Effects of nutrient and water restriction on thermal tolerance: A test of mechanisms and hypotheses

https://doi.org/10.1016/j.cbpa.2017.06.019Get rights and content

Abstract

Nutritional deprivation or desiccation can influence thermal tolerance by impacting the insects' ability to evaporatively cool, maintain cell membrane integrity and conduct protective or repair processes. Recovery from chilling is also linked to the re-establishment of iono- and osmo-regulatory homeostasis. Here, using Mediterranean fruit fly (Ceratitis capitata, Diptera: Tephritidae), we manipulated water and nutrient availability to test the mechanistic expectation that changes in whole organism lipid and water content can elicit variation in cold or heat tolerance (scored as chill coma recovery time and heat knockdown time). We measured body condition (body water and lipid content) as well as heat shock protein 70 gene (hsp70) and protein (HSP70) levels. A significant reduction in body water content with water restriction did not translate into differences in chill coma recovery. When nutrient restriction was coupled with water deprivation, this resulted in a significant reduction (− 54%) of heat knockdown time in females but male flies were unaffected. There was no evidence for an hsp70 or HSP70 response under any of the stress treatments and therefore no correlation with heat or cold tolerance. Heat hardening decreased all hsp levels. Therefore, although body water and total body lipid content differed between the treatment groups, the contribution of these factors to thermal tolerance was inconsistent with mechanistic expectations in heat knockdown time and insignificant for chill coma recovery. These results therefore highlight that the effects of resource restriction on thermal limits in insects are mechanistically more complex than previous models of stress resistance have suggested.

Introduction

Traits of thermal tolerance or resistance are strongly correlated with insect species' geographic distributions (reviewed in Addo-Bediako et al., 2000), and assessing the factors influencing these traits in the field is important (e.g. Sinclair and Roberts, 2005, Cooper et al., 2008, Terblanche et al., 2011). Most studies use laboratory-based methodology to control for a host of intrinsic and extrinsic factors that can influence thermal tolerance estimates, such as recent thermal history (e.g. season, acclimation), age or sex (e.g. Fischer et al., 2010). The impact of body condition on tolerance estimates is typically not the subject of investigation (but see e.g. Terblanche et al., 2008, Overgaard et al., 2012). As the temporary deprivation of food and/or water can have downstream impacts on organismal performance and fitness, tolerating these stressors is essential for insects in heterogeneous landscapes. Mathematical models have also highlighted how thermal traits may be influenced by nutritional and water status at the organismal level (Kearney et al., 2013). Complex interactions and fitness trade-offs occur between stressors in at least some instances (reviewed in Hoffmann et al., 2003, Mellanby, 1932, Bubliy et al., 2012a, Bubliy et al., 2012b, Overgaard et al., 2012, Kellermann et al., 2013, Karl et al., 2014, Boardman et al., 2015, Scharf et al., 2016). The different approaches typically employed, and range of stressors considered may elicit very different underlying physiological processes, or similar trait responses may arise from distinctly different mechanisms. Thus, determining the mechanistic underpinnings of multi-stressor responses to changing body condition is a crucial avenue for forecasting species responses to changing environments.

There are at least three major expectations for how stress resistance may respond under changing nutritional or hydration conditions at the organismal level. First, the organism may already have the necessary cellular biochemistry in place to fully withstand the stress, and therefore no effect would be detected on thermal tolerances. Second, exposure to stress (e.g. food or water deprivation) may result in upregulation of protective (or repair) mechanisms that allows the organisms to better withstand subsequent thermally stressful conditions. Finally, biochemical or physiological responses may be upregulated but are insufficient to compensate for the subsequent thermal stress. Predictable trait responses can be generated to differentiate among these possibilities.

