Acessibilidade / Reportar erro

Multiple genes contribute to anhydrobiosis (tolerance to extreme desiccation) in the nematode Panagrolaimus superbus

Abstract

The molecular basis of anhydrobiosis, the state of suspended animation entered by some species during extreme desiccation, is still poorly understood despite a number of transcriptome and proteome studies. We therefore conducted functional screening by RNA interference (RNAi) for genes involved in anhydrobiosis in the holo-anhydrobiotic nematode Panagrolaimus superbus. A new method of survival analysis, based on staining, and proof-of-principle RNAi experiments confirmed a role for genes involved in oxidative stress tolerance, while a novel medium-scale RNAi workflow identified a further 40 anhydrobiosis-associated genes, including several involved in proteostasis, DNA repair and signal transduction pathways. This suggests that multiple genes contribute to anhydrobiosis in P. superbus.

Keywords:
peroxiredoxin; kinase; RNAi; proteostasis

Introduction

A few species of bacteria, yeasts, plants and small invertebrates are capable of surviving extreme desiccation through a unique and outstanding strategy: anhydrobiosis. When facing severe drought, these species begin to dehydrate and, instead of dying, they accumulate intrinsically disordered proteins (such as LEA and TDPs; Boothby et al., 2017Boothby TC, Tapia H, Brozena AH, Piszkiewicz S, Smith AE, Giovannini I, Rebecchi L, Pielak GJ, Koshland D and Goldstein B (2017) Tardigrades use intrinsically disordered proteins to survive desiccation. Mol Cell 65:975-984.) and non-reducing disaccharides (such as trehalose and sucrose), which promote the vitrification of the internal cellular environment. This process results in a bioglass – an amorphous organic scaffold that completely arrests metabolism and preserves internal contents (Crowe et al., 1998Crowe JH, Carpenter JF and Crowe LM (1998) The role of vitrification in anhydrobiosis. Annu Rev Physiol 60:73-103.). It is also proposed that such disaccharides act as `water-replacement molecules’, directly interacting with proteins and membranes and helping to maintain their native structures (Sakurai et al., 2008Sakurai M, Furuki T, Akao K, Tanaka D, Nakahara Y, Kikawada T, Watanabe M and Okuda T (2008) Vitrification is essential for anhydrobiosis in an African chironomid, Polypedilum vanderplanki. Proc Natl Acad Sci U S A 105:5093-5098.). However, it is possible that other factors also contribute to the protection of the organism. In this “dry state” (anhydrobiosis itself), the organism is tolerant to several other physical stresses such as extremes of temperature and pressure, ultraviolet light and radiation (Tunnacliffe and Lapinski, 2003Tunnacliffe A and Lapinski J (2003) Resurrecting Van Leeuwenhoek’s rotifers: A reappraisal of the role of disaccharides in anhydrobiosis. Philos Trans R Soc Lond B Biol Sci 358:1755-1771.; Watanabe et al., 2006aWatanabe M, Sakashita T, Fujita A, Kikawada T, Nakahara Y, Hamada N, Horikawa DD, Wada S, Funayama T, Kobayashi Y, et al. (2006a) Estimation of radiation tolerance to high LET heavy ions in an anhydrobiotic insect, Polypedilum vanderplanki. Int J Radiat Biol 82:835-842.,bWatanabe M, Sakashita T, Fujita A, Kikawada T, Horikawa DD, Nakahara Y, Wada S, Funayama T, Hamada N, Kobayashi Y, et al. (2006b) Biological effects of anhydrobiosis in an African chironomid, Polypedilum vanderplanki on radiation tolerance. Int J Radiat Biol 82:587-592.).

In rotifers, anhydrobiosis arrests the “biological clock”, meaning that they do not age when in suspended animation. Therefore, the average lifespan is unaltered by desiccation, regardless of the time spent in the dry state (Ricci et al., 1987Ricci C, Vaghi L and Manzini ML (1987) Desiccation of rotifers (Macrotrachela quadricornifera) - Survival and reproduction. Ecology 68:1488-1494.). However, once a dried rotifer is rehydrated, the same animal needs an interval of at least 24 h before being subjected to desiccation again to be able to survive (Schramm and Becker, 1987Schramm U and Becker W (1987) Anhydrobiosis of the bdelloid rotifer Habrotrocha rosa (Aschelminthes). Z Mikrosk Anat Forsch 101:1-17.). Notably, variations in the life histories of different anhydrobiotic species may occur (Ricci and Caprioli, 1998Ricci C and Caprioli, M (1998) Stress during dormancy: Effect on recovery rates and life-history traits of anhydrobiotic animals. Aquat Ecol 32:353-359.).

Anhydrobiotic organisms are exposed to extreme water stress, which causes deleterious effects in the cell, including oxidative damage (Leprince et al., 1994Leprince O, Atherton NM, Deltour R and Hendry GAF (1994) The involvement of respiration in free radical processes during loss of desiccation tolerance in germinating Zea mays L. (an electron paramagnetic resonance study). Plant Physiol 104:1333-1339.; França et al., 2007França MB, Panek AD and Eleutherio EC (2007) Oxidative stress and its effects during dehydration. Comp Biochem Physiol A Mol Integr Physiol 146:621-631.). Oxidative stress refers to a biological condition in which there is an imbalance in the concentrations of oxidant species and antioxidants (Sies, 2000Sies H (2000) What is oxidative stress? In: Keaney JF (ed) Oxidative Stress and Vascular Disease. Kluwer Academic Publishers, Dordrecht, pp 1-8.). Organisms have developed adaptive mechanisms for cellular detoxification, including systems that repair or prevent damage caused by oxidants (Michiels et al., 1994Michiels C, Raes M, Toussaint O and Remacle J (1994) Importance of Se-glutathione peroxidase, catalase, and Cu/Zn-SOD for cell survival against oxidative stress. Free Radic Biol Med 17:235-248.). Among such defense systems in the anhydrobiotic nematode Panagrolaimus superbus are: protein DJ-1 (Culleton et al., 2015Culleton BA, Lall P, Kinsella GK, Doyle S, McCaffrey J, Fitzpatrick DA and Burnell AM (2015) A role for the Parkinson’s disease protein DJ-1 as a chaperone and antioxidant in the anhydrobiotic nematode Panagrolaimus superbus. Cell Stress Chaperones 20:121-137.), glutathione peroxidases (GP114) and peroxiredoxin (PER).

Although anhydrobiosis in animals was first described more than three centuries ago (Van Leeuwenhoek, 1702Van Leeuwenhoek A (1702) Letter to Hendrik van Bleyswijk, dated 9 February 1702.), its molecular basis is poorly understood. Transcriptome and proteome analyses in tardigrades (Milnesium tardigradum, Richtersius coronifer and Hypsibius dujardini), bdelloid rotifers (Adineta ricciae), nematodes (Aphelenchus avenae, Ditylenchus africanus, Plectus murrayi and Panagrolaimus superbus), insects (Polypedilum vanderplanki), algae and plants (Pyropia orbicularis, Myrothamnus flabellifolia and Boea hygrometrica) identified several genes and proteins that were up- or down-regulated by water loss (Adhikari et al., 2009Adhikari BN, Wall DH and Adams BJ (2009) Desiccation survival in an Antarctic nematode: Molecular analysis using expressed sequenced tags. BMC Genomics 10:69.; Haegeman et al., 2009Haegeman A, Jacob J, Vanholme B, Kyndt T, Mitreva M and Gheysen G (2009) Expressed sequence tags of the peanut pod nematode Ditylenchus africanus: The first transcriptome analysis of an Anguinid nematode. Mol Biochem Parasitol 167:32-40.; Mali et al., 2010Mali B, Grohme MA, Förster F, Dandekar T, Schnölzer M, Reuter D, Welnicz W, Schill RO and Frohme M (2010) Transcriptome survey of the anhydrobiotic tardigrade Milnesium tardigradum in comparison with Hypsibius dujardini and Richtersius coronifer. BMC Genomics 11:168.; Schokraie et al., 2010Schokraie E, Hotz-Wagenblatt A, Warnken U, Mali B, Frohme M, Förster F, Dandekar T, Hengherr S, Schill RO and Schnölzer M (2010) Proteomic analysis of tardigrades: Towards a better understanding of molecular mechanisms by anhydrobiotic organisms. PLoS One 5:e9502., 2012Schokraie E, Warnken U, Hotz-Wagenblatt A, Grohme MA, Hengherr S, Förster F, Schill RO, Frohme M, Dandekar T and Schnölzer M (2012) Comparative proteome analysis of Milnesium tardigradum in early embryonic state versus adults in active and anhydrobiotic state. PLoS One 7:e45682.; Boschetti et al., 2011Boschetti C, Pouchkina-Stantcheva N, Hoffmann P and Tunnacliffe A (2011) Foreign genes and novel hydrophilic protein genes participate in the desiccation response of the bdelloid rotifer Adineta ricciae. J Exp Biol 214:59-68.; Tyson et al., 2012Tyson T, O’Mahony Zamora G, Wong S, Skelton M, Daly B, Jones JT, Mulvihill ED, Elsworth B, Phillips M, Blaxter M, et al. (2012) A molecular analysis of desiccation tolerance mechanisms in the anhydrobiotic nematode Panagrolaimus superbus using expressed sequenced tags. BMC Res Notes 5:68.; Yamaguchi et al., 2012Yamaguchi A, Tanaka S, Yamaguchi S, Kuwahara H, Takamura C, Imajoh-Ohmi S, Horikawa DD, Toyoda A, Katayama T, Arakawa K, et al. (2012) Two novel heat-soluble protein families abundantly expressed in an anhydrobiotic tardigrade. PLoS One 7:e44209., Wang et al., 2014Wang C, Grohme MA, Mali B, Schill RO and Frohme M (2014) Towards decrypting cryptobiosis - analyzing anhydrobiosis in the tardigrade Milnesium tardigradum using transcriptome sequencing. PLoS One 9:e92663.; López-Cristoffanini et al., 2015López-Cristoffanini C, Zapata J, Gaillard F, Potin P, Correa JA and Contreras-Porcia L (2015) Identification of proteins involved in desiccation tolerance in the red seaweed Pyropia orbicularis (Rhodophyta, Bangiales). Proteomics 15:3954-3968.; Ma et al., 2015Ma C, Wang H, Macnish AJ, Estrada-Melo AC, Lin J, Chang Y, Reid MS and Jiang CZ (2015) Transcriptomic analysis reveals numerous diverse protein kinases and transcription factors involved in desiccation tolerance in the resurrection plant Myrothamnus flabellifolia. Hortic Res 22:15034.; Zhu et al., 2015Zhu Y, Wang B, Phillips J, Zhang ZN, Du H, Xu T, Huang LC, Zhang XF, Xu GH, Li WL, et al. (2015) Global transcriptome analysis reveals acclimation-primed processes involved in the acquisition of desiccation tolerance in Boea hygrometrica. Plant Cell Physiol 56:1429-1441.; Ryabova et al., 2017Ryabova A, Mukae K, Cherkasov A, Cornette R, Shagimardanova E, Sakashita T, Okuda T, Kikawada T and Gusev O (2017) Genetic background of enhanced radioresistance in an anhydrobiotic insect: Transcriptional response to ionizing radiations and desiccation. Extremophiles 21:109-120.). Such studies are, by their nature, correlative, and do not provide evidence for a functional role of the genes or proteins concerned. In the model nematode, Caenorhabditis elegans, whose dauer larvae alone are desiccation tolerant, a large number of mutants are available that can be used to study anhydrobiosis in this species (Erkut et al., 2011Erkut C, Penkov S, Khesbak H, Vorkel D, Verbavatz JM, Fahmy K and Kurzchalia TV (2011) Trehalose renders the dauer larva of Caenorhabditis elegans resistant to extreme desiccation. Curr Biol 21:1331-1336., 2013Erkut C, Vasilj A, Boland S, Habermann B, Shevchenko A and Kurzchalia TV (2013) Molecular strategies of the Caenorhabditis elegans dauer larva to survive extreme desiccation. PLoS One 8:e82473.). However, in other nematodes, the lack of such mutants has prompted researchers to use RNA interference (RNAi) techniques instead (Reardon et al., 2010Reardon W, Chakrabortee S, Pereira TC, Tyson T, Banton MC, Dolan KM, Culleton BA, Wise MJ, Burnell AM and Tunnacliffe A (2010) Expression profiling and cross-species RNA interference (RNAi) of desiccation-induced transcripts in the anhydrobiotic nematode Aphelenchus avenae. BMC Mol Biol 11:6.), an approach which is also possible in C. elegans (Gal et al., 2004Gal TZ, Glazer I and Koltai H (2004) An LEA group 3 family member is involved in survival of C. elegans during exposure to stress. FEBS Lett 577:21-26.; Erkut et al., 2013Erkut C, Vasilj A, Boland S, Habermann B, Shevchenko A and Kurzchalia TV (2013) Molecular strategies of the Caenorhabditis elegans dauer larva to survive extreme desiccation. PLoS One 8:e82473.).

