Next Article in Journal
Progesterone Promotes In Vitro Maturation of Domestic Dog Oocytes Leading to Successful Live Births
Previous Article in Journal
Effects of Carbonated Beverage Consumption on Oral pH and Bacterial Proliferation in Adolescents: A Randomized Crossover Clinical Trial
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Heat Shock Proteins (Hsps) in Cellular Homeostasis: A Promising Tool for Health Management in Crustacean Aquaculture

Aquatic Environmental Biotechnology and Nanotechnology (AEBN) Division, ICAR-Central Inland Fisheries Research Institute (CIFRI), Barrackpore 700120, India
*
Authors to whom correspondence should be addressed.
Life 2022, 12(11), 1777; https://doi.org/10.3390/life12111777
Submission received: 13 September 2022 / Revised: 27 October 2022 / Accepted: 28 October 2022 / Published: 3 November 2022
(This article belongs to the Section Animal Science)

Abstract

:
Heat shock proteins (Hsps) are a family of ubiquitously expressed stress proteins and extrinsic chaperones that are required for viability and cell growth in all living organisms. These proteins are highly conserved and produced in all cellular organisms when exposed to stress. Hsps play a significant role in protein synthesis and homeostasis, as well as in the maintenance of overall health in crustaceans against various internal and external environmental stresses. Recent reports have suggested that enhancing in vivo Hsp levels via non-lethal heat shock, exogenous Hsps, or plant-based compounds, could be a promising strategy used to develop protective immunity in crustaceans against both abiotic and biotic stresses. Hence, Hsps as the agent of being an immune booster and increasing disease resistance will present a significant advancement in reducing stressful conditions in the aquaculture system.

1. Introduction

From the past to the present, crustacean aquaculture has been increasing considerably in order to reach the higher food demands of the world. The total crustacean aquaculture production, from over 30 different species, was 8.4 MT valued at USD 61.06 billion, with an average annual growth rate of 9.92% per year since 2000 [1,2]. The marine shrimp currently dominate crustacean aquaculture at 5.51 MT or 65.3% of total crustaceans (valued at USD 34.2 billion), followed by freshwater crustaceans (2.53 MT or 29.9% of total crustaceans and are valued at USD 24.3 billion). Despite tremendous development, the aquaculture sector is confronted with various problems, some of which have a detrimental impact on its ability to achieve sustainable outcomes [3,4]. Disease outbreaks in aquaculture have been one of the most serious issues, causing significant output losses, sometimes up to 100%. The adverse environmental conditions in the culture systems caused by, for instance, extreme weather events and crowding stress are some of the major drivers of disease emergence. Pathogenic and non-pathogenic microbial populations coexist with farmed animals in an aquaculture production system, however, when they are stressed, acute disease and infections may develop [5,6]. It is noteworthy that not all pathogens cause significant mortality; some contribute to heavy losses through their negative impacts on growth, appetite, susceptibility to adverse environmental conditions, and many others. For instance, acute hepatopancreatic necrosis disease (AHPND) caused by Vibrio parahaemolyticus and Microsporidian Enterocytozoon hepatopenaei, a shrimp pathogen that was discovered in Thailand over ten years back, has now become a widespread and highly impacting pathogen responsible for growth retardation, yet without causing a significant effect on shrimp survival [7,8,9]. Based on the discussion made above and the evidence from earlier studies, it can be suggested that biological agents/pathogens and environmental factors interrelate in various complex ways with the host in order to cause diseases.
The culture of crustacean animals in the most optimal conditions is not often economically feasible; as such, there will always be a risk of infection and a need for effective disease control strategies in order to make the aquaculture industry more sustainable. Among the various drugs, antibiotics are the most popular and widely used chemotherapeutic agent used to control bacterial disease in crustacean aquaculture [10,11]. However, the wide use of antibiotics, and especially their application at subtherapeutic doses, inevitably results in the development of antibiotic-resistant bacteria. In fact, when antibiotics are used, the resistant bacterial strains can multiply rapidly due to the fact that their competitors (sensitive bacterial strains) are suppressed. Moreover, it is generally accepted that the acquisition of resistance is more rapid against bactericidal antibiotics. As antibiotic-resistant genes are often located on mobile genetic elements, such as in transferable plasmids and integrons, the antibiotic-resistant mobile genetic elements can be transmitted via horizontal gene transfer, not only to other bacteria but also to terrestrial bacteria, including humans and animals’ pathogens [12,13,14]. This resistance has already been reported in some human pathogenic bacteria, including Salmonella enterica serotype Typhimurium and Vibrio cholera. Given the worldwide trade in aquaculture products, health problems associated with antibiotic use are not limited to aquaculture food-producing countries but are in the countries importing antibiotic-treated food products [15,16]. Due to the magnitude of the challenges that the industry is faced with as it grows, new alternatives to the use of antimicrobials are urgently needed, since no anti-infective treatment appears to be capable of solving every problem.
Crustaceans, similar to other invertebrates, lack an adaptive immune system. Instead, they rely on their effective cellular and humoral innate immune system in order to defend against hostile microorganisms. In crustaceans, including shrimps, the major immune reactions occur in the hemolymph, which consists of three different principal types of hemocytes, i.e., hyaline, granular, and semi-granular hemocytes [17,18]. Over the past few years, several strategies/preventive measures to enhance non-specific immune defense systems and protect crustacean animals from pathogenic microbes, without using antibiotics, have been developed; further, some of them have even been applied successfully. For example, in regard to the monitoring and managing of microbes in the culture system, some of the promising alternatives are microbe-associated feeding strategies, (i.e., natural or plant-based compounds), prebiotics, probiotics, and synbiotics, improving welfare and stress prevention, and the use of natural immunostimulants. Interestingly, the heat shock proteins (Hsps) that have been reported to potentiate the generation of pro-oxidant activity are responsible for the induction of protective immune responses in crustacean species [19,20]. Hsps, also known as stress proteins and extrinsic chaperones, are a group of evolutionarily conserved proteins. Hsps are well known as molecular chaperones, and they are known for aiding nascent polypeptide folding and oligomerization; protecting proteins from irreversible denaturation; re-folding or degrading damaged proteins; translocating proteins into membrane-bound cell compartments; and contributing to disease resistance [19,20,21,22,23,24,25]. Hence, a holistic approach—which includes the environment, host, and pathogen—and develops pharmacological agents capable of inducing Hsp responses, could protect the animals at different stages of life and against a broad spectrum of infectious agents; further, this type of approach will also most likely be the most sustainable [26,27,28,29,30]. In this new concept of disease control strategy, measures that prevent disease are the most important health management options. The focus of the present review is to provide a details overview of Hsps and their potential role in host immune response, protein homeostasis, and cross-protection against abiotic and biotic stress through different regulatory pathways. In addition, we have also tried to explore the in vivo working mechanism of Hsps by using a model organism, A. franciscana (brine shrimp). This model is utilized through different regulatory pathways and possibilities for the purposes of developing compounds/molecules (e.g., plant-based compounds) that can induce the production of Hsps within the host, as well as confer protection against various abiotic and biotic stresses.