Desiccation prior to a heat or cold tolerance assay would generally be expected to reduce overall thermal tolerance by restricting available water resources that underlie multiple physiological traits (reviewed in Zera and Harshman, 2001). However, some studies have found a positive association between prior desiccation stress and heat tolerance (e.g. Benoit et al., 2009, Bubliy et al., 2012a), largely attributed to the upregulation of heat shock proteins (Hsps) that act as molecular chaperones for damaged proteins (Feder and Hofmann, 1999). However, the association is equivocal depending on stress type and exposure duration (e.g. Feder et al., 1992, Silbermann and Tatar, 2000, Boardman et al., 2013, Sørensen et al., 2013). Desiccation has been associated with improved cold hardiness by potentially increasing the osmolality of the tissue, which in turn, decreases the temperature at which the organism freezes (supercooling point; reviewed in Zachariassen, 1985). Cryoprotective dehydration, whereby an organism loses water during cold stress to avoid ice formation in tissues, is an important low temperature survival strategy found in many insects (e.g. Sørensen and Holmstrup, 2011; reviewed in Chown and Terblanche, 2006). However, if the organism does not freeze, then this increase in osmolality may detrimentally affect the individuals' ability to regain ion homeostasis, the latter of which is the major hypothesis describing insects' inability to recover from chill coma (e.g. Coello Alvarado et al., 2015, Olsson et al., 2016, reviewed in Overgaard and MacMillan, 2017).

Nutritional stress might influence thermal tolerance primarily through reductions in body lipid levels. Lipid levels and the composition thereof (e.g. long vs. short chains, number of double vs. single bonds) are thought to underpin the link between starvation and thermal tolerance in Drosophila (e.g. Hoffmann and Harshman, 1999, Hoffmann et al., 2005). For example, a negative association is found between starvation resistance and cold tolerance due to the importance of lipids in both traits (e.g. Hoffmann et al., 2005). Reductions in lipid levels are generally associated with starvation across all organisms (McCue, 2010) but may result in increased water uptake (e.g. Raubenheimer and Gäde, 1993) which can be beneficial for improving heat tolerance by increasing the water stores available for evaporative cooling or generally withstanding desiccation stress (Chown and Nicholson, 2004).

Here we provide a systematic, comprehensive assessment of these potential mechanisms that have been hypothesised to underlie the influence of water and nutritional status on thermal tolerance using the Mediterranean fruit fly, Ceratitis capitata (Wiedemann) (Diptera: Tephritidae). As a global agricultural pest there is considerable interest in better understanding its basal and plastic climatic stress resistance, and the underlying mechanisms thereof (De Meyer et al., 2008, Malacrida et al., 2007, Nyamukondiwa et al., 2010). Previous studies have measured chill coma recovery time and heat knockdown time (Weldon et al., 2011), shown a pronounced decrease in heat tolerance (measured as CTmax) following prior starvation (Nyamukondiwa and Terblanche, 2009) and plastic responses to prior thermal exposures (Nyamukondiwa et al., 2010) that were attributed to variation in hsp70 mRNA expression (Kalosaka et al., 2009, but see Pujol-Lereis et al., 2014). We exposed C. capitata to a full factorial experimental design of long term (i.e. several days as adults) restriction of nutrients, water and the combined restriction of both nutrient and water and then assayed chill coma recovery time and heat knockdown time. We also estimated body water and body lipid amounts, and the relative amount of mRNA hsp70, HSP70 protein and total protein concentration to explain any potential thermal tolerance responses. We predicted that restricting water or nutrient levels would have a detrimental effect on heat tolerance and, by contrast, be beneficial for cold tolerance. Furthermore, it can be predicted that if thermal tolerance is set in some mechanistic way by an organism's body condition, then combined stressors of nutrient and water restriction should have a greater reduction on thermal limits than either stressor applied individually, and these latter should reduce thermal limits more than a control or reference group. Heat tolerance is likely to be impacted by reduced water stores by limiting evaporative cooling and/or reducing the pool of water available to survive prolonged stress. Nutrient restriction may also be detrimental for heat tolerance by reducing the amount of stored lipids that can be metabolised for water. The change in saturation of cell membranes caused by reduced nutrient levels would be beneficial for cold tolerance. Reduced water levels would benefit cold tolerance by reducing the formation of ice within the tissue. However, if the nutrient restriction results in an increased uptake in water, heat tolerance may benefit from increased structural integrity of cell membranes i.e. through lipid saturation, and more available body water for evaporative cooling. Cross-tolerance may also be detected if prior depletion of nutrients or water results in heat shock protein production.