Identification of anhydrobiosis-related genes is a central requirement in the development of anhydrobiotic engineering, which aims to confer desiccation tolerance on dehydration-sensitive biological samples (cells, tissues, organs) (Chen et al., 2009Chen W, Zhang X, Liu M, Zhang J, Ye Y, Lin Y, Luyckx J and Qu J (2009) Trehalose protects against ocular surface disorders in experimental murine dry eye through suppression of apoptosis. Exp Eye Res 89:311-318.; García De Castro et al., 2000García De Castro A, Bredholt H, Strøm AR and Tunnacliffe A (2000) Anhydrobiotic engineering of gram-negative bacteria. Appl Environ Microbiol 66:4142-4124.; Li et al., 2012Li S, Chakraborty N, Borcar A, Menze MA, Toner M and Hand SC (2012) Late embryogenesis abundant proteins protect human hepatoma cells during acute desiccation. Proc Natl Acad Sci U S A 109:20859-20864.). Successful anhydrobiotic engineering would have multiple applications in agriculture (e.g., by rendering plants tolerant to drought) and medicine (e.g., preservation in the dry state of organs for transplant).

In this study we examined anhydrobiosis in P. superbus, a free-living nematode nearly 1 mm long that feeds on bacteria and was first described by Fuchs (1930)Fuchs G (1930) Neue an Borkenkäfer und Russelkäfer gebundene Nematoden, halbparasitische und Wohnungseinmieter. Zool Jb 59:586-608.. Members of the genus Panagrolaimus inhabit diverse niches, from the Antarctic, volcanic islands, temperate and semi-arid soils to terrestrial mosses (Shannon et al., 2005Shannon AJ, Browne JA, Boyd J, Fitzpatrick DA and Burnell AM (2005) The anhydrobiotic potential and molecular phylogenetics of species and strains of Panagrolaimus (Nematoda, Panagrolaimidae). J Exp Biol 208:2433-2445.; McGill et al., 2015McGill LM, Shannon AJ, Pisani D, Félix MA, Ramløv H, Dix I, Wharton DA and Burnell AM (2015) Anhydrobiosis and freezing-tolerance: Adaptations that facilitate the establishment of Panagrolaimus nematodes in polar habitats. PLoS One 10:e0116084.). P. superbus and C. elegans belong to the same order (Rhabditida) and are anatomically similar. However, the former is dioecious while the latter is typically hermaphroditic and has a faster populational growth rate. We have focused on P. superbus, rather than C. elegans, because the former nematode is: (i) holo-anhydrobiotic (Jönsson, 2005Jönsson KI (2005) The evolution of life histories in holo-anhydrobiotic animals: A first approach. Integr Comp Biol 45:764-770.) i.e., able to enter anhydrobiosis at any life stage, (ii) robustly desiccation tolerant (Shannon et al., 2005Shannon AJ, Browne JA, Boyd J, Fitzpatrick DA and Burnell AM (2005) The anhydrobiotic potential and molecular phylogenetics of species and strains of Panagrolaimus (Nematoda, Panagrolaimidae). J Exp Biol 208:2433-2445.) and (iii) does not demand extra/special laboratory procedures to obtain specific larval stages.

We first developed a new method for rapid and accurate assessment of survival to desiccation that could be used in a scalable screening procedure. To test this method and also to gain further information on the functional roles of glutathione peroxidase (a protein previously shown to be involved in anhydrobiosis; Reardon et al., 2010Reardon W, Chakrabortee S, Pereira TC, Tyson T, Banton MC, Dolan KM, Culleton BA, Wise MJ, Burnell AM and Tunnacliffe A (2010) Expression profiling and cross-species RNA interference (RNAi) of desiccation-induced transcripts in the anhydrobiotic nematode Aphelenchus avenae. BMC Mol Biol 11:6.) and peroxiredoxin (a biochemically related protein), we performed RNA interference by feeding on populations of P. superbus. We then developed a new medium-scale RNAi screening protocol to screen a panel of 97 target genes previously shown to be regulated during extreme desiccation in other anhydrobiotic species and found that knockdown of 40 of these genes adversely affects desiccation tolerance in P. superbus.

Materials and Methods

Nematode maintenance

Panagrolaimus superbus (strain DF5050) used in this study was first isolated from Surtsey Island (Iceland) by Björn Sohlenius (Sohlenius, 1988Sohlenius B (1988) Interactions between two species of Panagrolaimus in agar cultures. Nematologica 34:208-217.; Shannon et al., 2005Shannon AJ, Browne JA, Boyd J, Fitzpatrick DA and Burnell AM (2005) The anhydrobiotic potential and molecular phylogenetics of species and strains of Panagrolaimus (Nematoda, Panagrolaimidae). J Exp Biol 208:2433-2445.) and was maintained in incubators at 21 °C, in the dark, on NGM (Nematode Growth Medium) agar plates and fed with a layer of Escherichia coli (strain OP50).

Evaluation of staining for the determination of survival percentages

P. superbus worms were collected from maintenance plates (NGM agar covered with a layer of OP50) by rinsing with 5 mL of M9 buffer, transferred to 50 mL test tubes and left for 10 min to precipitate. Then, the supernatant was discarded and worms were transferred to new 50 mL tubes containing 10 mL of M9 buffer. This washing procedure was done three times to reduce the amount of OP50 bacteria. Worms were then separated into two 0.2 mL vials. One sample was heated at 70 °C for 10 min in a thermocycler with subsequent decanting for 10 min. Heating at this temperature is lethal to worms, providing a positive control for staining. The second sample was kept at rest for 20 min at room temperature.

Supernatant was then removed and 180 μL trypan blue or erythrosin B (both 0.4% w/v in M9 buffer) were added. Tubes were left at room temperature for four hours without agitation. After this period, supernatant was removed and 150 μL of M9 buffer were added; the total volume was transferred to 1.5 mL tubes, to which 850 μL of M9 was added. After gentle agitation, tubes were left without agitation for 10 min for worm decantation. This washing procedure (disposing supernatant and adding 950 μL of M9 buffer) was repeated three times to remove the dye. Subsequently, worms were placed on plastic plates.

For each treatment, three plates were produced: i) a sample of live worms, ii) a sample of dead worms, and iii) a mixture of live and dead worms. These plates were analyzed via light microscopy and images were captured using the program ScopePhoto to verify whether stained worms (bluish by trypan blue, pink by erythrosin B) were active, thus revealing any false positives. Similarly, a total and permanent absence of movements in unstained (live worms), representing false negatives, was also considered. The whole procedure was also performed for C. elegans worms. Staining for only one hour was also tested. Three biological replicates were performed (N > 100 worms per group, per replicate).

RNAi by feeding - PER and GP114

Partial cDNAs corresponding to a glutathione peroxidase (designated GP114, GenBank Accession Number GR881191) and peroxiredoxin (designated PER, GenBank Accession Number GR881190) in the L4440 vector (ampicillin resistance) were propagated in E. coli HT115 (Reardon et al., 2010Reardon W, Chakrabortee S, Pereira TC, Tyson T, Banton MC, Dolan KM, Culleton BA, Wise MJ, Burnell AM and Tunnacliffe A (2010) Expression profiling and cross-species RNA interference (RNAi) of desiccation-induced transcripts in the anhydrobiotic nematode Aphelenchus avenae. BMC Mol Biol 11:6.). These feeding strains, designated dsGP114 and dsPER, were grown in 50 mL tubes of liquid LB ampicillin (50 μg/mL) under agitation (210 rpm, 37 °C) overnight. Subsequently, tubes were centrifuged for 10 min at 3,500 x g. Pellets were resuspended with 600 μL of liquid LB ampicillin and inverted on petri dishes containing NGM agar and IPTG (1 mM). Plates were left for two days at room temperature to induce double-stranded RNA (dsRNA) expression by the bacteria.