2. Regulation of the Heat Shock Protein Response

Heat shock proteins (Hsps) are evolutionarily ancient and extremely conserved proteins found in almost all living organisms, ranging from archaebacteria, prokaryotes, and eukaryotes. Hsps play an essential role in regulating cellular metabolism in stressful conditions. These proteins are present in several intracellular locations, e.g., in the nucleus, mitochondria, endoplasmic reticulum (ER), and chloroplast and cytosol of eukaryotes [31,32,33]. In general, Hsps are classified into several families based on their function, molecular mass, and sequence homology (i.e., Hsp60, Hsp70, Hsp90, and sHsps) (Figure 1) and also, they can be grouped according to their nearest size family (e.g., Hsp84, Hsp85, and Hsp86 in the Hsp90 kDa family) [34].
Hsps plays a key role in cellular processes that occur during and after exposure to oxidative stress that is caused by hazardous environmental and/or microbial agents. As a result, the normal intracellular reducing environment is compromised, which leads to oxidation and aggregation of key proteins and DNA, ultimately resulting in cellular dysfunction. Due to their versatile functions, Hsps can intervene following oxidative stress at several levels [18,19,35]. Firstly, some Hsps, mainly the Hsp70 family members, play a crucial role in protein sorting and quality control via selecting and directing abnormal proteins to the proteasome or lysosomes for degradation; thus, Hsps aid the clearance of damaged proteins [36]. In some cases, where misfolded proteins need to be rescued, the same machinery facilitates the correct folding of damaged proteins. Moreover, the Hsp families, such as Hsp27, Hsp70, and Hsp90, can negatively regulate apoptosis via the binding and inhibiting of members of the apoptotic cascade. Some Hsps have immuno-enhancing actions (for details, see the section on Hsps and immunity). Among the different Hsp families, Hsp70 is the largest and most highly conserved of the stress protein families. At least 121 proteins have been characterized within this family, and cross-hybridization occurs across various species, such as in mammals, fish, and mollusks [37,38,39].
After the discovery of heat shock proteins, it was clear that the heat shock protein response requires a specific transcription factor [40,41,42,43,44]. In crustaceans, a few reports have suggested that the heat shock transcription factor plays a significant role in determining the heat shock response and also in tolerance against stressful conditions. Tan and Macrae [45] reported that the heat shock factor 1 (Hsf1) transcription factor induces stress tolerance (drying) in diapausing brine shrimp cysts and is responsible for the improved growth and survival of A. franciscana [45]. In another study, Sornchuer et al. [46] demonstrated that Hsf1 has an important role in the thermal stress response and regulates the transcription of heat shock proteins and immune-related genes in P. monodon [46]. However, we still need to carry out detailed investigations and obtain more information in order to characterize the mechanism of heat shock response regulation in crustaceans. Moreover, in eukaryotes, the heat shock response is well documented, and this study shows that the induction of Hsp transcription is mediated by a pre-existing transcription factor, i.e., the heat shock factor or the heat shock transcription factor (HSF) [47]. The HSF upon activation binds to the heat shock element (HSE) at the promoter region (5’ upstream end) of the Hsp gene and induces the transcription of Hsp [48]. The binding motif of the HSE is composed of nGAAn blocks (5 bp) in alternate orientation, and for the stable binding of the HSF and the HSE, at least three units are required [49]. In prokaryotes (e.g., E. coli), the σ32 regulatory protein is responsible for Hsp expression [50]. Additionally, as an alternate subunit of the bacterial RNA polymerase, σ32 replaces the σ70 normal regulatory proteins during heat stress [51]. HSF1 and σ32 share basic mechanistic properties, however their structure or sequence are not related; further, it has been found that protein homeostasis disturbance results in the activation of hsf1 and σ32 [52].
HSF genes have been reported from yeast (Saccharomyces cerevisiae), plants (Arabidopsis sp.), fruit flies (Drosophila sp.), chicken (Gallus sp.), A. franciscana, and P. monodon [53,54,55,56,57]. Sequence comparison of the HSF gene from different species showed that DNA binding and the oligomerization domain is strongly conserved. The HSF contains two highly conserved regions, i.e., a ~100 amino acids NH2-terminal DNA binding domain and an adjacent trimerization domain having 3 hydrophobic heptad repeats, leu zippers [33]. The activation of the HSF involves cellular factors as the intermediary sensors that regulate the activity of the HSF during non-stressful conditions [58]. In animals, the HSF is maintained as a monomeric form through transient interaction with Hsps. However, during stressful conditions, the HSF was released and formed trimers with the HSE, which resulted in the induction of Hsps. Subsequently, Hsps bind with denatured/misfolded protein aggregates and maintain the homeostatic condition in the cells [59,60,61,62,63].
Based on previous studies, a hypothetical illustration of the possible methods involved in the activation of the heat shock protein response, and its potential role in maintaining homeostasis and host health, is depicted in Figure 2. Briefly, the stressful stimuli, heat-shock-protein-inducing (Hspi) conditions (e.g., environmental, pathological, and physiological) [64,65], or compounds (e.g., plant-based, natural polymers, etc.) [66,67] develop oxidative status inside the host. Subsequently, it facilitates the phosphorylation and nuclear translocation of the HSF (in the native form present in the non-phosphorylated form attached with heat shock proteins such as Hsp90) [68] where (inside the hemocyte nucleus) the HSF binds with the HSE at the promoter region (5’ upstream end) of the Hsp gene and induces the transcription of the Hsp gene and production of Hsps [69,70]. The Hsp (e.g., Hsp70) functions as a non-covalently molecular chaperone bind with a hydrophobic exposed segment of unfolded proteins [71,72]. Further, this prevents the aggregation of inappropriate or unfolded proteins; inhibits the misfolding of the polypeptides; transports immature polypeptides to target organelle for the purposes of final packaging and repair, as well as denaturation or degradation of misfolded proteins through proteasomes or lysosomes (proteolysis), which cannot be repaired [73]; and, also, maintains protein homeostasis. The extracellular Hsp also functions as a chaperokine and binds with Toll-like receptors (e.g., TLR-4), which are expressed on hemocytes that lead to the maturation and activation of hemocytes [74,75,76]. The activated hemocytes induce the production of astakines that binds with the binding site of a pathogen in the host cell and decrease the chance of the pathogen attaching to its host cells [77,78,79]. The hemocytes (activated by Hsps) also induce the expression of transglutaminase (Tgase) [80], which form a blood clot in the presence of Ca2+ and prevents the spread of pathogens and increases the expression of antimicrobial peptides (e.g., crustin, lysozyme, etc.) that has bactericidal activity against Gram-positive and Gram-negative bacteria [81,82,83,84]. The p38 (mitogen-activated protein kinase, MAPK) activated by hemocytes, increases the expression of antimicrobial peptides (e.g., crustin, lysozyme, etc.) and has a critical role in defense against bacterial and viral infection [85,86,87,88,89]. Hemocytes activate the ProPO cascade in the presence of Ca2+, leading to melanization and further killing of the pathogens [90,91,92,93].

3. Factors Modulate Heat Shock Protein Response

The induction of Hsps in response to cellular stressors was initially considered a short-term functional response, with a range of essential housekeeping and cytoprotective functions. These stressors, in general, induce protein damage and increase the susceptibility of host animals further to subsequent stressful conditions. However, accumulating pieces of evidence, over the past few decades, have suggested that Hsps play a significant role in the regulation of the immune response in invertebrates (Figure 3 and Table 1). The heat shock protein expressions influenced by either abiotic or biotic stresses are summarized in the section below.

3.1. Environmental or Abiotic Stresses

The physiological status of crustaceans is greatly influenced by their environmental conditions. A slight variation in environmental parameters creates a stressful condition that attenuates the immune system and increases the susceptibility of animals to microbial infection.

3.1.1. Temperature

Temperature is considered an important abiotic stressor, as a slight change in water temperature can affect the body physiology and health of crustaceans [105,106,107]. Interestingly, heat shock proteins are amongst the most significant proteins that are induced by hypo and hyperthermia and their role in protection against thermal stress has been well documented [108,109,110,111,112]. Among the Hsp multigenic family, few proteins are expressed at extremely low levels under normal conditions, while the transcription of most Hsps increased significantly in response to stresses, e.g., stress-inducible proteins (Hsp70). However, Hsps that are expressed constitutively under normal conditions, and may be upregulated during stress conditions, are generally known as heat shock cognate proteins, e.g., Hsc70 [113,114,115,116]. Hsps play a central role in thermotolerance by promoting growth at moderately high temperatures and protecting the organism from mortality at extremely high temperatures [117]. In general, Hsps are induced in both hyperthermia and hypothermia conditions. For example, an increase in the water temperature has been found to induce the production of Hsp70 in Ferropenaeus chinensis, Chinese white shrimp [118], Gammarus pulex, freshwater crustacean [119], and A. franciscana, [120]. Moreover, increased Hsp (Hsp90 and Hsp40) expression was also reported during 6 h of cold shock at 1 and 6 °C in A. franciscana [121]. Hsp70 has been demonstrated to play an important role in protecting cells from damage in S. paramamosain in response to thermal stress (this was shown in an increase in 11 °C from normal growth temperature) [122]. In addition, the adult A. franciscana, when exposed to sub-lethal heat shock (37 °C for 30 min), induces the transcription of Hsp70, Hsp67, and Hsc70, resulting in an improved tolerance of brine shrimp to high temperatures [123]. Interestingly, marine invertebrates are very sensitive to high temperatures and there are a myriad of reports suggesting that Hsp70 is upregulated in response to heat stress [124,125,126]. These studies indicate that Hsps improve thermal tolerance in crustaceans and provide protection in both hyperthermia and hypothermia conditions.

3.1.2. Salinity

The concentration of dissolved inorganic salt concentrations or the salinity in the water is reported to affect the osmoregulation of crustaceans and induce cellular damage, including a deleterious effect on the folding and transformation of polypeptides [127,128,129,130]. Moreover, the osmotic-stress-induced Hsp production plays a crucial role in the maintenance of biological processes, as well as the protection of crustaceans against stressful conditions [131,132,133,134,135]. The study by Yang et al. [79] reported that high salinity stress increased the expression of Hsp70 and that this could lead to enhanced resistance in S. paramamosain against changes in water salinity. The possible mechanism behind the protective action of Hsps against salinity stress is that the osmotic change increases the metabolism rate and enhances the stress response resulting in increased Hsp production [136]. The increased Hsp response subsequently enhances the immune response, including lysozyme, phenoloxidase, and peroxide activity and provides rapid protection against osmotic stress until the organic osmolytes are fully accumulated [137,138]. Water salinity was reported as an important factor for the purposes of natural growth of crustaceans; further, studies on P. trituberculatus have shown that variable salinity significantly influences larval development [139]. Xu and Qin [61] found that Hsp60 has an important role in both the cellular and humoral stress response of the swimming crab, P. trituberculatus, and that these responses regulate the salinity stress via an intrinsic pathway. Further, these responses also play an essential role in protecting the swimming crab against salinity stress. In addition, a few reports have suggested that an increase in the expression of Hsp70 enhances immune response and confers protection to A. franciscana against hypersalinity stress [140].

3.1.3. Environmental Pollutants

The environmental pollutants induced heat shock proteins (Hsps) expression in crustaceans are the most frequently studied in the literature [141,142,143]. The study of heat shock proteins in invertebrates started in the 1990s, and the first observation conducted on Hsps was made by Köhler et al. [144]. The study showed that the exposure of three diplopods (Tachypodoiulus niger, Cylindroiulus punctatus, and Glomeris marginata), one isopod (Oniscus asellus) and two slugs (Arion ater and Deroceras reticulatum) to heavy metals/molluscicides resulted in the increased expression of Hsp70 [145].
Studies on the effect of environmental pollutants on crustaceans demonstrated that the Hsp gene expression is induced by several chemical stresses. For example, this can be found in: nonylphenol (NP) (used in the polymer industry); bisphenol A diglycidyl ether (BPA) (intermediate in the production of polycarbonate and epoxy resins) [146]; 17α-ethynyl estradiol (EE) (synthetic estrogen) [147]; bis(2-ethylhexyl) phthalate (DEHP) (plasticizer in polymer products); endosulfan (ES) (organochlorine insecticide); chloropyriphos (CP) (organophosphorus insecticide); paraquat dichloride (PQ) (oxygen radical generating herbicide); Cadmium (Cd); lead (Pb) and potassium dichromate (Cr) (heavy metals); and benzo[a] pyrene (BaP) (polycyclic aromatic hydrocarbon) [148]. Moreover, for instance, the sublethal concentration of endosulfan has been reported to enhance the synthesis of Hsp70 and Hsp90 in monsoon river prawn, i.e., M. malcolmsonii and P. monodon [149,150,151,152]. In another study, the mixture of environmental pollutant chemicals has been reported to modulate the physiological, as well as immunity and survival responses in crustaceans. For instance, Park and Kwak [89] demonstrated that the application of bisphenol A (BPA) and 4-nonlphenol (NP)—an endocrine disrupting chemical (EDCs)—at different concentrations and at different time intervals (12, 24, 48, and 96 h) induce the expression of Hsp90; further, when exposed to BPA and NP, the marine crab, Charybdis japonica, has a significantly increased survival [54]. Additionally, chemical stress was reported to induce the production of Hsps, which helps in maintaining the homeostasis and structural integrity of cells [153,154].