Section snippets

Experimental treatments

Ceratitis capitata (Diptera: Tephritidae) pupae were obtained from laboratory stocks maintained in large laboratory cultures at Citrus Research International (CRI) at Nelspruit, South Africa that have been established for c. 15 years (or ± 180 generations). Colonies were last partially substituted with wild-caught flies in 2011 (A. Manrakhan, pers. comm.). One-day old adults were separated by sex into groups of 10 flies, placed into ventilated 5 L plastic containers at 25 °C ± 1 °C (LE-509 incubator,

Thermal tolerance

There was no significant effect of the nutrient or water restriction on chill coma recovery time in C. capitata (Fig. 1A, Table 1; means ± SE chill coma recovery of control groups: females, 8.82 ± 1.38 min; males, 9.55 ± 1.30 min). There were also no overall differences in chill coma recovery between the sexes in this trait (Fig. 1A, Table 1). In contrast, heat knockdown time was significantly influenced by experimental treatments (Table 1), (means ± SE heat knockdown time of control groups: females, 8.62

Discussion

Lipid amount and composition, and associated water availability, feature most prominently among the list of implicated mechanisms of cross-tolerance with stress resistance, along with the generic stress response of heat shock proteins. Cross-tolerance, or protective mechanisms that arise from prior exposure to a different stressor (e.g. Sinclair et al., 2013) is specific to the stress experienced, while the genetic background of the population is also known to substantially influence results

Conclusions

Despite an extensive body of evidence of cross-tolerance and nutrient and water restriction on estimates of thermal tolerance, particularly for cold tolerance traits (e.g. Bubliy et al., 2012a, Bubliy et al., 2012b), we found no positive correlation between nutrition and desiccation stress and cold tolerance in C. capitata but i) a positive association between nutrient restriction and heat tolerance and ii) a negative association with the combined nutrient and water restriction for heat

Acknowledgements

The authors thank Janina von Diest for early technical assistance, Dr. Aruna Manrankhan at CRI for providing specimens, the Central Analytical Facility at Stellenbosch University for assistance with RT-PCR analysis, the Biochemistry Department for access to plate reader for HSP70 protein absorbance and the anonymous referees that provided constructive criticism of an earlier version of this manuscript. KAM was supported by the Claude Leon foundation. LB was supported by an NRF Innovation

References (59)

  • C. Nyamukondiwa et al.

    Thermal tolerance in adult Mediterranean and Natal fruit flies (Ceratitis capitata and Ceratitis rosa): effects of age, gender and feeding status

    J. Therm. Biol.

    (2009)
  • L.M. Pujol-Lereis et al.

    Analysis of survival, gene expression and behavior following chill-coma in the medfly Ceratitis capitata: effects of population heterogeneity and age

    J. Insect Physiol.

    (2014)
  • D. Raubenheimer et al.

    Compensatory water intake by locusts (Locusta migratoria): implications for mechanisms regulating drink size

    J. Insect Physiol.

    (1993)
  • M.E. Salvucci et al.

    Heat shock proteins in whiteflies, an insect that accumulates sorbitol in response to heat stress

    J. Therm. Biol.

    (2000)
  • B.J. Sinclair et al.

    Acclimation, shock and hardening in the cold

    J. Therm. Biol.

    (2005)
  • J.G. Sørensen et al.

    Cryoprotective dehydration is widespread in Arctic springtails

    J. Insect Physiol.

    (2011)
  • G. Stephanou et al.

    Heat shock response in Ceratitis capitata

    Comp. Biochem. Physiol. B

    (1983)
  • J.S. Terblanche et al.

    Thermal tolerance in a south-east African population of the tsetse fly Glossina pallidipes (Diptera, Glossinidae): implications for forecasting climate change impacts

    J. Insect Physiol.

    (2008)
  • C.W. Weldon et al.

    Time-course for attainment and reversal of acclimation to constant temperature in two Ceratitis species

    J. Therm. Biol.

    (2011)
  • A. Addo-Bediako et al.

    Thermal tolerance, climatic variability and latitude

    Proc. Biol. Sci. B

    (2000)
  • O.A. Bubliy et al.