Worms were then collected from OP50 and washed with M9, as previously described. Subsequently, a small amount of worms was transferred to either plates containing bacteria expressing dsRNA against GP114 or PER. They were left for nearly 15 days to ensure silencing of the entire population. HT115 bacteria containing a GFP gene cloned in the L4440 vector (L4440::GFP, referred to as “GFP”) was used as a negative control. Populations fed with L4440::GP114 bacteria are referred to as “dsGP114”, “GP114 knockdown” or “GP114-silenced”; similarly for PER, mutatis mutandis. The terms “GP114” and “PER” were used to refer to the corresponding genes/cDNAs. Three biological replicates were performed (N = 200 worms for each treatment, for each replicate).

Desiccation challenge

Worms were submitted to desiccation challenge according to Shannon et al. (2005)Shannon AJ, Browne JA, Boyd J, Fitzpatrick DA and Burnell AM (2005) The anhydrobiotic potential and molecular phylogenetics of species and strains of Panagrolaimus (Nematoda, Panagrolaimidae). J Exp Biol 208:2433-2445.. Briefly, silenced worms were immobilized on 0.45 μm Supor filter membranes (Sigma Aldrich) by vacuum filtration with a Sartorius funnel, placed in 1.5 mL test tubes and then subjected to the following conditions: 98% relative humidity (RH) for 24 h over a saturated solution of copper sulphate (unless otherwise stated); 10% RH for 24 h over dry silica gel and pre-hydration in 100% RH for 24 h in distilled water vapour. Rehydration was achieved by adding 1.5 mL of M9 buffer to the samples. Survival percentage was measured by staining with erythrosin B. Three biological replicates were performed (N > 100 worms per group, per replicate).

Assessing the roles of PER and GP114 as antioxidants

PER- and GP114-silenced worms (in M9 buffer) were subjected to oxidative stress by adding hydrogen peroxide (H2O2, Synth) to the following final concentrations: 0 μM (zero), 1 μM, 10 μM, 100 μM, 1 mM, 10 mM, 20 mM and 40 mM. The final volume in all tubes was 100 μL. These values were selected according to previous studies on C. elegans (Larsen, 1993Larsen PL (1993) Aging and resistance to oxidative damage in Caenorhabditis elegans. Proc Natl Acad Sci U S A 90:8905-8909.).

Samples were then homogenized by mild agitation and incubated at 20 °C for 24 h. After this period, the supernatant was removed and 1 mL of erythrosin B (0.4% w/v) was added and left for four hours. Worms were then washed three times with M9 buffer and survival percentages were determined (number of unstained worms/total number of worms). Three biological replicates were performed (N = 200 per group, per replicate).

Screening for anhydrobiosis-related genes in P. superbus

Selection of targets

A total of 97 potential targets were considered for the screening experiments. The first group comprised 33 kinase-related cDNAs, obtained from a mixed population of P. superbus and cloned in the pDNR-Lib vector (Clontech), kindly provided by Dr. Trevor Tyson (Van Andel Institute, USA). They correspond to all genes whose “target codes” end with a “K”, in Table 1. These targets were selected because signaling processes are likely to be very important for entry into anhydrobiosis. The second group (all other genes in Table 1) comprised 64 genes shown to be up-regulated during anhydrobiosis in other animal species (Adhikari et al., 2009Adhikari BN, Wall DH and Adams BJ (2009) Desiccation survival in an Antarctic nematode: Molecular analysis using expressed sequenced tags. BMC Genomics 10:69.; Haegeman et al., 2009Haegeman A, Jacob J, Vanholme B, Kyndt T, Mitreva M and Gheysen G (2009) Expressed sequence tags of the peanut pod nematode Ditylenchus africanus: The first transcriptome analysis of an Anguinid nematode. Mol Biochem Parasitol 167:32-40.; Mali et al., 2010Mali B, Grohme MA, Förster F, Dandekar T, Schnölzer M, Reuter D, Welnicz W, Schill RO and Frohme M (2010) Transcriptome survey of the anhydrobiotic tardigrade Milnesium tardigradum in comparison with Hypsibius dujardini and Richtersius coronifer. BMC Genomics 11:168.; Schokraie et al., 2010Schokraie E, Hotz-Wagenblatt A, Warnken U, Mali B, Frohme M, Förster F, Dandekar T, Hengherr S, Schill RO and Schnölzer M (2010) Proteomic analysis of tardigrades: Towards a better understanding of molecular mechanisms by anhydrobiotic organisms. PLoS One 5:e9502., 2012Schokraie E, Warnken U, Hotz-Wagenblatt A, Grohme MA, Hengherr S, Förster F, Schill RO, Frohme M, Dandekar T and Schnölzer M (2012) Comparative proteome analysis of Milnesium tardigradum in early embryonic state versus adults in active and anhydrobiotic state. PLoS One 7:e45682.; Cornette et al., 2010Cornette R, Kanamori Y, Watanabe M, Nakahara Y, Gusev O, Mitsumasu K, Kadono-Okuda K, Shimomura M, Mita K, Kikawada T, et al. (2010) Identification of anhydrobiosis-related genes from an expressed sequence tag database in the cryptobiotic midge Polypedilum vanderplanki (Diptera; Chironomidae). J Biol Chem 285:35889-35999.; Boschetti et al., 2012Boschetti C, Carr A, Crisp A, Eyres I, Wang-Koh Y, Lubzens E, Barraclough TG, Micklem G and Tunnacliffe A (2012) Biochemical diversification through foreign gene expression in bdelloid rotifers. PLoS Genet 8:e1003035.; Tyson et al., 2012Tyson T, O’Mahony Zamora G, Wong S, Skelton M, Daly B, Jones JT, Mulvihill ED, Elsworth B, Phillips M, Blaxter M, et al. (2012) A molecular analysis of desiccation tolerance mechanisms in the anhydrobiotic nematode Panagrolaimus superbus using expressed sequenced tags. BMC Res Notes 5:68.; Yamaguchi et al., 2012Yamaguchi A, Tanaka S, Yamaguchi S, Kuwahara H, Takamura C, Imajoh-Ohmi S, Horikawa DD, Toyoda A, Katayama T, Arakawa K, et al. (2012) Two novel heat-soluble protein families abundantly expressed in an anhydrobiotic tardigrade. PLoS One 7:e44209.). These targets were selected by considering the following aspects: (i) they should be induced in at least one species during anhydrobiosis and (ii) there should be homolog(s) within the P. superbus EST library (Tyson et al., 2012Tyson T, O’Mahony Zamora G, Wong S, Skelton M, Daly B, Jones JT, Mulvihill ED, Elsworth B, Phillips M, Blaxter M, et al. (2012) A molecular analysis of desiccation tolerance mechanisms in the anhydrobiotic nematode Panagrolaimus superbus using expressed sequenced tags. BMC Res Notes 5:68.).

Table 1
Functional identification of anhydrobiosis-related genes via RNAi. All the 40 genes whose knockdown lead to statistically significant decreases in survival percentage ≤10% compared to control group (i.e., worms soaked with RNA duplexes against GFP, normalized as 100% survival) are listed here in gray. Nine genes (marked with asterisks) presented statistically significant decreases lower than 10% (One-Way ANOVA). The remaining 48 targets that did not lead to statistically significant reductions in the initial screenings and are listed in white.

Production of long double-stranded RNAs (dsRNAs)

dsRNAs were generated for all 33 kinase-related cDNAs. Briefly, 500 ng target cDNAs (cloned in pDNR-Lib vector) were subjected to PCR using primers which correspond to flanking sequences in the vector and include a T7 promoter tail (Table S1). PCR was performed with GoTaq DNA Polymerase (Promega) in 50 μL under the following conditions: 94 °C for 5 min , followed by 33 cycles of 94 °C for 30 s, 53 °C for 30 s, 72 °C for 1 min, and an extension step at 72 °C for 10 min. The resulting amplicons were precipitated with isopropanol, resuspended with ultrapure water and submitted to in vitro transcription (Yu et al., 2002Yu JY, DeRuiter SL and Turner DL (2002) RNA interference by expression of short-interfering RNAs and hairpin RNAs in mammalian cells. Proc Natl Acad Sci U S A 99:6047-6052.) using TranscriptAid T7 High yield transcription kit; ThermoScientific, followed by DNase I treatment, according to the manufacturer’s instructions. dsRNAs were diluted to 1 μg/μL with ultrapure water and Tris HCl (pH 6.8) was added to a final concentration of 5 mM.

dsRNAs were also generated for 14 other genes from P. superbus (Table S1, target codes 1-4, 6-11, 15-17 and 21). Initially, their corresponding cDNAs were obtained by RT-PCR (using the same conditions described in section 2.6.5) and cloned into vector pCR2.1 TOPO (Invitrogen). These cloned sequences were then used as templates for a second round of PCR (performed as for pDNR-Lib, described above), but now using T7-gene specific primers (sequences on Table S1; targets 1 to 21). The resulting amplicons could be readily used for in vitro transcription (as previously described), yielding dsRNAs.

Design of dicer substrates

Dicer substrates were designed for 50 targets (genes whose target codes start with “si”, on Table S2), using the freeware Strand Analysis (Pereira et al., 2007Pereira TC, Pascoal VDB, Secolin R, Rocha CS, Maia IG and Lopes-Cendes I (2007) Strand Analysis, a free online program for the computational identification of the best RNA interference (RNAi) targets based on Gibbs free energy. Genet Mol Biol 30:1206-1208.) and extended three nucleotides at each end. These molecules are 27 RNA duplexes, with two nucleotide 3′ overhangs and phosphate groups at the 5′ ends. As a negative control, we designed a dicer substrate against GFP (accession number X83960). Dicer substrates were purchased from Sigma-Aldrich; their sequences are listed in Table S2.

RNA interference by soaking with siRNAs/dsRNAs

RNAi was triggered by soaking 200 - 600 worms (per biological replicate per target) for 24 h, in the dark, with long dsRNAs at a final concentration of 0.8 μg/μL (soaking volume: 35 μL) or dicer substrates (siRNAs) at final concentration of 1 μM (soaking volume: 100 μL) and kept in the dark for 24 h without agitation at 21 °C. Three biological replicates were performed.