3.2. Biotic Stresses

The heat shock proteins that are induced by biotic stresses play a very crucial role in protein folding, immune enhancement, and cross-protection against infectious diseases [155,156,157]. Although, there are several reports that demonstrate that Hsps are easily induced by abiotic stresses including heat, salinity, etc. [158], very little information is available on crustaceans’ Hsp response against biotic stresses, including bacteria, parasites, and viruses. Some of the recent research findings have suggested that members of the Hsp70 family have been identified in crustaceans, which are involved in the response to biotic stresses, mainly bacteria, parasites, and viruses [159,160]. Zhou et al. [119] demonstrated that L. vannamei when challenged with V. alginolyticus (Gram-negative) and S. aureus (Gram-positive) bacteria have a significantly increased expression of L. vannamei Hsp60 (LvHSP60) and Hsp70 (LvHSP70) gene in the gills, hepatopancreas, and hemocytes. In another study, temporal transcription of LvHSP70, following the white spot syndrome virus (WSSV) challenge, has been reported to induce an anti-WSSV innate immune response in L. vannamei [161].
In the swimming crab, Portunus trituberculatus, transcription of the P. trituberculatus, the Hsp70 (PtHsp70) gene was shown to increase very rapidly in response to the bacterial challenge with V. alginolyticus [162]. Similar findings were reported by Yang et al. [19], i.e., the fact that there was an increased transcription of Hsp70 in hemocytes of S. paramamosain after the V. alginolyticus challenge and were involved in generating cross-protection in the mud crab. Additionally, Hsp90, which plays a crucial role in protein biosynthesis, signal transduction, and immune responses, was also shown to induce protective immunity in crustaceans. Huang et al. [20] analyzed the role of Hsp90 in S. paramamosain in response to microbial infection, and the results showed that the transcription of SpHSP90 was upregulated in mud crabs after being challenged with Staphylococcus aureus, white spot syndrome virus (wssv), and V. harveyi [20].
Hsps exert their physiological effect via assisting in the formation of new polypeptides as well as in the protecting and maintaining of the host cell polypeptides and naïve proteins from denaturation during microbial infection [163,164,165]. In crustaceans, the host immune response against microbial infection [133] is often associated with reactive oxygen species (ROS) production [166]. It has been demonstrated that in L. vannamei, after bacterial challenge, the subsequent induction of the host immune response leads to an increase in ROS levels [167]. While microbial infection-induced ROS mostly has antimicrobial activity, the production of ROS can result in the denaturation of proteins (proteotoxicity) in the host cell itself; as such, in this condition, the induced Hsps display a cytoprotective role and act as chaperone proteins in order to maintain the protein homeostasis and preserve cellular structures [168].

4. An A. franciscana Model System to Investigate the Role of Hsps in Crustaceans

The A. franciscana is a small branchiopod crustacean that is highly osmotolerant and reported from several harsh environmental conditions worldwide [169,170,171,172]. They live in an environment of severe hypersalinity, high levels of ultraviolet radiation, fluctuating oxygen concentration, and extreme temperature [173,174]. The oviparous development in A. franciscana leads to the production of hard-shell covering diapause cysts, which are composed of stress-tolerant metabolically inactive embryos stalled at gastrulation and that which can remain in stasis for several years [175,176,177,178]. However, when the diapause cysts were immersed under appropriate conditions in seawater with aeration and temperatures, the hard-shell raptures and cysts develop, releasing swimming larvae within 24 h (Figure 4) [178,179,180].
Apart from its interesting life history, A. franciscana are non-selective filter feeders (which can be grown in a wide range of feed resources), have a rapid generation cycle (the cyst grows to adult in 20–30 days), require very low space for growth (hence have a relatively smaller cost to culture), and developmental stages are well characterized. This aspect of their stages being well characterized, for instance, is shown in the fact that the A. franciscana produces encysted gastrulae cysts during oviparous development, while ovoviviparous development provides live larvae in both sexual and asexual (parthenogenetic) stages. Additionally, gnotobiotic (germ-free) culture conditions (allowing full control over the host-associated microbial communities) and advanced molecular techniques including qPCR and RNAi are well established in A. franciscana, which makes this species an exceptional model organism that can be used in order to investigate the host–pathogen relationship and to study the biological activity of protective compounds [181,182,183,184,185,186,187,188,189,190]. In addition, cysts of A. franciscana can be stored for a couple of years in the fridge and, after terminating the diapause stage, the cysts from different generations can be used and hatched all together, simultaneously. This excellent facility of storage and hatching, permits one to perform experiments on demand, which can help to avoid or minimize environmental influences. Above all, the genome sequence of A. franciscana showed that it shares a very high homology with shrimps and other crustaceans’ genomes (Figure 5). Therefore, there is a high possibility that the outcome of studies based on A. franciscana would provide a fundamental basis to understand the host–pathogen interactions in other commercially important shrimp species.
The induction of Hsp production inside the host in order to control diseases in aquaculture was investigated using the model organism, brine shrimp larvae [191,192]. In 1988, Miller and McLennan observed the presence of heat shock proteins in the early developmental stages of brine shrimp, i.e., in encysted gastrula embryos (cysts) and newly hatched nauplius larvae. They have reported that the larvae exhibited induced thermotolerance, which is associated with the synthesis and upregulation of heat shock proteins [193]. The role of heat shock protein response in adult A. franciscana, in response to high temperature (including LT50 determination, enhanced thermotolerance, and increased production of the Hsp70 family stress protein) was studied by Frankenberg et al. [95]. Results demonstrated that Hsp70 family proteins (mainly Hsp67 and Hsc70) levels were significantly upregulated during sublethal heat shock (37 °C for 30 min). The A. franciscana exposed to Zn-control/Cd-control treatment induced the expression of Hsp and increased the partitioning of Cd to the tropically available metal (TAM), which could result in the bio-enhancement of Cd trophic transfer to predators, which, in turn, leads to the suppression of Zn accumulation in A. franciscana [194].
Later, a gnotobiotic (germ-free) culture system was developed for brine shrimp and several studies have demonstrated that this provides a fully controlled and excellent host–pathogen environment and facilitates the determining of the effect of external stimuli on the host (Figure 5) [195]. The gnotobiotic system also avoids the interference generated by host-associated microorganisms as well as shifts in the composition of microbial diversity [196,197,198,199,200]. The effect of non-lethal heat shock (NLHS) on host survival, immune response, and protection against stressful conditions were also studied by several researchers. Results showed that NLHS increased the transcription of Hsp70, which enhances the immune response and provides cross-protection from environmental, physiological, and microbial stress in A. franciscana [201,202,203,204,205,206]. In another study, Sung et al. [64] demonstrated that non-pathogenic CAG 629 and CAG 626 E. coli strains, when heat shocked and administered through feed to Artemia larvae, resulted in enhanced Hsp70 expression (~two folds) and subsequent protection against V. campbellii infection [207]. In another study, the bacterial strains GR 8 (Cytophaga sp.), LVS 3 (Aeromonas hydrophila), LVS 2 (Bacillus sp.), LVS 8 (Vibrio sp.), and GR 10 (Roseobacter sp.) overproducing DnaK were reported to protect and improve gnotobiotic Artemia resistance to V. campbellii infection [208]. Baruah et al. [23] demonstrated that a heat shock protein inducer, Tex-OE®, enhances the production of Hsp70 and increased the survival of A. franciscana nauplii against thermal and salinity stresses. Further, it was, therefore, concluded that the protective effect of Tex-OE® is mediated by an enhanced production of Hsp70 [209]. In addition, the production of heat shock protein 70 (Hsp70) in A. franciscana in response to non-lethal heat shock (30 min exposure to 37 °C) increased the tolerance of brine shrimp to zinc and cadmium metal exposure [210]. Later, several studies investigated the molecular chaperone activity of Hsps in A. franciscana through an in vivo RNA interference (RNAi) technique [211,212]. The results showed that the injection of dsRNA to A. franciscana (before fertilization) resulted in a complete knockdown of the expression of small heat shock proteins (i.e., p26) and reduced the resistance of cysts to desiccation and freezing [213,214]. In another study, the knockdown of dsRNA Hsp70 was reported to downregulate the expression of Hsp70 mRNA, thereby resulting in an reduced survival of shrimp and brine shrimp larvae against pathogenic V. campbellii and AHPND-causing V. parahaemolyticus challenges [215,216,217]. Taken together, these studies confirm that Hsps possibly play important roles in enhanced protection in crustaceans against stressful conditions by stabilizing and refolding unfolded or denatured proteins, as well as in enhancing their innate defense system.