    Plastic responses of four environmental stresses and cross-resistance in a laboratory population of Drosophila melanogaster

    Funct. Ecol.

    (2012)
  • O.A. Bubliy et al.

    Humidity affects genetic architecture of heat resistance in Drosophila melanogaster

    J. Evol. Biol.

    (2012)
  • J.R. Carey et al.

    Relationship of age patterns of fecundity to mortality, longevity, and lifetime reproduction in a large cohort of Mediterranean fruit fly females

    J. Gerontol. Biol. Sci.

    (1998)
  • S.L. Chown et al.

    Insect Physiological Ecology: Mechanisms and Patterns

    (2004)
  • S.L. Chown et al.

    Phenotypic variance, plasticity and heritability estimates of critical thermal limits depend on methodological context

    Funct. Ecol.

    (2009)
  • S. Clusella-Trullas et al.

    Effects of temperature on heat-shock responses and survival of two species of marine invertebrates from sub-Antarctic Marion Island

    Antarct. Sci.

    (2014)
  • M. De Meyer et al.

    Ecological niches and potential geographic distributions of Mediterranean fruit fly (Ceratitis capitata) and Natal fruit fly (Ceratitis rosa)

    J. Biogeogr.

    (2008)
  • M.E. Feder et al.

    Heat-shock proteins, molecular chaperones, and the stress response: evolutionary and ecological physiology

    Annu. Rev. Physiol.

    (1999)
  • J.H. Feder et al.

    The consequences of expressing Hsp70 in Drosophila cells at normal temperatures

    Genes Dev.

    (1992)
  • Cited by (41)

    • Consistent heat tolerance under starvation across seasonal morphs in Mycalesis mineus (Lepidoptera: Nymphalidae)

      2022, Comparative Biochemistry and Physiology -Part A : Molecular and Integrative Physiology
      Citation Excerpt :

      However, it should be noted that the endpoint at larval spitting of green fluid was taken simply for its universality and ease of detection. It could be possible that such fluid was linked to body water content that increased with body mass and food availability (but see Mitchell et al., 2017; Overgaard et al., 2012). Ecological relevance of this physiological endpoint will require further investigation.

    • Geographic variation in acclimation responses of thermal tolerance in South African diving beetles (Dytiscidae: Coleoptera)

      2021, Comparative Biochemistry and Physiology -Part A : Molecular and Integrative Physiology
      Citation Excerpt :

      In the lab we transferred them to glass aquariums (5 L of distilled water) with vegetation and stones and fed them with frozen Artemia. Estimates of thermal resistance are typically not confounded by starvation and dehydration stress during assays over the kinds of time-scales we consider here (Terblanche et al., 2011; Overgaard et al., 2012; Mitchell et al., 2017). We haphazardly assigned them to one of two groups for each population to be held at either 20 °C or 26 °C for 7–8 days, both temperatures well within the range that any of the species we study here experience in the field in the austral summer (following e.g., Hoffman and Watson, 1993; Klok and Chown, 2003).

    • Water deprivation drives intraspecific variability in lizard heat tolerance

      2020, Basic and Applied Ecology
      Citation Excerpt :

      For instance, in vinegar flies (Drosophilidae) tolerance to high temperatures decreased from humid to dry conditions despite that its heritability peaked in arid conditions (Bubliy, Kristensen, Kellermann, & Loeschcke, 2012). By contrast, considering that animals can obtain water via food resources and directly drinking, fruit flies (Tephritidae) deprived of access to water showed decreased heat knockdown time, but only if combined with food deprivation, and treatments did not affect the expression profiles of heat-shock proteins (Mitchell, Boardman, Clusella-Trullas, & Terblanche, 2017). Given that insects are highly sensitive to variation in rainfall and water availability, owing to their small body size that drives a high surface area to volume ratio, and thus high susceptibility to evaporative water loss and limited capacity to carry water (Chown, Sørensen, & Terblanche, 2011), it could be argued that these experiments should bear limited relevance for vertebrate ectotherms.

    View all citing articles on Scopus
    1

    Current address: Department of Entomology and Nematology, University of Florida, Gainesville, USA.

    View full text