Confirmation of gene silencing by semi-quantitative RT-PCR

We selected a few representative targets to perform semi-quantitative RT-PCR to evaluate gene silencing. We also assessed a representative gene shown not to be involved in anhydrobiosis (si86 – whose silencing did not lead to decrease in survival) to show that a lack of decrease in survival after desiccation is not due to ineffective gene silencing.

Initially, total RNA from worms was extracted using the TRIzol reagent (Invitrogen) according to the manufacturer’s guidelines. RNA samples were quantified by spectrophotometry and subsequently diluted in ultra pure water (RNase free) to yield a final concentration of 1 μg/μL. All RNA samples were pre-treated with DNase I (Fermentas) following a modified version of the manufacturer’s protocol: one unit of enzyme (1 h at 37 °C), followed by addition of another unit of enzyme (1 h at 37 °C). Reverse transcription reactions (RT) were performed using ImProm-IITM kit (Promega) and random primers (500 ng) in a final volume of 20 μL, according to the manufacturer. PCR was then performed using the GoTaqR DNA Polymerase kit (Promega) according to the manufacturer’s instructions. PCR was performed using 2 μL of RT and 25 picomoles of each gene-specific primer (forward or reverse) or β-actin (separate tubes) in a final volume of 50 μL.

All PCR reactions were performed under the following conditions: 94 °C for 5 min, followed by 33 cycles of 94 °C for 30 s, 57 °C for 30 s, 72 °C for 1 min, and a final extension at 72 °C for 10 min. Amplifications curves were analyzed to guarantee that a plateau was not reached under these conditions for the tested genes. PCR products were resolved in 1% agarose gel stained with Sybr Safe (Invitrogen). Band densitometry was done using IMAGE J software. Data normalization was done by dividing the value obtained for the silenced gene by the value found for β-actin for the corresponding sample. For each target, semi-quantitative RT-PCR was performed in technical triplicates, each one consisting of a pool of nearly 600 worms.

Nucleotide sequences of each primer are listed in Table S1. PCR conditions were the same for all genes analyzed, except for target si86: 23 cycles.

Lethality assay

In order to determine whether the knockdown alone causes a decrease in viability, worms were soaked for 24 h with RNA duplexes for all 97 targets (or fed with dsPER/dsGP114) and survival percentages were determined by staining. This procedure aims to guarantee that any decrease in survival percentage (compared to control group) is due to the disruption of the process of anhydrobiosis rather than an unrelated lethality. Three biological replicates were performed (N = 200 per group, per replicate).

Statistical analyses

All experiments were performed in biological triplicates (or quadruplicates) and data are presented as mean values and standard deviations. Statistical analyses were performed using Student’s t-test, z-test or one-way ANOVA (with Tukey’s or Dunn’s post-hoc tests) or Mann-Whitney Rank Sum test with SigmaStat software. Statistical differences were considered significant when p ≤ 0.05. Only those genes giving a decrease in survival percentage >10% after RNAi/desiccation challenge were considered “anhydrobiosis-related”.

Results

Evaluation of staining for the determination of survival percentages

Both tested compounds (erythrosin B and trypan blue) are commonly used for staining cells, and were successfully used to indicate viability in whole worms (Figure S1 A-F). Heat-killed worms were strongly and completely stained, while live animals remained unstained even after four hours of soaking in dye solution. In a very few cases, faint local staining was observed in live nematodes, probably indicating local tissue damage.

After desiccation challenge, all developmental stages (eggs, larvae and adult worms) were stained when dead. No false negatives were observed. However, staining patterns varied: in many cases we observed intense whole-body staining, but in some cases partial staining (of the anterior, middle or posterior body regions) were seen in moving worms (false positives). Light staining occurred in some larvae and adults, without movement, which we considered dead. Staining for four hours or one hour was equally effective. It is possible that desiccation might make membranes leaky without killing worms (generating the few observed false positives). However, we believe that the observed “whole-body staining” reflects a degree of membrane damage that is incompatible with life, and thus, these worms can be considered dead. Additionally, such categorisation was applied to all groups (experimental and control), allowing an unbiased analysis.

Involvement of a peroxiredoxin gene in P. superbus anhydrobiosis

Previously, a glutathione peroxidase gene (GP114) was shown by RNAi to be involved in P. superbus anhydrobiosis (Reardon et al., 2010Reardon W, Chakrabortee S, Pereira TC, Tyson T, Banton MC, Dolan KM, Culleton BA, Wise MJ, Burnell AM and Tunnacliffe A (2010) Expression profiling and cross-species RNA interference (RNAi) of desiccation-induced transcripts in the anhydrobiotic nematode Aphelenchus avenae. BMC Mol Biol 11:6.). To test whether the staining method could demonstrate the effect of gene silencing on survival after desiccation, we first fed nematodes bacteria expressing GP114 dsRNA. In addition, we also used RNAi to examine the role of another P. superbus antioxidant gene, encoding a peroxiredoxin (PER), in anhydrobiosis.

Gene silencing was confirmed by semi-quantitative RT-PCR, revealing an average reduction in mRNA levels of 71% for PER and 61% for GP114 transcripts, compared to a control group (Figure 1A). Prior to desiccation, gene silencing had no effect on nematode viability (Figure 1B), demonstrating that PER and GP114 are not essential genes under the tested conditions. After desiccation, all groups displayed increased mortality (Figure 1C), but PER and GP114 groups were significantly more sensitive than the GFP control. In particular, PER knockdown resulted in a 66% decrease in viability compared to the control. As well as confirming a role for glutathione peroxidase, these data also show for the first time the participation of a peroxiredoxin (PER) in nematode anhydrobiosis. An EST-based study on P. superbus identified two clusters for peroxiredoxin genes and three for glutathione peroxidase (Tyson et al., 2012Tyson T, O’Mahony Zamora G, Wong S, Skelton M, Daly B, Jones JT, Mulvihill ED, Elsworth B, Phillips M, Blaxter M, et al. (2012) A molecular analysis of desiccation tolerance mechanisms in the anhydrobiotic nematode Panagrolaimus superbus using expressed sequenced tags. BMC Res Notes 5:68.). Therefore, it is possible that the lower survival percentage observed when silencing peroxiredoxin, compared to glutathione peroxidase, is due to the lower compensation capacity within the first protein family. These findings are consistent with oxidative stress being a significant component of the various stress vectors experienced by desiccating nematodes, as indicated for other organisms (Haegeman et al., 2009Haegeman A, Jacob J, Vanholme B, Kyndt T, Mitreva M and Gheysen G (2009) Expressed sequence tags of the peanut pod nematode Ditylenchus africanus: The first transcriptome analysis of an Anguinid nematode. Mol Biochem Parasitol 167:32-40.; Cornette et al., 2010Cornette R, Kanamori Y, Watanabe M, Nakahara Y, Gusev O, Mitsumasu K, Kadono-Okuda K, Shimomura M, Mita K, Kikawada T, et al. (2010) Identification of anhydrobiosis-related genes from an expressed sequence tag database in the cryptobiotic midge Polypedilum vanderplanki (Diptera; Chironomidae). J Biol Chem 285:35889-35999.). This experiment also validates the staining method for the assessment of survival of desiccation.

Figure 1
Involvement of peroxiredoxin and glutathione peroxidase in anhydrobiosis in P. superbus. A) Molecular analysis by semi-quantitative RT-PCR revealed an average reduction of 71% of PER RNA transcripts and 61% of GP114 in worms subjected to RNAi by feeding compared to the control group corresponding to worms fed with bacteria expressing dsRNA against GFP gene, which is not associated with anhydrobiosis (* p ≤ 0.05, one-way ANOVA). B) Viability tests before desiccation. Average survival percentages obtained for different treatments. Groups did not statistically differ and remained above 95% survival. C) Survival tests after extreme desiccation. Silencing PER promoted a 66% reduction in survival percentage when compared to control group; GP114 promoted a 21% reduction (* p≤0.05; one-way ANOVA).

PER and GP114 act as antioxidants

Knockdown of PER and GP114 had little or no effect on the morphology, development, fertility and behaviour of P. superbus (data not shown), as anticipated. However, we would expect gene silencing to compromise the ability of nematodes to combat oxidative stress, and we therefore tested this by exposing control and experimental groups to increasing concentrations of hydrogen peroxide for 24 h.

Slightly fluctuating responses (around the value observed for 0 μM) are noted for both treatments (PER and GP114) up to 10 mM and statistically significant differences (between control and respective experimental group) may be observed from 10 μM. Curiously, although a decrease in survival can be observed at 10 and 20 mM for PER, it is not seen at 40 mM, possibly due to compensation by other members of this gene family at higher concentrations (Figure 2A). On the other hand, GP114 also shows a decrease between experimental group and respective control at 20 mM, which is still present at 40 mM (Figure 2B), suggesting that a similar compensatory mechanism is not present, or is less effective. Taken together, these findings suggest that both PER and GP114 are involved in controlling oxidative stress, a situation known to occur during dehydration.

Figure 2
PER and GP114 enzymes combat oxidative stress. Statistically significant decreases in survival percentages were observed in both treatments (glutathione peroxidase- and peroxiredoxin-silenced worms) exposed to hydrogen peroxide. (* p<0.05; t-test between control group (GFP) and corresponding experimental group).

Evaluation of soaking as a means of triggering RNAi in P. superbus

We decided to determine whether immersing P. superbus in solutions containing long (>100 bp) and short (27 bp) RNA duplexes was also effective in promoting knockdown. We tested a few representative targets to show that, as judged by semi-quantitative RT-PCR, successful gene silencing was achieved by soaking with 27 bp RNA duplexes (known as dicer substrates) at 1 μM for 24 h (Figure 3 and Figure S2). Soaking in solutions of RNA duplexes resulted in gene silencing in a dose-dependent manner (up to 1 μM), although nonspecific effects began to emerge at high concentrations (10 μM, data not shown). Successful RNAi using dicer substrates was confirmed by its effect on target mRNA levels and by phenotypical analyses (knockdown of ifb-1 and actin genes, data not shown).

Figure 3
Molecular confirmation of gene knockdown by semi-quantitative RT-PCR. Three targets were selected to confirm gene silencing by soaking with dicer substrates, two of which lead to a decrease in survival after desiccation (si46 and si76) and one of which did not (si86). * p<0.05 (Tukey, compared to control group).