5. Heat Shock Protein, a Promising Candidate to Enhance Immunity and Prevent Diseases in Crustaceans

The role of Hsps, produced by both prokaryotic and eukaryotic organisms, in eliciting immune responses and inducing resistance to diseases has been well established in various animal and human models. However, in crustaceans, the studies into the effects of Hsps in generating protective immunity against infection stress are accumulating [218,219,220]. Based on these accumulating pieces of evidence, it can be suggested that Hsps are potential health-beneficial biomolecules that could be targeted in order to develop a disease-control strategy in aquaculture animals.
The Hsp70 family comprises the most well-characterized Hsps. The induction of Hsp expression could enhance crustacean immunity as shown in other animals that are exposed to heat stress. Shrimp (L. vannamei) that are exposed to chronic NLHS showed a higher expression of LvHSP70, LvHSP90, and immune-related genes (i.e., LvproPO and LvCrustin [221,222,223,224,225]). The high expression of LvproPO and hemocyanin was also observed in shrimp that were exposed to acute NLHS. Moreover, shrimp exposed to either acute or chronic NLHS had a higher survival rate than that of the non-heated shrimp control when they were challenged with VPAHPND. In addition, in regard to bacterial infection, the exposure of shrimp to NLHS could reduce WSSV infection as shown by the decreased viral copy number and the decreased cumulative mortality of WSSV-infected L. vannamei. In Artemia, it was also reported that NLHS could protect the animals against deleterious bacterial challenges. Furthermore, the functions of HSP70, HSC70, and HSC70-5 have been demonstrated to use the RNA interference (RNAi) technique. Iryani et al. [36] used RNAi to verify the role of HSP70 in protecting the nauplii of A. franciscana against abiotic and biotic stressors. The survival of nauplii lacking HSP70, compared with that of those with a functional HSP70, was decreased by 41% during heat stress and 34% upon V. campbellii infection. These results suggest that Hsps plays an important role in maintaining protein homeostasis by functioning as a molecular chaperone, while enhancing the host innate immune response against bacterial infection [226,227,228,229,230]. The effect of Hsps in regulating immune-related gene expression in crustaceans was not only investigated by suppression of HSP gene expression, but also by the injection of recombinant Hsps. In L. vannamei, injected with the recombinant DnaK followed by V. campbellii challenge, the transcript expression of immune genes TGase-1 and LvproPO-2 and endogenous HSP70 (LvHSP70) were clearly affected. Similarly, P. monodon, which was first injected with DnaK and then injected 1 h later with V. harveyi, showed a significant increase in proPO transcript expression. The study involving LvHSP70 injection showed that HSP70 could induce several immune pathways in the shrimp. Direct injection of rLvHSP70 into shrimp muscle demonstrated that rLvHSP70 enters hemocyte cells and localizes to both the cytoplasm and nucleus, while also accumulating in the plasma membrane. Thus, both gene knockdown and recombinant protein injection elicited similar results, suggesting a novel mechanism underlying the role of LvHSP70 in the activation of the shrimp immune system. Moreover, feeding A. franciscana with E. coli producing ArHsp70 or DnaK proteins showed a high survival rate in a Vibrio challenge assay. The observed effects could be due to the enhancement of the Artemia immune system as phenoloxidase activity was found to be increased by these proteins [231,232,233,234].
The Hsp family that is of main interest for disease control is HSP70, however the sHsps of HSP90 and HSP60, as well as the co-chaperone HSP40, appear to ameliorate infection by pathogens as well. sHsps provide oligomeric platforms for the ATP-independent binding of structurally perturbed proteins, preventing their irreversible denaturation when cells are stressed. HSP90, HSP70, and HSP60 are stress-induced and they have the ability to protect proteins from irreversible denaturation. However, the major function of these chaperone families is to bind and fold nascent proteins through ATP-driven allosteric rearrangement, although the molecular structure and mechanism of action for each chaperone differ. Hsps function cooperatively by forming intracellular networks of chaperones, co-chaperones, and accessory proteins. sHsp monomers, consisting of a conserved α-crystallin domain flanked by an amino-terminal sequence and a carboxyl-terminal extension [235,236,237,238], assemble into oligomers. The α-crystallin domain contributes to the dimerization of monomers and substrate binding, activities that depend on the amino- and carboxyl-terminal regions for the greatest efficiency [239,240,241]. Further, sHsp oligomers either disassemble or undergo structural rearrangement during stress, increasing surface hydrophobicity and enhancing reactions with substrate proteins [242,243,244,245,246]. Proteins released from sHsps when the stress passes, either refold spontaneously or refold with the assistance of an ATP-dependent Hsp, such as HSP70. The primary role of sHsps during exposure to stress, including infection, is to protect proteins from irreversible denaturation.
Under acute thermal stress in the culture environment, it was found that the white shrimp HSP40 (LvHSP40) transcript levels were significantly induced in muscle, gill, and the hepatopancreas. The expression profiles of four HSP genes (LvHSP60, LvHSP70, LvHSC70, and LvHSP90) of L. vannamei were significantly induced, and the transcription level of LvHSP70 was the most sensitive to temperature fluctuations. Moreover, HSP60 and HSP10 genes from Scrippsiella trochoidea were rapidly upregulated upon exposure to both low and high temperatures [247,248,249,250]. In pathological situations, such as necrotic cell death, Hsps can be released into the extracellular environment with cellular proteins in order to induce autoimmunity by receptor-mediated activation of the innate immune response. In shrimp, there is evidence that the Hsps are highly expressed in response to pathogen infection. The LvHsp60 protein from L. vannamei was significantly upregulated in the gills, hepatopancreas, and in the hemocytes after being challenged with either Gram-positive and Gram-negative bacteria.
In addition, it has been shown that using a plant-based polyphenolic compound, phloroglucinol, induces HSP70 production and protects the brine shrimp A. franciscana and freshwater prawn M. rosenbergii against bacterial infection. Treatment of the brine shrimp with phloroglucinol was shown to result in significant upregulation of DSCAM, proPO, Peroxinectin, HSP90, and HSP70 and downregulation of LGBP. Moreover, the phenolic compound carvacrol also induces the expression of HSP72 in A. franciscana. Further, the induction of HSP72 enhances Artemia larvae tolerance to lethal heat stress or pathogenic V. harveyi. Similarly, treating A. franciscana with pyrogallol at an optimum concentration could induce protective effects against V. harveyi infection [40,41,178,179,180,181,182]. Taken together, these findings suggest that Hsps possibly contribute to immune defense against infections by regulating the immune system in crustaceans.

6. Conclusions and Future Perspective

Aquaculture is a rapidly growing food-producing sector. The sector has grown at an average rate of 8.9% per year since 1970, compared to only 1.2% for capture fisheries and 2.8% for terrestrial farmed meat-production systems over the same period. However, the intensive development of the aquaculture industry has been accompanied by an increase in environmental impacts. The production process generates substantial amounts of polluted effluent, containing uneaten feed, and feces. Discharges from aquaculture into the aquatic environments contain nutrients, as well as various organic and inorganic compounds such as ammonium, phosphorus, dissolved organic carbon, and organic matter. The high levels of nutrients cause environmental deterioration of the receiving water bodies. In addition, the drained water may increase the occurrence of pathogenic microorganisms and introduce invading pathogen species. Additionally, the degraded aquatic environment makes the animal more susceptible to disease outbreaks. For instance, over the last decade of disease outbreaks, in particular, bacterial diseases have brought socio-economic and environmental unsustainability to the crustacean aquaculture industry. As estimated by the FAO, the economic losses from a disease outbreak in the aquaculture industry are over USD 9 billion per year, which is approximately 15% of the value of world-farmed fish and shellfish production. The traditional methods applied so far in the mitigation of stressful conditions, such as disinfectants and antibiotics, have had very limited success. Among the various drugs, antibiotics are the most popular and widely used chemotherapeutic agent used to control bacterial disease in crustacean aquaculture. However, the wide use of antibiotics, and especially the application at subtherapeutic doses, can eliminate both pathogenic and beneficial bacteria and inevitably results in the development of resistant strains of antibiotics that can create environmental problems in the ecosystems. In addition, the presence of antibiotic residues in commercialized products of aquaculture constitutes an additional problem for human health, as this can generate problems of allergy, toxicity, and results in the alteration of the human gut’s normal microflora. Hence, the development of natural products or plant-derived compounds is needed in order to mitigate stressful conditions in aquatic environments and enhance the immune reactivity of crustaceans.
Interestingly, to avoid the problems generated by environmental change, crustaceans possess their own specific behavioral and physiological adaptive mechanism, i.e., the enhanced heat shock protein response, that plays a significant role by stabilizing the protein structure and function, as well as generating protective immunity against abiotic and biotic stresses. However, when the adverse environmental condition is prolonged, it surpasses the natural defense capacity and affects growth, survival, metabolism, reproduction, and immunity. Hence, the management strategies, e.g., the use of heat shock protein-inducing (Hspi) compounds may offer a promising viable and sustainable means to maintain homeostasis in aquatic animals and avoid integrative and or multiple stresses. Natural products from medicinal plants and marine seaweeds are considered potential alternatives for the prevention of stressful conditions in crustacean aquaculture. Plant-based compounds are identified in order to possess the characteristic of enhancing the heat shock protein within the animal in a non-invasive manner. These compounds/molecules are also commonly called heat shock protein inducers (Hspi). Heat shock proteins, especially Hsp70, are reported to be functionally involved in providing cross-protection in crustacean species. For instance, the upregulation of Hsp70 production in shrimps as a general stress response protects the animals from subsequent secondary heterologous environmental and physiological insults. In general, the molecular chaperone activity—which maintains protein homeostasis by protecting the nascent polypeptides from misfolding, assisting in the assembly and disassembly of macromolecular complexes, as well as facilitating co- and post-translational folding and regulating translocation—is documented as a protective function of Hsp70. Additionally, Hsp70 is also reported to induce thermotolerance, protect against osmotic stress, prevent oxidative toxicity and damage, and improve tolerance against microbial infection. These observations clearly illustrated that Hsps play a significant role in host immunity and health. In conclusion, we can say that the development of a compound or molecule that could have a possible application as a Hsp inducer would be a holistic approach that could be used to generate tolerance in the crustacean species against subsequent deleterious environmental stresses. In addition, it may be a suitable candidate for use as an anti-stress agent in crustacean aquaculture.

Author Contributions

Conceptualization, V.K. and S.R.; methodology, V.K.; validation, V.K. and S.R.; writing—original draft preparation, V.K. and S.R.; writing—review and editing, B.K.B. and B.K.D.; visualization, V.K.; supervision, B.K.B. and B.K.D.; project administration, B.K.B. and B.K.D.; funding acquisition, B.K.B. and B.K.D. All authors have read and agreed to the published version of the manuscript.

Funding

This research received no external funding.

