression network analysis
.
Stat. Appl. Genet. Mol. Biol.
4
,
Article17
.

Competing interests

The authors declare no competing or financial interests.

This is an Open Access article distributed under the terms of the Creative Commons Attribution License (https://creativecommons.org/licenses/by/4.0), which permits unrestricted use, distribution and reproduction in any medium provided that the original work is properly attributed.

Supplementary information