Functional identification of anhydrobiosis-related genes in P. superbus

We combined the use of the staining method for assessment of nematode viability with the soaking method for induction of RNAi in a medium-scale screening experiment to identify P. superbus genes associated with anhydrobiosis. We selected a panel of 97 genes, from which 40 genes (Figure 4 and Table 1) showed reduced survival (20% on average) after knockdown and desiccation challenge. This level of decrease has been previously observed in other nematodes (Reardon et al., 2010Reardon W, Chakrabortee S, Pereira TC, Tyson T, Banton MC, Dolan KM, Culleton BA, Wise MJ, Burnell AM and Tunnacliffe A (2010) Expression profiling and cross-species RNA interference (RNAi) of desiccation-induced transcripts in the anhydrobiotic nematode Aphelenchus avenae. BMC Mol Biol 11:6.; Erkut et al., 2013Erkut C, Vasilj A, Boland S, Habermann B, Shevchenko A and Kurzchalia TV (2013) Molecular strategies of the Caenorhabditis elegans dauer larva to survive extreme desiccation. PLoS One 8:e82473.) and revealed to be consistent with involvement in anhydrobiosis.

Figure 4
Identification of anhydrobiosis-related genes in P. superbus. All the 40 genes here indicated lead to statistically significant reductions >10% in survival percentage (z-test, p<0.05 compared to control group) after knockdown and desiccation.

Approximately half of the positive targets (24 genes) were related to cell signalling (kinase domain-containing proteins). The remaining sixteen genes encode several classes of proteins, including proteases, ribosomal proteins, structural proteins, aquaporins, DNA repair enzymes and molecular chaperones.

Discussion

P. superbus is a well-studied anhydrobiotic nematode that is amenable to RNAi (Shannon et al., 2008Shannon AJ, Tyson T, Dix I, Boyd J and Burnell AM (2008) Systemic RNAi mediated gene silencing in the anhydrobiotic nematode Panagrolaimus superbus. BMC Mol Biol 9:58.) and therefore a gene silencing approach can be taken to test the involvement of candidate genes in anhydrobiosis. However, to achieve this on anything more than a small scale requires some improvement to screening procedures, particularly in assessing survival of desiccation. Prior to the work in this report, desiccation tolerance was determined by observation under the microscope of “movement” as a survival criterion. However, this has its drawbacks, since absence of mobility is not necessarily evidence of death, and therefore the surviving fraction may be underestimated; it is also time- and labor-intensive. We describe here an alternative staining method, which is simple, fast, cheap and provides unequivocal survival data; it is particularly suited to medium- or large-scale screening experiments. Of the two stains tested, erythrosin B may be slightly preferable, as it has been used as a food colorant for many years and therefore carries fewer safety concerns.

The developed approach could be used to deepen our understanding on the functional roles of two P. superbus genes, involved in oxidative stress, in the process of desiccation tolerance. GP114 has previously been validated as an anhydrobiosis-related gene using RNAi (Reardon et al., 2010Reardon W, Chakrabortee S, Pereira TC, Tyson T, Banton MC, Dolan KM, Culleton BA, Wise MJ, Burnell AM and Tunnacliffe A (2010) Expression profiling and cross-species RNA interference (RNAi) of desiccation-induced transcripts in the anhydrobiotic nematode Aphelenchus avenae. BMC Mol Biol 11:6.); we confirmed this here and also showed the involvement of PER, demonstrating a role for peroxiredoxins in anhydrobiosis for the first time. As their name suggests, peroxiredoxins reduce hydrogen peroxide levels and thus help to limit oxidative damage caused by water loss, including lipid peroxidation, protein oxidation and DNA mutations, which may otherwise compromise cell function, eventually culminating in cell death (Hansen et al., 2006Hansen JM, Go YM and Jones DP (2006) Nuclear and mitochondrial compartmentation of oxidative stress and redox signaling. Annu Rev Pharmacol Toxicol 46:215-234.). The decrease in viability (~60%, compared to control group) after desiccation of PER-silenced nematodes reveals that peroxiredoxin activity is important for successful anhydrobiosis in P. superbus. However, the complexity of this phenomenon means that many other processes must be involved, and many studies have highlighted the importance of non-reducing disaccharide accumulation, LEA proteins, heat shock proteins and other molecular adaptations (see Erkut et al., 2013Erkut C, Vasilj A, Boland S, Habermann B, Shevchenko A and Kurzchalia TV (2013) Molecular strategies of the Caenorhabditis elegans dauer larva to survive extreme desiccation. PLoS One 8:e82473. for a systems approach in C. elegans and Burnell and Tunnacliffe (2011)Burnell AM and Tunnacliffe A (2011) Gene induction and desiccation stress in nematodes. In: Perry RN and Wharton (eds) Molecular and Physiological Basis of Nematode Survival. CABI Press, Wallingford, pp 126-156. for an earlier review of nematode anhydrobiosis).

Once we had validated the staining protocol by showing the involvement of GP114 and PER in P. superbus anhydrobiosis, our next step was to develop a practical approach to screen a larger set of genes. Therefore, we decided to evaluate the efficiency of RNAi by soaking in this species, since it is faster than feeding methods (24 h, instead of several days on feeding plates), uses less space (0.2 - 1.5 mL tubes, instead of 60 - 90 mm plates), demands fewer consumables and does not require cloning cDNAs in special feeding vectors. Original studies in C. elegans showed that SID-1, a transmembrane protein expressed in the pharynx, is responsible for the uptake of long dsRNA (Feinberg and Hunter, 2003Feinberg EH and Hunter CP (2003) Transport of dsRNA into cells by the transmembrane protein SID-1. Science 301:1545-1547.) and short siRNAs at high concentrations (Issa et al., 2005Issa Z, Grant WN, Stasiuk S and Shoemaker CB (2005) Development of methods for RNA interference in the sheep gastrointestinal parasite, Trichostrongylus colubriformis. Int J Parasitol 35:935-940.; Shih et al., 2009Shih JD, Fitzgerald MC, Sutherlin M and Hunter CP (2009) The SID-1 double-stranded RNA transporter is not selective for dsRNA length. RNA15:384-390.). The observation of nonspecific effects at extremely high concentrations of dicer substrates (10 μM) is probably due to: (i) interference in endogenous pathways, such as microRNA biosynthesis (Grimm et al., 2006Grimm D, Streetz KL, Jopling CL, Storm TA, Pandey K, Davis CR, Marion P, Salazar F and Kay MA (2006) Fatality in mice due to oversaturation of cellular microRNA/short hairpin RNA pathways. Nature 441:537-541.), (ii) off-targeting, i.e., silencing other genes, or/and (iii) general interference in the transcriptome (Jackson et al., 2003Jackson AL, Bartz SR, Schelter J, Kobayashi SV, Burchard J, Mao M, Li B, Cavet G and Linsley PS (2003) Expression profiling reveals off-target gene regulation by RNAi. Nat Biotechnol 21:635-637.).

We then initiated a medium-scale screening experiment in P. superbus with 97 genes implicated in anhydrobiosis using RNAi by soaking and observed a decrease in survival for nearly half of them. Many of these gene sequences, together with their possible roles in anhydrobiosis, were first discussed by Tyson et al. (2012)Tyson T, O’Mahony Zamora G, Wong S, Skelton M, Daly B, Jones JT, Mulvihill ED, Elsworth B, Phillips M, Blaxter M, et al. (2012) A molecular analysis of desiccation tolerance mechanisms in the anhydrobiotic nematode Panagrolaimus superbus using expressed sequenced tags. BMC Res Notes 5:68. who identified them in P. superbus after generating an EST library. Since this library was constructed using mixed populations of worms (no specific developmental stages) under normal humidity conditions, this study was not able to determine whether the genes identified were in fact involved in anhydrobiosis. Here we will discuss some of the genes which were not explicitly mentioned in previous work (Tyson et al., 2012Tyson T, O’Mahony Zamora G, Wong S, Skelton M, Daly B, Jones JT, Mulvihill ED, Elsworth B, Phillips M, Blaxter M, et al. (2012) A molecular analysis of desiccation tolerance mechanisms in the anhydrobiotic nematode Panagrolaimus superbus using expressed sequenced tags. BMC Res Notes 5:68.) and which our RNAi experiments suggest to have a functional role in P. superbus anhydrobiosis.

The most abundant EST found in P. superbus (Tyson et al., 2012Tyson T, O’Mahony Zamora G, Wong S, Skelton M, Daly B, Jones JT, Mulvihill ED, Elsworth B, Phillips M, Blaxter M, et al. (2012) A molecular analysis of desiccation tolerance mechanisms in the anhydrobiotic nematode Panagrolaimus superbus using expressed sequenced tags. BMC Res Notes 5:68.) encodes sxp/Ral-2 protein, a small (16–21 kDa) basic protein with a common domain of unknown function, which is highly expressed in parasitic nematodes and is secreted onto the surface of the worm cuticle. Several cuticle proteins are differentially expressed during desiccation in diverse species (Adhikari et al., 2009Adhikari BN, Wall DH and Adams BJ (2009) Desiccation survival in an Antarctic nematode: Molecular analysis using expressed sequenced tags. BMC Genomics 10:69.; Cornette et al., 2010Cornette R, Kanamori Y, Watanabe M, Nakahara Y, Gusev O, Mitsumasu K, Kadono-Okuda K, Shimomura M, Mita K, Kikawada T, et al. (2010) Identification of anhydrobiosis-related genes from an expressed sequence tag database in the cryptobiotic midge Polypedilum vanderplanki (Diptera; Chironomidae). J Biol Chem 285:35889-35999.), which are possibly involved in modification of cuticle permeability and which, along with aquaporins, surface lipids (Wharton et al., 2008Wharton DA, Petrone L, Duncan A and McQuillan AJ (2008) A surface lipid may control the permeability slump associated with entry into anhydrobiosis in the plant parasitic nematode Ditylenchus dipsaci. J Exp Biol 211:2901-2908.) and behavioral responses (worm coiling/clumping), might promote controlled water loss during dehydration. We were able to validate an anatomically related polypeptide: the `cuticle collagen protein LON-3’ (target code “si31”), a polypeptide involved in C. elegans body shape and probably targeted by TGF-beta signaling (Suzuki et al., 2002Suzuki Y, Morris GA, Han M and Wood WB (2002) A cuticle collagen encoded by the lon-3 gene may be a target of TGF-beta signaling in determining Caenorhabditis elegans body shape. Genetics 162:1631-1639.). Notably, the surface of C. elegans displays indentations in its circumference spaced about one micrometer apart, defining rings called annuli. It is suggested that annuli may function as pleats, allowing the cuticle to fold on the inner radius of a bend and extend over the outer radius (Riddle et al., 1997Riddle DL, Blumenthal T, Meyer BJ and Priess JR (1997). C. elegans II. 2nd edition. Cold Spring Harbor Laboratory Press, New York, 1222 p.). It was recently shown that C. elegans lon-2 mutants present with wider annuli and a decrease in furrow depth (Essmann et al., 2017Essmann CL, Elmi M, Shaw M, Anand GM, Pawar VM and Srinivasan MA (2017) In-vivo high resolution AFM topographic imaging of Caenorhabditis elegans reveals previously unreported surface structures of cuticle mutants. Nanomedicine 13:183-189.). Therefore, morphological changes promoted by lon-3 (which is closely related to lon-2) possibly involve alterations in annuli also, thereby altering the total body surface and controlling water loss.