Acknowledgments

Authors are thankful to ICAR-Central Inland Fisheries Research Institute (ICAR-CIFRI) and other technical and supporting staff for the financial and technical support.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. FAO. State of Fisheries and Aquaculture in the World; FAO: Rome, Italy, 2019; ISBN 9781424464968. [Google Scholar]
  2. Tacon, A.G.J.; Lemos, D.; Metian, M. Fish for Health: Improved Nutritional Quality of Cultured Fish for Human Consumption. Rev. Fish. Sci. Aquac. 2020, 28, 449–458. [Google Scholar] [CrossRef]
  3. Roy, S.; Kumar, V.; Mitra, A.; Manna, R.K.; Suresh, V.R.; Homechaudhuri, S. Amylase and protease activity in shrimps and prawn of sundarbans, West bengal, India. Indian J. Geo-Mar. Sci. 2018, 47, 53–59. [Google Scholar]
  4. FAO. State of Fisheries and Aquaculture in the World; FAO: Rome, Italy, 2018; ISBN 978-92-5-130562-1. [Google Scholar]
  5. Robert, J.; Jancovich, J.K. Recombinant Ranaviruses for Studying Evolution of Host–Pathogen Interactions in Ectothermic Vertebrates. Viruses 2016, 8, 187. [Google Scholar] [CrossRef] [Green Version]
  6. Soltani, M.; Ghosh, K.; Hoseinifar, S.H.; Kumar, V.; Lymbery, A.J.; Roy, S.; Ringø, E. Genus Bacillus, promising probiotics in aquaculture: Aquatic animal origin, bio-active components, bioremediation and efficacy in fish and shellfish. Rev. Fish. Sci. Aquac. 2019, 27, 331–379. [Google Scholar] [CrossRef] [Green Version]
  7. Defoirdt, T.; Boon, N.; Bossier, P.; Verstraete, W. Disruption of bacterial quorum sensing: An unexplored strategy to fight infections in aquaculture. Aquaculture 2004, 240, 69–88. [Google Scholar] [CrossRef]
  8. Kumar, V.; Roy, S.; Meena, D.K.; Sarkar, U.K. Application of Probiotics in Shrimp Aquaculture: Importance, Mechanisms of Action, and Methods of Administration. Rev. Fish. Sci. Aquac. 2016, 24, 342–368. [Google Scholar] [CrossRef]
  9. Roy, S.; Kumar, V.; Bossier, P.; Norouzitallab, P.; Vanrompay, D. Phloroglucinol Treatment Induces Transgenerational Epigenetic Inherited Resistance against Vibrio Infections and Thermal Stress in a Brine Shrimp (Artemia franciscana) Model. Front. Immunol. 2019, 10, 2745. [Google Scholar] [CrossRef] [Green Version]
  10. Coombes, J.L.; Robey, E.A. Dynamic imaging of host-pathogen interactions in vivo. Nat. Rev. Immunol. 2010, 10, 353–364. [Google Scholar] [CrossRef]
  11. Kumar, V.; Roy, S.; Baruah, K.; Van Haver, D.; Impens, F.; Bossier, P. Environmental conditions steer phenotypic switching in acute hepatopancreatic necrosis disease-causing Vibrio parahaemolyticus, affecting PirAVP/PirB VP toxins production. Environ. Microbiol. 2020, 22, 4212–4230. [Google Scholar] [CrossRef]
  12. Smith, V.J.; Chisholm, J.R. Non-cellular immunity in crustaceans. Fish Shellfish Immunol. 1992, 2, 1–31. [Google Scholar] [CrossRef]
  13. Söderhäll, K.; Cerenius, L. Role of the prophenoloxidase-activating system in invertebrate immunity. Curr. Opin. Immunol. 1998, 10, 23–28. [Google Scholar] [CrossRef]
  14. Sritunyalucksana, K.; Söderhäll, K. The proPO and clotting system in crustaceans. Aquaculture 2000, 191, 53–69. [Google Scholar] [CrossRef]
  15. Roy, S.; Bossier, P.; Norouzitallab, P.; Vanrompay, D. Trained immunity and perspectives for shrimp aquaculture. Rev. Aquac. 2020, 12, 2351–2370. [Google Scholar] [CrossRef]
  16. Norouzitallab, P.; Baruah, K.; Vanrompay, D.; Bossier, P. Teaching Shrimps Self-Defense to Fight Infections. Trends Biotechnol. 2018, 37, 16–19. [Google Scholar] [CrossRef]
  17. Churchward, G. The two faces of Janus: Virulence gene regulation by CovR/S in group A streptococci. Mol. Microbiol. 2007, 64, 34–41. [Google Scholar] [CrossRef]
  18. Dong, H.; Zheng, X.; Kumar, V.; Roy, S.; Duan, Y.; Gao, H.; Zhang, J. Dietary supplementation of teprenone potentiates thermal and hypoxia tolerance as well as cellular stress protection of Epinephelus coioides juveniles reared under multiple stressors. Aquaculture 2020, 514, 734413. [Google Scholar] [CrossRef]
  19. Yang, Y.; Ye, H.; Huang, H.; Li, S.; Liu, X.; Zeng, X.; Gong, J. Expression of Hsp70 in the mud crab, Scylla paramamosain in response to bacterial, osmotic, and thermal stress. Cell Stress Chaperones 2013, 18, 475–482. [Google Scholar] [CrossRef] [Green Version]
  20. Huang, A.-M.; Geng, Y.; Wang, K.-Y.; Zeng, F.; Liu, Q.; Wang, Y.; Sun, Y.; Liu, X.-X.; Zhou, Y. Molecular Cloning and Expression Analysis of Heat Shock Protein 90 (Hsp90) of the Mud Crab, Scylla Paramamosain. J. Agric. Sci. 2013, 5, 1. [Google Scholar] [CrossRef] [Green Version]
  21. Sung, Y.Y.; Pineda, C.; MacRae, T.H.; Sorgeloos, P.; Bossier, P. Exposure of gnotobiotic Artemia franciscana larvae to abiotic stress promotes heat shock protein 70 synthesis and enhances resistance to pathogenic Vibrio campbellii. Cell Stress Chaperones 2008, 13, 59–66. [Google Scholar] [CrossRef] [Green Version]
  22. Baruah, K.; Ranjan, J.; Sorgeloos, P.; Bossier, P. Efficacy of heterologous and homologous heat shock protein 70s as protective agents to Artemia franciscana challenged with Vibrio campbellii. Fish Shellfish Immunol. 2010, 29, 733–739. [Google Scholar] [CrossRef] [Green Version]
  23. Baruah, K.; Norouzitallab, P.; Roberts, R.J.; Sorgeloos, P.; Bossier, P. A novel heat-shock protein inducer triggers heat shock protein 70 production and protects Artemia franciscana nauplii against abiotic stressors. Aquaculture 2012, 334–337, 152–158. [Google Scholar] [CrossRef]
  24. Sung, Y.Y.; Van Damme, E.J.; Sorgeloos, P.; Bossier, P. Non-lethal heat shock protects gnotobiotic Artemia franciscana larvae against virulent Vibrios. Fish Shellfish Immunol. 2007, 22, 318–326. [Google Scholar] [CrossRef]
  25. Baruah, K.; Phong, H.P.P.D.; Norouzitallab, P.; Defoirdt, T.; Bossier, P. The gnotobiotic brine shrimp (Artemia franciscana) model system reveals that the phenolic compound pyrogallol protects against infection through its prooxidant activity. Free Radic. Biol. Med. 2015, 89, 593–601. [Google Scholar] [CrossRef]
  26. Baruah, K.; Huy, T.T.; Norouzitallab, P.; Niu, Y.; Gupta, S.K.; De Schryver, P.; Bossier, P. Probing the protective mechanism of poly-ß-hydroxybutyrate against vibriosis by using gnotobiotic Artemia franciscana and Vibrio campbellii as host-pathogen model. Sci. Rep. 2015, 5, 9427. [Google Scholar] [CrossRef] [Green Version]
  27. Baruah, K.; Norouzitallab, P.; Linayati, L.; Sorgeloos, P.; Bossier, P. Reactive oxygen species generated by a heat shock protein (Hsp) inducing product contributes to Hsp70 production and Hsp70-mediated protective immunity in Artemia franciscana against pathogenic vibrios. Dev. Comp. Immunol. 2014, 46, 470–479. [Google Scholar] [CrossRef]
  28. Niu, Y.; Norouzitallab, P.; Baruah, K.; Dong, S.; Bossier, P. A plant-based heat shock protein inducing compound modulates host–pathogen interactions between Artemia franciscana and Vibrio campbellii. Aquaculture 2014, 430, 120–127. [Google Scholar] [CrossRef]
  29. Roy, S. Modulating Innate Immune Memory in Brine Shrimp (Artemia franciscana) and in Giant Freshwater Prawn (Macrobrachium rosenbergii). Ph.D. Thesis, University of Ghent, Ghent, Belgium, 2020. [Google Scholar]
  30. Kumar, V. Acute Hepatopancreatic Necrosis Disease (Ahpnd) in Shrimp: Virulence, Pathogenesis and Mitigation Strategies. Ph.D. Thesis, University of Ghent, Ghent, Belgium, 2020. [Google Scholar]
  31. Nover, L.; Scharf, K.D.; Neumann, D. Cytoplasmic heat shock granules are formed from precursor particles and are associated with a specific set of mRNAs. Mol. Cell. Biol. 1989, 9, 1298–1308. [Google Scholar] [CrossRef]
  32. Parsell, D.A.; Lindquist, S. The Function of Heat-Shock Proteins in Stress Tolerance: Degradation and Reactivation of Damaged Proteins. Annu. Rev. Genet. 1993, 27, 437–496. [Google Scholar] [CrossRef]
  33. Craig, E.A.; Weissman, J.S.; Horwich, A.L. Heat shock proteins and molecular chaperones: Mediators of protein conformation and turnover in the cell. Cell 1994, 78, 365–372. [Google Scholar] [CrossRef]
  34. Dong, H.; Roy, S.; Zheng, X.; Kumar, V.; Das, B.K.; Duan, Y.; Sun, Y.; Zhang, J. Dietary teprenone enhances non-specific immunity, antioxidative response and resistance to hypoxia induced oxidative stress in Lateolabrax maculatus. Aquaculture 2021, 533, 736126. [Google Scholar] [CrossRef]
  35. Lindquist, S.; Craig, E.A. The Heat-Shock Proteins. Annu. Rev. Genet. 1988, 22, 631–677. [Google Scholar] [CrossRef] [PubMed]