Schokraie et al. (2010)Schokraie E, Hotz-Wagenblatt A, Warnken U, Mali B, Frohme M, Förster F, Dandekar T, Hengherr S, Schill RO and Schnölzer M (2010) Proteomic analysis of tardigrades: Towards a better understanding of molecular mechanisms by anhydrobiotic organisms. PLoS One 5:e9502. found three different cathepsins (K, Z and L1) during a proteomic study in the tardigrade M. tardigradum. One of them, cathepsin-L-like cysteine proteinase (target code “si25”), which is a ubiquitous protease in eukaryotes, is associated with desiccation tolerance in P. superbus. The parasitic nematode Parelaphostrongylus tenuis expresses a cathepsin B cysteine protease homolog which is abundant in larval stages. Although it is less abundant in larval stage L3, it is still predominant during this developmental phase. Curiously, L3 is the phase when the parasite leaves the intermediate host (snail) and it was observed that, to some extent, L3 is desiccation tolerant, allowing persistence in the environment (Duffy et al., 2006Duffy MS, Cevasco DK, Zarlenga DS, Sukhumavasi W and Appleton JA (2006) Cathepsin B homologue at the interface between a parasitic nematode and its intermediate host. Infect Immun 74:1297-1304.). Although most cathepsins degrade autophagosomal content, cathepsin L also degrades lysosomal membrane components (Kaminskyy and Zhivotovsky, 2012Kaminskyy V and Zhivotovsky B (2012) Proteases in autophagy. Biochim Biophys Acta 1824:44-50.). Therefore, this enzyme could be involved in a general turnover of damaged proteins after desiccation. Several other proteases have been implicated in desiccation tolerance in diverse species, including (i) serine endopeptidases and aminopeptidases in the resurrection plant Ramonda serbica (Kidric et al., 2014Kidric M, Sabotic J and Stevanovic B (2014) Desiccation tolerance of the resurrection plant Ramonda serbica is associated with dehydration-dependent changes in levels of proteolytic activities. Plant Physiol 171:998-1002.), (ii) ATP-dependent ClpXP protease in the bacterium Staphylococcus aureus (Chaibenjawong and Foster, 2011Chaibenjawong P and Foster SJ (2011) Desiccation tolerance in Staphylococcus aureus. Arch Microbiol 193:125-135.), and (iii) carboxy-terminal protease (CtpA) in Rhizobium leguminosarum (Gilbert et al., 2007Gilbert KB, Vanderlinde EM and Yost CK (2007) Mutagenesis of the carboxy terminal protease CtpA decreases desiccation tolerance in Rhizobium leguminosarum. FEMS Microbiol Lett 272:65-74.), indicating an important role in proteome turnover mediated by proteinases.

Since P. superbus is a fast desiccation strategist, i.e. it may enter anhydrobiosis without preconditioning, it is likely that a substantial proportion of its proteome is constitutively primed to enter the dry state. Therefore, the relatively large number of kinases (24) revealed by our RNAi experiments to be important for anhydrobiosis may act as essential signaling regulators or as activators of the proteome, by phosphorylating their substrates within a very short period, allowing rapid entry in the dry state. Other studies have also demonstrated a high number of such proteins involved in dehydration/drought/desiccation tolerance in different plant species, including 229 kinases in chrysanthemum plants (Xu et al., 2013Xu Y, Gao S, Yang Y, Huang M, Cheng L, Wei Q, Fei Z, Gao J and Hong B (2013) Transcriptome sequencing and whole genome expression profiling of chrysanthemum under dehydration stress. BMC Genomics 14:662.) and over 460 kinases in the resurrection plant Myrothamnus flabellifoliathe (Ma et al., 2015Ma C, Wang H, Macnish AJ, Estrada-Melo AC, Lin J, Chang Y, Reid MS and Jiang CZ (2015) Transcriptomic analysis reveals numerous diverse protein kinases and transcription factors involved in desiccation tolerance in the resurrection plant Myrothamnus flabellifolia. Hortic Res 22:15034.). Still within this context, adenylate kinase (target code “19K”), an enzyme involved in the biosynthesis of ATP, was demonstrated to be up-regulated in a drought-tolerant genotype of tomato (Gong et al., 2010Gong P, Zhang J, Li H, Yang C, Zhang C, Zhang X, Khurram Z, Zhang Y, Wang T, Fei Z, et al. (2010) Transcriptional profiles of drought-responsive genes in modulating transcription signal transduction, and biochemical pathways in tomato. J Exp Bot 61:3563-3575.). In P. superbus, accumulation of ATP during dehydration is probably a key aspect of the whole process, providing a readily available source of energy during rehydration, when cells need energy to resume activities, but are not fully capable of generating it.

An important aspect of our approach relies on the fact that since our panel of 97 targets were genes differentially expressed during desiccation in different anhydrobiotic species (the nematode Plectus murrayi, the dipteran Polypedilum vanderplanki and two tardigrades Hypsibius dujardini and Milnesium tardigradum), one might expect that most (if not all) of them would be shown to be anhydrobiosis-related in P. superbus. However, our study confirmed the association of only 40 of them. This result may reflect that different anhydrobiotic species adopt different biochemical strategies and/or molecular programs to promote desiccation tolerance. For example, although many anhydrobiotic animals accumulate trehalose during desiccation, bdelloid rotifers do not (Lapinski and Tunnacliffe, 2003Lapinski J and Tunnacliffe A (2003) Anhydrobiosis without trehalose in bdelloid rotifers. FEBS Lett 553:387-390.), while plants often accumulate sucrose (Zhang et al., 2016Zhang Q, Song X and Bartels D (2016) Enzymes and metabolites in carbohydrate metabolism of desiccation tolerant plants. Proteomes 4:40.). Moreover, P. superbus is a fast strategist (i.e., it is able to enter anhydrobiosis in the absence of preconditioning), while other species demand a slow desiccation protocol (Shannon et al., 2005Shannon AJ, Browne JA, Boyd J, Fitzpatrick DA and Burnell AM (2005) The anhydrobiotic potential and molecular phylogenetics of species and strains of Panagrolaimus (Nematoda, Panagrolaimidae). J Exp Biol 208:2433-2445.). Thus, we cannot unequivocally rule out the involvement of the other 57 assessed targets in anhydrobiosis (Table 1, in white) since their roles may be minor (secondary) within anhydrobiosis, or be compensated by other genes, demanding other genetic analyses (e.g., CRISPR-mediated single- or multiple-knockouts, which is not established for this species) to determine it.

The set of genes identified here, along with other similar functional studies, might be used for the development of anhydrobiotic engineering (García De Castro et al., 2000García De Castro A, Bredholt H, Strøm AR and Tunnacliffe A (2000) Anhydrobiotic engineering of gram-negative bacteria. Appl Environ Microbiol 66:4142-4124.), a research field which aims to render cells and whole organisms tolerant to dehydration. Several anhydrobiosis-based approaches have recently been developed in order to preserve biological samples at room temperature for longer periods, including the stabilization of RNA molecules (Hernandez et al., 2009Hernandez GE, Mondala TS and Head SR (2009) Assessing a novel room-temperature RNA storage medium for compatibility in microarray gene expression analysis. Biotechniques 47:667-670.) and poxviral/adenoviral vaccines without refrigeration (Alcock et al., 2010Alcock R, Cottingham MG, Rollier CS, Furze J, De Costa SD, Hanlon M, Spencer AJ, Honeycutt JD, Wyllie DH, Gilbert SC, et al. (2010) Long-term thermostabilization of live poxviral and adenoviral vaccine vectors at supraphysiological temperatures in carbohydrate glass. Sci Transl Med 2:19ra12.). Further challenges encompass the development of transgenic plants tolerant to extreme drought, a recurrent and increasing challenge in agriculture and food production. This may be even more relevant considering the expected population growth over the next decades and the predictions of a rise in global temperature. Heterologous expression of just one desiccation-related protein is sufficient to promote a significant increase in drought tolerance (Wang et al., 2017Wang B, Du H, Zhang Z, Xu W and Deng X (2017) BhbZIP60 from Resurrection Plant Boea hygrometrica is an mRNA splicing-activated endoplasmic reticulum stress regulator involved in drought tolerance. Front Plant Sci 8:245.). For example, Liu et al. (2009)Liu X, Wang Z, Wang LL, Wu RH, Phillips J and Deng X (2009) LEA 4 group genes from the resurrection plant Boea hygrometrica confer dehydration tolerance in transgenic tobacco. Plant Sci 176:90-98. produced transgenic tobacco plants expressing LEA proteins derived from the resurrection plant Boea hygrometrica. Transgenic plants expressing BhLEA1 (Boea hygrometrica LEA1) or BhLEA2 were submitted to water stress and, compared to a control group, displayed (i) higher water content, (ii) higher activities of photosystem II, superoxide dismutase and peroxidase, (iii) lower membrane permeability and (iv) stabilization of several proteins including ribulose-bisphosphate carboxylase (large subunit).

On the other hand, medicine might also benefit from discoveries on the molecular basis of anhydrobiosis. This might be achieved via strategies based on the concept of ‘DNA vaccines’, which promote the transient expression of heterologous proteins within the human body. This approach allows, for example, the expression of virus-derived proteins in the body, which, in turn, trigger the immune system to produce corresponding antibodies and promote protection (Dowd et al., 2016Dowd KA, Ko SY, Morabito KM, Yang ES, Pelc RS, DeMaso CR, Castilho LR, Abbink P, Boyd M and Nityanandam R, et al. (2016) Rapid development of a DNA vaccine for Zika virus. Science 354:237-240.). Such an approach might be used to express anhydrobiosis-related genes for novel purposes. For example, Hashimoto et al. (2016)Hashimoto T, Horikawa DD, Saito Y, Kuwahara H, Kozuka-Hata H, Shin-I T, Minakuchi Y, Ohishi K, Motoyama A, Aizu T, et al. (2016) Extremotolerant tardigrade genome and improved radiotolerance of human cultured cells by tardigrade-unique protein. Nat Commun 7:12808. reported several unique genes in the anhydrobiotic tardigrate Ramazzottius varieornatus. One of these was shown to suppress X-ray-induced DNA damage by nearly 40% and to improve radiotolerance when expressed in human cell cultures. Therefore, heterologous expression of the “DNA repair protein RAD51 homolog 1” (target si42, identified in our present study) via strategies based on the concept of `DNA vaccines’ might eventually be useful for people submitted to radiotherapy during treatment against cancer. More elaborate and complex strategies might allow, in the future, the expression of several anhydrobiosis-related proteins and the preservation in the dry state and at room temperature of human organs for transplant until a compatible patient is found.

Finally, the current work, along with the impending publication of the P. superbus genome by other groups, helps to establish P. superbus as a nematode model for anhydrobiosis. Comparative studies between P. superbus and C. elegans, which have a fundamentally different approach to desiccation tolerance, may shed light on the evolution of desiccation tolerance in these proximal species.

Conclusions

We have found evidence for the participation of 40 anhydrobiosis-related genes in P. superbus. Our data, along with transcriptomic and proteomic analyses, help unveil the genetic scaffold underlying extreme desiccation tolerance, as well as the development of anhydrobiotic engineering.

Acknowledgments

This work was supported by Fundação de Amparo a Pesquisa do Estado de São Paulo (FAPESP, process 2008/54236-7) and Conselho Nacional de Desenvolvimento Científico e Tecnológico (CNPq, process 472427/2008-3). CCSE and TAJS were recipients of student fellowships from CAPES (Coordenação de Aperfeiçoamento de Pessoal de Nível Superior); TCP, GVG, TFA and GB were recipients of fellowships from FAPESP (Processes 2009/01520-2; 2010/14905-7; 2010/11756-0; 2012/06441-6). AT received support from the European Research Council in the form of an Advanced Investigator Award (AdG 233232). We also thank Trevor Tyson for providing P. superbus kinase cDNAs.