  36. Iryani, M.T.M.; MacRae, T.H.; Panchakshari, S.; Tan, J.; Bossier, P.; Wahid, M.E.A.; Sung, Y.Y. Knockdown of heat shock protein 70 (Hsp70) by RNAi reduces the tolerance of Artemia franciscana nauplii to heat and bacterial infection. J. Exp. Mar. Biol. Ecol. 2017, 487, 106–112. [Google Scholar] [CrossRef]
  37. Roberts, R.J.; Agius, C.; Saliba, C.; Bossier, P.; Sung, Y.Y. Heat shock proteins (chaperones) in fish and shellfish and their potential role in relation to fish health: A review. J. Fish Dis. 2010, 33, 789–801. [Google Scholar] [CrossRef] [PubMed]
  38. Sung, Y.Y.; Macrae, T.H.; Sorgeloos, P.; Bossier, P. Stress response for disease control in aquaculture. Rev. Aquac. 2011, 3, 120–137. [Google Scholar] [CrossRef]
  39. Robert, J. Evolution of heat shock protein and immunity. Dev. Comp. Immunol. 2003, 27, 449–464. [Google Scholar] [CrossRef]
  40. Norouzitallab, P.; Baruah, K.; Muthappa, D.M.; Bossier, P. Non-lethal heat shock induces HSP70 and HMGB1 protein production sequentially to protect Artemia franciscana against Vibrio campbellii. Fish Shellfish Immunol. 2015, 42, 395–399. [Google Scholar] [CrossRef]
  41. Norouzitallab, P.; Baruah, K.; Biswas, P.; Vanrompay, D.; Bossier, P. Probing the phenomenon of trained immunity in invertebrates during a transgenerational study, using brine shrimp Artemia as a model system. Sci. Rep. 2016, 6, 21166. [Google Scholar] [CrossRef] [Green Version]
  42. Lindquist, S. The Heat-Shock Response. Annu. Rev. Biochem. 1986, 55, 1151–1191. [Google Scholar] [CrossRef]
  43. Richter, K.; Haslbeck, M.; Buchner, J. The Heat Shock Response: Life on the Verge of Death. Mol. Cell 2010, 40, 253–266. [Google Scholar] [CrossRef]
  44. Courgeon, A.-M.; Maisonhaute, C.; Best-Belpomme, M. Heat shock proteins are induced by cadmium in Drosophila cells. Exp. Cell Res. 1984, 153, 515–521. [Google Scholar] [CrossRef]
  45. Tan, J.; Macrae, T.H. Stress tolerance in diapausing embryos of Artemia franciscana is dependent on heat shock factor 1 (Hsf1). PLoS ONE 2018, 13, e0200153. [Google Scholar] [CrossRef] [PubMed]
  46. Sornchuer, P.; Junprung, W.; Yingsunthonwattana, W.; Tassanakajon, A. Heat shock factor 1 regulates heat shock proteins and immune-related genes in Penaeus monodon under thermal stress. Dev. Comp. Immunol. 2018, 88, 19–27. [Google Scholar] [CrossRef] [PubMed]
  47. Yura, T.; Tobe, T.; Ito, K.; Osawa, T. Heat shock regulatory gene (htpR) of Escherichia coli is required for growth at high temperature but is dispensable at low temperature. Proc. Natl. Acad. Sci. USA 1984, 81, 6803–6807. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  48. Michel, G.P.; Starka, J. Effect of ethanol and heat stresses on the protein pattern of Zymomonas mobilis. J. Bacteriol. 1986, 165, 1040–1042. [Google Scholar] [CrossRef] [Green Version]
  49. Ananthan, J.; Goldberg, A.L.; Voellmy, R. Abnormal Proteins Serve as Eukaryotic Stress Signals and Trigger the Activation of Heat Shock Genes. Science 1986, 232, 522–524. [Google Scholar] [CrossRef]
  50. Csermely, P.; Schnaider, T.; So, C.; Prohászka, Z.; Nardai, G. The 90-kDa Molecular Chaperone Family: Structure, function, and clinical applications. a comprehensive review. Pharmacol. Ther. 1998, 79, 129–168. [Google Scholar] [CrossRef]
  51. Bozaykut, P.; Ozer, N.K.; Karademir, B. Regulation of protein turnover by heat shock proteins. Free Radic. Biol. Med. 2014, 77, 195–209. [Google Scholar] [CrossRef]
  52. Aridon, P.; Geraci, F.; Turturici, G.; D’Amelio, M.; Savettieri, G.; Sconzo, G. Protective Role of Heat Shock Proteins in Parkinson’s Disease. Neurodegener. Dis. 2011, 8, 155–168. [Google Scholar] [CrossRef]
  53. Capy, P.; Gasperi, G.; Biémont, C.; Bazin, C. Stress and transposable elements: Co-evolution or useful parasites? Heredity 2000, 85, 101–106. [Google Scholar] [CrossRef] [Green Version]
  54. Horowitz, A.; Horowitz, S. Disease control in shrimp aquaculture from a microbial ecology perspective. In The New Wave, Proceedings of the Special Session on Sustainable Shrimp Culture, Aquaculture; Browdy, C.L., Jory, D.E., Eds.; World Aquaculture Society, 2001; pp. 199–218. Available online: https://ag.arizona.edu/azaqua/tilapia/tilapia_shrimp/moriarty.PDF (accessed on 27 August 2022).
  55. De La Vega, E.; Degnan, B.; Hall, M.R.; Cowley, J.A.; Wilson, K.J. Quantitative real-time RT-PCR demonstrates that handling stress can lead to rapid increases of gill-associated virus (GAV) infection levels in Penaeus monodon. Dis. Aquat. Org. 2004, 59, 195–203. [Google Scholar] [CrossRef] [Green Version]
  56. de la Vega, E.; Hall, M.R.; Degnan, B.M.; Wilson, K.J. Short-term hyperthermic treatment of Penaeus monodon increases expression of heat shock protein 70 (HSP70) and reduces replication of gill associated virus (GAV). Aquaculture 2006, 253, 82–90. [Google Scholar] [CrossRef]
  57. Vidal, O.M.; Granja, C.B.; Aranguren, F.; Brock, J.A.; Salazar, M. A Profound Effect of Hyperthermia on Survival of Litopenaeus vannamei Juveniles Infected with White Spot Syndrome Virus. J. World Aquac. Soc. 2001, 32, 364–372. [Google Scholar] [CrossRef]
  58. Ellis, R.J.; Van der Vies, S.M. Molecular chaperones. Annu. Rev. Biochem. 1991, 60, 321–347. [Google Scholar] [CrossRef]
  59. Tomanek, L.; Somero, G. Evolutionary and acclimation-induced variation in the heat-shock responses of congeneric marine snails (genus Tegula) from different thermal habitats: Implications for limits of thermotolerance and biogeography. J. Exp. Biol. 1999, 202, 2925–2936. [Google Scholar] [CrossRef] [PubMed]
  60. Tomanek, L. Variation in the heat shock response and its implication for predicting the effect of global climate change on species’ biogeographical distribution ranges and metabolic costs. J. Exp. Biol. 2010, 213, 971–979. [Google Scholar] [CrossRef] [Green Version]
  61. Xu, Q.; Qin, Y. Molecular cloning of heat shock protein 60 (PtHSP60) from Portunus trituberculatus and its expression response to salinity stress. Cell Stress Chaperones 2012, 17, 589–601. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  62. Baruah, K.; Ranjan, J.; Sorgeloos, P.; MacRae, T.H.; Bossier, P. Priming the prophenoloxidase system of Artemia franciscana by heat shock proteins protects against Vibrio campbellii challenge. Fish Shellfish Immunol. 2011, 31, 134–141. [Google Scholar] [CrossRef]
  63. Zhu, Y.; Zhu, G.; Guo, Q.; Zhu, Z.; Wang, C.; Liu, Z. A Comparative Proteomic Analysis of Pinellia ternata Leaves Exposed to Heat Stress. Int. J. Mol. Sci. 2013, 14, 20614–20634. [Google Scholar] [CrossRef] [Green Version]
  64. Sung, Y.Y.; Dhaene, T.; Defoirdt, T.; Boon, N.; MacRae, T.H.; Sorgeloos, P.; Bossier, P. Ingestion of bacteria overproducing DnaK attenuates Vibrio infection of Artemia franciscana larvae. Cell Stress Chaperones 2009, 14, 603–609. [Google Scholar] [CrossRef] [Green Version]
  65. Jakob, U.; Muse, W.; Eser, M.; Bardwell, J.C.A. Chaperone Activity with a Redox Switch. Cell 1999, 96, 341–352. [Google Scholar] [CrossRef] [Green Version]
  66. Kumsta, C.; Jakob, U. Redox-Regulated Chaperones. Biochemistry 2009, 48, 4666–4676. [Google Scholar] [CrossRef] [PubMed]
  67. Sharma, S.; Chakraborty, K.; Müller, B.K.; Astola, N.; Tang, Y.-C.; Lamb, D.C.; Hayer-Hartl, M.; Hartl, F.U. Monitoring Protein Conformation along the Pathway of Chaperonin-Assisted Folding. Cell 2008, 133, 142–153. [Google Scholar] [CrossRef] [Green Version]
  68. Walter, S.; Buchner, J. Molecular chaperones—Cellular machines for protein folding. Angew. Chem. Int. Ed. 2002, 41, 1098–1113. [Google Scholar] [CrossRef]
  69. Mayer, M.P.; Bukau, B. Hsp70 chaperones: Cellular functions and molecular mechanism. Cell. Mol. Life Sci. 2005, 62, 670. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  70. Flaherty, K.M.; McKay, D.B.; Kabsch, W.; Holmes, K.C. Similarity of the three-dimensional structures of actin and the ATPase fragment of a 70-kDa heat shock cognate protein. Proc. Natl. Acad. Sci. USA 1991, 88, 5041–5045. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  71. Lo, J.; Hayashi, M.; Woo-Kim, S.; Tian, B.; Huang, J.; Fearns, C.; Takayama, S.; Zapata, J.M.; Yang, Y.; Lee, J. Tid1, a co-chaperone of the heat shock 70 protein and the mammalian counterpart of the Drosophila tumor suppressor l(2)tid, is critical for early embryonic development and cell survival. Mol. Cell. Biol. 2004, 24, 2226–2236. [Google Scholar] [CrossRef] [Green Version]
  72. Flaherty, K.M.; DeLuca-Flaherty, C.; McKay, D.B. Three-dimensional structure of the ATPase fragment of a 70K heat-shock cognate protein. Nature 1990, 346, 623–628. [Google Scholar] [CrossRef]
  73. Kiang, J.G. Heat Shock Protein 70 kDa Molecular Biology, Biochemistry, and Physiology. Pharmacol. Ther. 1998, 80, 183–201. [Google Scholar] [CrossRef]
  74. Kregel, K.C. Heat shock proteins: Modifying factors in physiological stress responses and acquired thermotolerance. J. Appl. Physiol. 2002, 92, 2177–2186. [Google Scholar] [CrossRef] [Green Version]
  75. Cottin, D.; Shillito, B.; Chertemps, T.; Thatje, S.; Léger, N.; Ravaux, J. Comparison of heat-shock responses between the hydrothermal vent shrimp Rimicaris exoculata and the related coastal shrimp Palaemonetes varians. J. Exp. Mar. Biol. Ecol. 2010, 393, 9–16. [Google Scholar] [CrossRef] [Green Version]
  76. Cottin, D.; Roussel, D.; Foucreau, N.; Hervant, F.; Piscart, C. Disentangling the effects of local and regional factors on the thermal tolerance of freshwater crustaceans. Die Nat. 2012, 99, 259–264. [Google Scholar] [CrossRef] [PubMed]
  77. Yost, H.; Lindquist, S. RNA splicing is interrupted by heat shock and is rescued by heat shock protein synthesis. Cell 1986, 45, 185–193. [Google Scholar] [CrossRef]
  78. Huang, L.-H.; Wang, C.-Z.; Kang, L. Cloning and expression of five heat shock protein genes in relation to cold hardening and development in the leafminer, Liriomyza sativa. J. Insect Physiol. 2009, 55, 279–285. [Google Scholar] [CrossRef]
  79. Yang, J.; Mu, Y.; Dong, S.; Jiang, Q.; Yang, J. Changes in the expression of four heat shock proteins during the aging process in Brachionus calyciflorus (rotifera). Cell Stress Chaperones 2013, 19, 33–52. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  80. Hartl, F.U.; Martin, J. Molecular chaperones in cellular protein folding. Bioessays 1995, 16, 689–692. [Google Scholar] [CrossRef] [PubMed]
  81. Bukau, B.; Horwich, A.L. The Hsp70 and Hsp60 Chaperone Machines. Cell 1998, 92, 351–366. [Google Scholar] [CrossRef] [Green Version]
  82. Hartl, F.U.; Hayer-Hartl, M. Molecular Chaperones in the Cytosol: From Nascent Chain to Folded Protein. Science 2002, 295, 1852–1858. [Google Scholar] [CrossRef] [Green Version]
  83. Beckmann, R.P.; Mizzen, L.E.; Welch, W.J. Interaction of Hsp 70 with Newly Synthesized Proteins: Implications for Protein Folding and Assembly. Science 1990, 248, 850–854. [Google Scholar] [CrossRef]
  84. Palleros, D.R.; Welch, W.J.; Fink, A.L. Interaction of hsp70 with unfolded proteins: Effects of temperature and nucleotides on the kinetics of binding. Proc. Natl. Acad. Sci. USA 1991, 88, 5719–5723. [Google Scholar] [CrossRef] [Green Version]
  85. Sadis, S.; Hightower, L.E. Unfolded proteins stimulate molecular chaperone Hsc70 ATPase by accelerating ADP/ATP exchange. Biochemistry 1992, 31, 9406–9412. [Google Scholar] [CrossRef]
  86. Frydman, J.; Nimmesgern, E.; Ohtsuka, K.; Hartl, F.U. Folding of nascent polypeptide chains in a high molecular mass assembly with molecular chaperones. Nature 1994, 370, 111–117. [Google Scholar] [CrossRef] [PubMed]
  87. Gething, M.-J.; Sambrook, J. Protein folding in the cell. Nature 1992, 335, 33–45. [Google Scholar] [CrossRef] [PubMed]
  88. Chen, Z.; Zhou, T.; Wu, X.; Hong, Y.; Fan, Z.; Li, H. Influence of cytoplasmic heat shock protein 70 on viral infection of Nicotiana benthamiana. Mol. Plant Pathol. 2008, 9, 809–817. [Google Scholar] [CrossRef]
  89. Park, H.; Lee, J.; Huh, S.; Seo, J.; Choi, E. Hsp72 functions as a natural inhibitory protein of c-Jun N-terminal kinase. EMBO J. 2001, 20, 446–456. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  90. Dong, Y.; Dong, S. Induced thermotolerance and expression of heat shock protein 70 in sea cucumber Apostichopus japonicus. Fish. Sci. 2008, 74, 573–578. [Google Scholar] [CrossRef]
  91. Wu, G.; Harris, M.K.; Guo, J.-Y.; Wan, F.-H. Response of multiple generations of beet armyworm, Spodoptera exigua (Hübner), feeding on transgenic Bt cotton. J. Appl. Èntomol. 2009, 133, 90–100. [Google Scholar] [CrossRef]
  92. Tsan, M.-F.; Gao, B. Heat shock protein and innate immunity. Cell. Mol. Immunol. 2004, 1, 274–279. [Google Scholar]
  93. Jolesch, A.; Elmer, K.; Bendz, H.; Issels, R.D.; Noessner, E. Hsp70, a messenger from hyperthermia for the immune system. Eur. J. Cell Biol. 2012, 91, 48–52. [Google Scholar] [CrossRef] [Green Version]
  94. Bedulina, D.S.; Evgen’Ev, M.B.; Timofeyev, M.A.; Protopopova, M.V.; Garbuz, D.G.; Pavlichenko, V.V.; Luckenbach, T.; Shatilina, Z.M.; Axenov-Gribanov, D.V.; Gurkov, A.N.; et al. Expression patterns and organization of the hsp70 genes correlate with thermotolerance in two congener endemic amphipod species (Eulimnogammarus cyaneus and E. verrucosus) from Lake Baikal. Mol. Ecol. 2013, 22, 1416–1430. [Google Scholar] [CrossRef]
  95. Frankenberg, M.; Jackson, S.; Clegg, J. The heat shock response of adult Artemia franciscana. J. Therm. Biol. 2000, 25, 481–490. [Google Scholar] [CrossRef]
  96. Liu, J.; Yang, W.-J.; Zhu, X.-J.; Karouna-Renier, N.K.; Rao, R.K. Molecular cloning and expression of two HSP70 genes in the prawn, Macrobrachium rosenbergii. Cell Stress Chaperones 2004, 9, 313–323. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  97. Chang, E.S.; Chang, S.A.; Keller, R.; Reddy, P.S.; Snyder, J.; Spees, J.L.; Chang, E.S.; Chang, S.A.; Keller, R.; Reddy, P.S.; et al. Quantification of Stress in Lobsters: Crustacean Hyperglycemic Hormone, Stress Proteins, and Gene Expression. Am. Zool. 1999, 39, 487–495. [Google Scholar] [CrossRef]
  98. Tedengren, M.; Olsson, B.; Reimer, O.; Brown, D.C.; Bradley, B.P. Heat pretreatment increases cadmium resistance and HSP 70 levels in Baltic Sea mussels. Aquat. Toxicol. 2000, 48, 1–12. [Google Scholar] [CrossRef]
  99. Baruah, K.; Norouzitallab, P.; Shihao, L.; Sorgeloos, P.; Bossier, P. Feeding truncated heat shock protein 70s protect Artemia franciscana against virulent Vibrio campbellii challenge. Fish Shellfish Immunol. 2013, 34, 183–191. [Google Scholar] [CrossRef]
  100. Kumar, V.; Baruah, K.; Nguyen, D.V.; Smagghe, G.; Vossen, E.; Bossier, P. Phloroglucinol-Mediated Hsp70 Production in Crustaceans: Protection against Vibrio parahaemolyticus in Artemia franciscana and Macrobrachium rosenbergii. Front. Immunol. 2018, 9, 1091. [Google Scholar] [CrossRef] [Green Version]
  101. Junprung, W.; Supungul, P.; Tassanakajon, A. HSP70 and HSP90 are involved in shrimp Penaeus vannamei tolerance to AHPND-causing strain of Vibrio parahaemolyticus after non-lethal heat shock. Fish Shellfish Immunol. 2017, 60, 237–246. [Google Scholar] [CrossRef]
  102. Spees, J.L.; Chang, S.A.; Snyder, M.J.; Chang, E.S. Osmotic Induction of Stress-Responsive Gene Expression in the LobsterHomarus americanus. Biol. Bull. 2002, 203, 331–337. [Google Scholar] [CrossRef]
  103. Yan, L.; Cerny, R.L.; Cirillo, J.D. Evidence that hsp90 Is Involved in the Altered Interactions of Acanthamoeba castellanii Variants with Bacteria. Eukaryot. Cell 2004, 3, 567–578. [Google Scholar] [CrossRef] [Green Version]
  104. Holmes, J.L.; Sharp, S.Y.; Hobbs, S.; Workman, P. Silencing of HSP90 Cochaperone AHA1 Expression Decreases Client Protein Activation and Increases Cellular Sensitivity to the HSP90 Inhibitor 17-Allylamino-17-Demethoxygeldanamycin. Cancer Res. 2008, 68, 1188–1197. [Google Scholar] [CrossRef] [Green Version]
  105. Welch, W.J.; Feramisco, J.R. Purification of the major mammalian heat shock proteins. J. Biol. Chem. 1982, 257, 14949–14959. [Google Scholar] [CrossRef]
  106. Aligue, R.; Akhavan-Niak, H.; Russell, P. A role for Hsp90 in cell cycle control: Wee1 tyrosine kinase activity requires inter-action with Hsp90. EMBO J. 1994, 13, 6099–6106. [Google Scholar] [CrossRef] [PubMed]
  107. Jakob, U.; Lilie, H.; Meyer, I.; Buchner, J. Transient interaction of Hsp90 with early unfolding intermediates of citrate synthase: Implications for heat shock in vivo. J. Biol. Chem. 1995, 270, 7288–7294. [Google Scholar] [CrossRef] [PubMed]
  108. Pearl, L.H.; Prodromou, C. Structure and Mechanism of the Hsp90 Molecular Chaperone Machinery. Annu. Rev. Biochem. 2006, 75, 271–294. [Google Scholar] [CrossRef] [PubMed]
  109. Reddy, P.S.; Thirulogachandar, V.; Vaishnavi, C.; Aakrati, A.; Sopory, S.K.; Reddy, M.K. Molecular characterization and expression of a gene encoding cytosolic Hsp90 from Pennisetum glaucum and its role in abiotic stress adaptation. Gene 2011, 474, 29–38. [Google Scholar] [CrossRef] [PubMed]