References

  • Adhikari BN, Wall DH and Adams BJ (2009) Desiccation survival in an Antarctic nematode: Molecular analysis using expressed sequenced tags. BMC Genomics 10:69.
  • Alcock R, Cottingham MG, Rollier CS, Furze J, De Costa SD, Hanlon M, Spencer AJ, Honeycutt JD, Wyllie DH, Gilbert SC, et al. (2010) Long-term thermostabilization of live poxviral and adenoviral vaccine vectors at supraphysiological temperatures in carbohydrate glass. Sci Transl Med 2:19ra12.
  • Boschetti C, Carr A, Crisp A, Eyres I, Wang-Koh Y, Lubzens E, Barraclough TG, Micklem G and Tunnacliffe A (2012) Biochemical diversification through foreign gene expression in bdelloid rotifers. PLoS Genet 8:e1003035.
  • Boschetti C, Pouchkina-Stantcheva N, Hoffmann P and Tunnacliffe A (2011) Foreign genes and novel hydrophilic protein genes participate in the desiccation response of the bdelloid rotifer Adineta ricciae J Exp Biol 214:59-68.
  • Boothby TC, Tapia H, Brozena AH, Piszkiewicz S, Smith AE, Giovannini I, Rebecchi L, Pielak GJ, Koshland D and Goldstein B (2017) Tardigrades use intrinsically disordered proteins to survive desiccation. Mol Cell 65:975-984.
  • Burnell AM and Tunnacliffe A (2011) Gene induction and desiccation stress in nematodes. In: Perry RN and Wharton (eds) Molecular and Physiological Basis of Nematode Survival. CABI Press, Wallingford, pp 126-156.
  • Chaibenjawong P and Foster SJ (2011) Desiccation tolerance in Staphylococcus aureus Arch Microbiol 193:125-135.
  • Chen W, Zhang X, Liu M, Zhang J, Ye Y, Lin Y, Luyckx J and Qu J (2009) Trehalose protects against ocular surface disorders in experimental murine dry eye through suppression of apoptosis. Exp Eye Res 89:311-318.
  • Cornette R, Kanamori Y, Watanabe M, Nakahara Y, Gusev O, Mitsumasu K, Kadono-Okuda K, Shimomura M, Mita K, Kikawada T, et al. (2010) Identification of anhydrobiosis-related genes from an expressed sequence tag database in the cryptobiotic midge Polypedilum vanderplanki (Diptera; Chironomidae). J Biol Chem 285:35889-35999.
  • Crowe JH, Carpenter JF and Crowe LM (1998) The role of vitrification in anhydrobiosis. Annu Rev Physiol 60:73-103.
  • Culleton BA, Lall P, Kinsella GK, Doyle S, McCaffrey J, Fitzpatrick DA and Burnell AM (2015) A role for the Parkinson’s disease protein DJ-1 as a chaperone and antioxidant in the anhydrobiotic nematode Panagrolaimus superbus Cell Stress Chaperones 20:121-137.
  • Garcia de Castro A, Lapinski J and Tunnacliffe A (2000) Anhydrobiotic engineering. Nat Biotechnol 18:473.
  • Dowd KA, Ko SY, Morabito KM, Yang ES, Pelc RS, DeMaso CR, Castilho LR, Abbink P, Boyd M and Nityanandam R, et al. (2016) Rapid development of a DNA vaccine for Zika virus. Science 354:237-240.
  • Duffy MS, Cevasco DK, Zarlenga DS, Sukhumavasi W and Appleton JA (2006) Cathepsin B homologue at the interface between a parasitic nematode and its intermediate host. Infect Immun 74:1297-1304.
  • Erkut C, Penkov S, Khesbak H, Vorkel D, Verbavatz JM, Fahmy K and Kurzchalia TV (2011) Trehalose renders the dauer larva of Caenorhabditis elegans resistant to extreme desiccation. Curr Biol 21:1331-1336.
  • Erkut C, Vasilj A, Boland S, Habermann B, Shevchenko A and Kurzchalia TV (2013) Molecular strategies of the Caenorhabditis elegans dauer larva to survive extreme desiccation. PLoS One 8:e82473.
  • Essmann CL, Elmi M, Shaw M, Anand GM, Pawar VM and Srinivasan MA (2017) In-vivo high resolution AFM topographic imaging of Caenorhabditis elegans reveals previously unreported surface structures of cuticle mutants. Nanomedicine 13:183-189.
  • Feinberg EH and Hunter CP (2003) Transport of dsRNA into cells by the transmembrane protein SID-1. Science 301:1545-1547.
  • França MB, Panek AD and Eleutherio EC (2007) Oxidative stress and its effects during dehydration. Comp Biochem Physiol A Mol Integr Physiol 146:621-631.
  • Fuchs G (1930) Neue an Borkenkäfer und Russelkäfer gebundene Nematoden, halbparasitische und Wohnungseinmieter. Zool Jb 59:586-608.
  • Gal TZ, Glazer I and Koltai H (2004) An LEA group 3 family member is involved in survival of C. elegans during exposure to stress. FEBS Lett 577:21-26.
  • García De Castro A, Bredholt H, Strøm AR and Tunnacliffe A (2000) Anhydrobiotic engineering of gram-negative bacteria. Appl Environ Microbiol 66:4142-4124.
  • Gilbert KB, Vanderlinde EM and Yost CK (2007) Mutagenesis of the carboxy terminal protease CtpA decreases desiccation tolerance in Rhizobium leguminosarum FEMS Microbiol Lett 272:65-74.
  • Gong P, Zhang J, Li H, Yang C, Zhang C, Zhang X, Khurram Z, Zhang Y, Wang T, Fei Z, et al. (2010) Transcriptional profiles of drought-responsive genes in modulating transcription signal transduction, and biochemical pathways in tomato. J Exp Bot 61:3563-3575.
  • Grimm D, Streetz KL, Jopling CL, Storm TA, Pandey K, Davis CR, Marion P, Salazar F and Kay MA (2006) Fatality in mice due to oversaturation of cellular microRNA/short hairpin RNA pathways. Nature 441:537-541.
  • Haegeman A, Jacob J, Vanholme B, Kyndt T, Mitreva M and Gheysen G (2009) Expressed sequence tags of the peanut pod nematode Ditylenchus africanus: The first transcriptome analysis of an Anguinid nematode. Mol Biochem Parasitol 167:32-40.
  • Hansen JM, Go YM and Jones DP (2006) Nuclear and mitochondrial compartmentation of oxidative stress and redox signaling. Annu Rev Pharmacol Toxicol 46:215-234.
  • Hashimoto T, Horikawa DD, Saito Y, Kuwahara H, Kozuka-Hata H, Shin-I T, Minakuchi Y, Ohishi K, Motoyama A, Aizu T, et al. (2016) Extremotolerant tardigrade genome and improved radiotolerance of human cultured cells by tardigrade-unique protein. Nat Commun 7:12808.
  • Hernandez GE, Mondala TS and Head SR (2009) Assessing a novel room-temperature RNA storage medium for compatibility in microarray gene expression analysis. Biotechniques 47:667-670.
  • Issa Z, Grant WN, Stasiuk S and Shoemaker CB (2005) Development of methods for RNA interference in the sheep gastrointestinal parasite, Trichostrongylus colubriformis Int J Parasitol 35:935-940.
  • Jackson AL, Bartz SR, Schelter J, Kobayashi SV, Burchard J, Mao M, Li B, Cavet G and Linsley PS (2003) Expression profiling reveals off-target gene regulation by RNAi. Nat Biotechnol 21:635-637.
  • Jönsson KI (2005) The evolution of life histories in holo-anhydrobiotic animals: A first approach. Integr Comp Biol 45:764-770.
  • Kaminskyy V and Zhivotovsky B (2012) Proteases in autophagy. Biochim Biophys Acta 1824:44-50.
  • Kidric M, Sabotic J and Stevanovic B (2014) Desiccation tolerance of the resurrection plant Ramonda serbica is associated with dehydration-dependent changes in levels of proteolytic activities. Plant Physiol 171:998-1002.
  • Lapinski J and Tunnacliffe A (2003) Anhydrobiosis without trehalose in bdelloid rotifers. FEBS Lett 553:387-390.
  • Larsen PL (1993) Aging and resistance to oxidative damage in Caenorhabditis elegans Proc Natl Acad Sci U S A 90:8905-8909.
  • Leprince O, Atherton NM, Deltour R and Hendry GAF (1994) The involvement of respiration in free radical processes during loss of desiccation tolerance in germinating Zea mays L. (an electron paramagnetic resonance study). Plant Physiol 104:1333-1339.
  • Li S, Chakraborty N, Borcar A, Menze MA, Toner M and Hand SC (2012) Late embryogenesis abundant proteins protect human hepatoma cells during acute desiccation. Proc Natl Acad Sci U S A 109:20859-20864.
  • Liu X, Wang Z, Wang LL, Wu RH, Phillips J and Deng X (2009) LEA 4 group genes from the resurrection plant Boea hygrometrica confer dehydration tolerance in transgenic tobacco. Plant Sci 176:90-98.
  • López-Cristoffanini C, Zapata J, Gaillard F, Potin P, Correa JA and Contreras-Porcia L (2015) Identification of proteins involved in desiccation tolerance in the red seaweed Pyropia orbicularis (Rhodophyta, Bangiales). Proteomics 15:3954-3968.
  • Ma C, Wang H, Macnish AJ, Estrada-Melo AC, Lin J, Chang Y, Reid MS and Jiang CZ (2015) Transcriptomic analysis reveals numerous diverse protein kinases and transcription factors involved in desiccation tolerance in the resurrection plant Myrothamnus flabellifolia Hortic Res 22:15034.
  • Mali B, Grohme MA, Förster F, Dandekar T, Schnölzer M, Reuter D, Welnicz W, Schill RO and Frohme M (2010) Transcriptome survey of the anhydrobiotic tardigrade Milnesium tardigradum in comparison with Hypsibius dujardini and Richtersius coronifer BMC Genomics 11:168.
  • McGill LM, Shannon AJ, Pisani D, Félix MA, Ramløv H, Dix I, Wharton DA and Burnell AM (2015) Anhydrobiosis and freezing-tolerance: Adaptations that facilitate the establishment of Panagrolaimus nematodes in polar habitats. PLoS One 10:e0116084.
  • Michiels C, Raes M, Toussaint O and Remacle J (1994) Importance of Se-glutathione peroxidase, catalase, and Cu/Zn-SOD for cell survival against oxidative stress. Free Radic Biol Med 17:235-248.
  • Pereira TC, Pascoal VDB, Secolin R, Rocha CS, Maia IG and Lopes-Cendes I (2007) Strand Analysis, a free online program for the computational identification of the best RNA interference (RNAi) targets based on Gibbs free energy. Genet Mol Biol 30:1206-1208.
  • Reardon W, Chakrabortee S, Pereira TC, Tyson T, Banton MC, Dolan KM, Culleton BA, Wise MJ, Burnell AM and Tunnacliffe A (2010) Expression profiling and cross-species RNA interference (RNAi) of desiccation-induced transcripts in the anhydrobiotic nematode Aphelenchus avenae BMC Mol Biol 11:6.
  • Ricci C and Caprioli, M (1998) Stress during dormancy: Effect on recovery rates and life-history traits of anhydrobiotic animals. Aquat Ecol 32:353-359.
  • Ricci C, Vaghi L and Manzini ML (1987) Desiccation of rotifers (Macrotrachela quadricornifera) - Survival and reproduction. Ecology 68:1488-1494.
  • Riddle DL, Blumenthal T, Meyer BJ and Priess JR (1997). C. elegans II. 2nd edition. Cold Spring Harbor Laboratory Press, New York, 1222 p.
  • Ryabova A, Mukae K, Cherkasov A, Cornette R, Shagimardanova E, Sakashita T, Okuda T, Kikawada T and Gusev O (2017) Genetic background of enhanced radioresistance in an anhydrobiotic insect: Transcriptional response to ionizing radiations and desiccation. Extremophiles 21:109-120.
  • Sakurai M, Furuki T, Akao K, Tanaka D, Nakahara Y, Kikawada T, Watanabe M and Okuda T (2008) Vitrification is essential for anhydrobiosis in an African chironomid, Polypedilum vanderplanki Proc Natl Acad Sci U S A 105:5093-5098.
  • Schramm U and Becker W (1987) Anhydrobiosis of the bdelloid rotifer Habrotrocha rosa (Aschelminthes). Z Mikrosk Anat Forsch 101:1-17.
  • Schokraie E, Hotz-Wagenblatt A, Warnken U, Mali B, Frohme M, Förster F, Dandekar T, Hengherr S, Schill RO and Schnölzer M (2010) Proteomic analysis of tardigrades: Towards a better understanding of molecular mechanisms by anhydrobiotic organisms. PLoS One 5:e9502.
  • Schokraie E, Warnken U, Hotz-Wagenblatt A, Grohme MA, Hengherr S, Förster F, Schill RO, Frohme M, Dandekar T and Schnölzer M (2012) Comparative proteome analysis of Milnesium tardigradum in early embryonic state versus adults in active and anhydrobiotic state. PLoS One 7:e45682.
  • Shannon AJ, Browne JA, Boyd J, Fitzpatrick DA and Burnell AM (2005) The anhydrobiotic potential and molecular phylogenetics of species and strains of Panagrolaimus (Nematoda, Panagrolaimidae). J Exp Biol 208:2433-2445.
  • Shannon AJ, Tyson T, Dix I, Boyd J and Burnell AM (2008) Systemic RNAi mediated gene silencing in the anhydrobiotic nematode Panagrolaimus superbus BMC Mol Biol 9:58.
  • Shih JD, Fitzgerald MC, Sutherlin M and Hunter CP (2009) The SID-1 double-stranded RNA transporter is not selective for dsRNA length. RNA15:384-390.
  • Sies H (2000) What is oxidative stress? In: Keaney JF (ed) Oxidative Stress and Vascular Disease. Kluwer Academic Publishers, Dordrecht, pp 1-8.
  • Sohlenius B (1988) Interactions between two species of Panagrolaimus in agar cultures. Nematologica 34:208-217.
  • Suzuki Y, Morris GA, Han M and Wood WB (2002) A cuticle collagen encoded by the lon-3 gene may be a target of TGF-beta signaling in determining Caenorhabditis elegans body shape. Genetics 162:1631-1639.
  • Tunnacliffe A and Lapinski J (2003) Resurrecting Van Leeuwenhoek’s rotifers: A reappraisal of the role of disaccharides in anhydrobiosis. Philos Trans R Soc Lond B Biol Sci 358:1755-1771.
  • Tyson T, O’Mahony Zamora G, Wong S, Skelton M, Daly B, Jones JT, Mulvihill ED, Elsworth B, Phillips M, Blaxter M, et al. (2012) A molecular analysis of desiccation tolerance mechanisms in the anhydrobiotic nematode Panagrolaimus superbus using expressed sequenced tags. BMC Res Notes 5:68.
  • Van Leeuwenhoek A (1702) Letter to Hendrik van Bleyswijk, dated 9 February 1702.
  • Wang C, Grohme MA, Mali B, Schill RO and Frohme M (2014) Towards decrypting cryptobiosis - analyzing anhydrobiosis in the tardigrade Milnesium tardigradum using transcriptome sequencing. PLoS One 9:e92663.
  • Wang B, Du H, Zhang Z, Xu W and Deng X (2017) BhbZIP60 from Resurrection Plant Boea hygrometrica is an mRNA splicing-activated endoplasmic reticulum stress regulator involved in drought tolerance. Front Plant Sci 8:245.
  • Watanabe M, Sakashita T, Fujita A, Kikawada T, Nakahara Y, Hamada N, Horikawa DD, Wada S, Funayama T, Kobayashi Y, et al. (2006a) Estimation of radiation tolerance to high LET heavy ions in an anhydrobiotic insect, Polypedilum vanderplanki Int J Radiat Biol 82:835-842.
  • Watanabe M, Sakashita T, Fujita A, Kikawada T, Horikawa DD, Nakahara Y, Wada S, Funayama T, Hamada N, Kobayashi Y, et al. (2006b) Biological effects of anhydrobiosis in an African chironomid, Polypedilum vanderplanki on radiation tolerance. Int J Radiat Biol 82:587-592.
  • Wharton DA, Petrone L, Duncan A and McQuillan AJ (2008) A surface lipid may control the permeability slump associated with entry into anhydrobiosis in the plant parasitic nematode Ditylenchus dipsaci J Exp Biol 211:2901-2908.
  • Xu Y, Gao S, Yang Y, Huang M, Cheng L, Wei Q, Fei Z, Gao J and Hong B (2013) Transcriptome sequencing and whole genome expression profiling of chrysanthemum under dehydration stress. BMC Genomics 14:662.
  • Yamaguchi A, Tanaka S, Yamaguchi S, Kuwahara H, Takamura C, Imajoh-Ohmi S, Horikawa DD, Toyoda A, Katayama T, Arakawa K, et al. (2012) Two novel heat-soluble protein families abundantly expressed in an anhydrobiotic tardigrade. PLoS One 7:e44209.
  • Yu JY, DeRuiter SL and Turner DL (2002) RNA interference by expression of short-interfering RNAs and hairpin RNAs in mammalian cells. Proc Natl Acad Sci U S A 99:6047-6052.
  • Zhang Q, Song X and Bartels D (2016) Enzymes and metabolites in carbohydrate metabolism of desiccation tolerant plants. Proteomes 4:40.
  • Zhu Y, Wang B, Phillips J, Zhang ZN, Du H, Xu T, Huang LC, Zhang XF, Xu GH, Li WL, et al. (2015) Global transcriptome analysis reveals acclimation-primed processes involved in the acquisition of desiccation tolerance in Boea hygrometrica Plant Cell Physiol 56:1429-1441.
  • Associate Editor: Igor Schneider

Publication Dates

  • Publication in this collection
    06 Nov 2017
  • Date of issue
    Oct-Dec 2017

History

  • Received
    07 Feb 2017
  • Accepted
    20 July 2017
Sociedade Brasileira de Genética Rua Cap. Adelmio Norberto da Silva, 736, 14025-670 Ribeirão Preto SP Brazil, Tel.: (55 16) 3911-4130 / Fax.: (55 16) 3621-3552 - Ribeirão Preto - SP - Brazil
E-mail: editor@gmb.org.br