  110. Fu, D.; Chen, J.; Zhang, Y.; Yu, Z. Cloning and expression of a heat shock protein (HSP) 90 gene in the haemocytes of Crassostrea hongkongensis under osmotic stress and bacterial challenge. Fish Shellfish Immunol. 2011, 31, 118–125. [Google Scholar] [CrossRef]
  111. Xu, D.; Sun, L.; Liu, S.; Zhang, L.; Yang, H. Polymorphisms of heat shock protein 90 (Hsp90) in the sea cucumber Apostichopus japonicus and their association with heat-resistance. Fish Shellfish Immunol. 2014, 41, 428–436. [Google Scholar] [CrossRef]
  112. Zhao, H.; Yang, H.; Zhao, H.; Chen, M.; Wang, T. The molecular characterization and expression of heat shock protein 90 (Hsp90) and 26 (Hsp26) cDNAs in sea cucumber (Apostichopus japonicus). Cell Stress Chaperones 2011, 16, 481–493. [Google Scholar] [CrossRef] [Green Version]
  113. Zhu, J.-Y.; Wu, G.-X.; Ye, G.-Y.; Hu, C. Heat shock protein genes (hsp20, hsp75 and hsp90) from Pieris rapae: Molecular cloning and transcription in response to parasitization by Pteromalus puparum. Insect Sci. 2012, 20, 183–193. [Google Scholar] [CrossRef]
  114. Quintana, F.J.; Cohen, I.R. Heat Shock Proteins as Endogenous Adjuvants in Sterile and Septic Inflammation. J. Immunol. 2005, 175, 2777–2782. [Google Scholar] [CrossRef] [Green Version]
  115. Vabulas, R.M.; Ahmad-Nejad, P.; da Costa, C.; Miethke, T.; Kirschning, C.J.; Häcker, H.; Wagner, H. Endocytosed HSP60s Use Toll-like Receptor 2 (TLR2) and TLR4 to Activate the Toll/Interleukin-1 Receptor Signaling Pathway in Innate Immune Cells. J. Biol. Chem. 2001, 276, 31332–31339. [Google Scholar] [CrossRef] [Green Version]
  116. Choresh, O.; Ron, E.; Loya, Y. The 60-kDa Heat Shock Protein (HSP60) of the Sea Anemone Anemonia viridis: A Potential Early Warning System for Environmental Changes. Mar. Biotechnol. 2001, 3, 501–508. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  117. Choresh, O.; Loya, Y.; Müller, W.E.; Wiedenmann, J.; Azem, A. The mitochondrial 60-kDa heat shock protein in marine invertebrates: Biochemical purification and molecular characterization. Cell Stress Chaperones 2004, 9, 38–48. [Google Scholar] [CrossRef]
  118. Clayton, M.E.; Steinmann, R.; Fent, K. Different expression patterns of heat shock proteins hsp 60 and hsp 70 in zebra mussels (Dreissena polymorpha) exposed to copper and tributyltin. Aquat. Toxicol. 2000, 47, 213–226. [Google Scholar] [CrossRef]
  119. Zhou, J.; Wang, W.-N.; He, W.-Y.; Zheng, Y.; Wang, L.; Xin, Y.; Liu, Y.; Wang, A.-L. Expression of HSP60 and HSP70 in white shrimp, Litopenaeus vannamei in response to bacterial challenge. J. Invertebr. Pathol. 2010, 103, 170–178. [Google Scholar] [CrossRef]
  120. Huang, W.-J.; Leu, J.-H.; Tsau, M.-T.; Chen, J.-C.; Chen, L.-L. Differential expression of LvHSP60 in shrimp in response to environmental stress. Fish Shellfish Immunol. 2011, 30, 576–582. [Google Scholar] [CrossRef]
  121. Sun, Y.; Macrae, T.H. Small heat shock proteins: Molecular structure and chaperone function. Cell. Mol. Life Sci. 2005, 62, 2460–2476. [Google Scholar] [CrossRef]
  122. Mchaourab, H.S.; Godar, J.A.; Stewart, P.L. Structure and Mechanism of Protein Stability Sensors: Chaperone Activity of Small Heat Shock Proteins. Biochemistry 2009, 48, 3828–3837. [Google Scholar] [CrossRef] [Green Version]
  123. Laganowsky, A.; Benesch, J.; Landau, M.; Ding, L.; Sawaya, M.; Cascio, D.; Huang, Q.; Robinson, C.; Horwitz, J.; Eisenberg, D. Crystal structures of truncated alphaA and alphaB crystallins reveal structural mechanisms of polydispersity important for eye lens function. Protein Sci. 2010, 19, 1031–1043. [Google Scholar] [CrossRef] [Green Version]
  124. Hilario, E.; Martin, F.J.M.; Bertolini, M.C.; Fan, L. Crystal Structures of Xanthomonas Small Heat Shock Protein Provide a Structural Basis for an Active Molecular Chaperone Oligomer. J. Mol. Biol. 2011, 408, 74–86. [Google Scholar] [CrossRef] [Green Version]
  125. Waters, E.R. The evolution, function, structure, and expression of the plant sHSPs. J. Exp. Bot. 2013, 64, 391–403. [Google Scholar] [CrossRef] [Green Version]
  126. Kriehuber, T.; Rattei, T.; Weinmaier, T.; Bepperling, A.; Haslbeck, M.; Buchner, J. Independent evolution of the core domain and its flanking sequences in small heat shock proteins. FASEB J. 2010, 24, 3633–3642. [Google Scholar] [CrossRef] [PubMed]
  127. Van Montfort, R.L.; Basha, E.; Friedrich, K.L.; Slingsby, C.; Vierling, E. Crystal structure and assembly of a eukaryotic small heat shock protein. Nat. Genet. 2001, 8, 1025–1030. [Google Scholar] [CrossRef]
  128. Horwitz, J. Alpha-crystallin. Exp. Eye Res. 2003, 76, 145–153. [Google Scholar] [CrossRef]
  129. Haslbeck, M.; Franzmann, T.; Weinfurtner, D.; Buchner, J. Some like it hot: The structure and function of small heat-shock proteins. Nat. Struct. Mol. Biol. 2005, 12, 842–846. [Google Scholar] [CrossRef]
  130. Mogk, A.; Deuerling, E.; Vorderwülbecke, S.; Vierling, E.; Bukau, B. Small heat shock proteins, ClpB and the DnaK system form a functional triade in reversing protein aggregation. Mol. Microbiol. 2003, 50, 585–595. [Google Scholar] [CrossRef] [Green Version]
  131. Lee, G.J.; Roseman, A.M.; Saibil, H.R.; Vierling, E. A small heat shock protein stably binds heat-denatured model substrates and can maintain a substrate in a folding-competent state. EMBO J. 1997, 16, 659–671. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  132. Cashikar, A.G.; Duennwald, M.; Lindquist, S.L. A Chaperone Pathway in Protein Disaggregation: Hsp26 alters the nature of protein aggregates to facilitate reactivation by Hsp104s. J. Biol. Chem. 2005, 280, 23869–23875. [Google Scholar] [CrossRef] [Green Version]
  133. Liberek, K.; Lewandowska, A.; Ziętkiewicz, S. Chaperones in control of protein disaggregation. EMBO J. 2008, 27, 328–335. [Google Scholar] [CrossRef] [Green Version]
  134. Jackson, S.A.; Clegg, J.S. Ontogeny of low molecular weight stress protein p26 during early embryogenesis of the brine shrimp, Artemia franciscana. Dev. Growth Differ. 1996, 38, 153–160. [Google Scholar] [CrossRef]
  135. Clegg, J.S. Stress-related proteins compared in diapause and in activated, anoxic encysted embryos of the animal extremophile, Artemia franciscana. J. Insect Physiol. 2011, 57, 660–664. [Google Scholar] [CrossRef]
  136. King, A.M.; Macrae, T.H. The Small Heat Shock Protein p26 Aids Development of Encysting Artemia Embryos, Prevents Spontaneous Diapause Termination and Protects against Stress. PLoS ONE 2012, 7, e43723. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  137. King, A.M.; Toxopeus, J.; MacRae, T.H. Functional differentiation of small heat shock proteins in diapause-destined Artemia embryos. FEBS J. 2013, 280, 4761–4772. [Google Scholar] [CrossRef] [PubMed]
  138. Chen, T.; Villeneuve, T.S.; Garant, K.A.; Amons, R.; MacRae, T.H. Functional characterization of artemin, a ferritin homolog synthesized in Artemia embryos during encystment and diapause. FEBS J. 2007, 274, 1093–1101. [Google Scholar] [CrossRef]
  139. Hu, Y.; Bojikova-Fournier, S.; King, A.M.; MacRae, T.H. The structural stability and chaperone activity of artemin, a ferritin homologue from diapause-destined Artemia embryos, depend on different cysteine residues. Cell Stress Chaperones 2010, 16, 133–141. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  140. King, A.M.; Toxopeus, J.; MacRae, T.H. Artemin, a Diapause-Specific Chaperone, Contributes to the Stress Tolerance of Artemia franciscana Cysts and Influences Their Release from Females. J. Exp. Biol. 2014, 217, 1719–1724. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  141. Macrae, T.H. Stress tolerance during diapause and quiescence of the brine shrimp, Artemia. Cell Stress Chaperones 2015, 21, 9–18. [Google Scholar] [CrossRef] [Green Version]
  142. Dai, L.; Chen, D.-F.; Liu, Y.-L.; Zhao, Y.; Yang, F.; Yang, J.-S.; Yang, W.-J. Extracellular Matrix Peptides of Artemia Cyst Shell Participate in Protecting Encysted Embryos from Extreme Environments. PLoS ONE 2011, 6, e20187. [Google Scholar] [CrossRef]
  143. Wu, B.J.; Kingston, R.E.; Morimoto, R.I. Human HSP70 promoter contains at least two distinct regulatory domains. Proc. Natl. Acad. Sci. USA 1986, 83, 629–633. [Google Scholar] [CrossRef] [Green Version]
  144. Köhler, H.R.; Triebskorn, R.; Stöcker, W.; Kloetzel, P.-M.; Alberti, G. The 70 kD heat shock protein (hsp 70) in soil invertebrates: A possible tool for monitoring environmental toxicants. Arch. Environ. Contam. Toxicol. 1992, 22, 334–338. [Google Scholar] [CrossRef]
  145. Viswanathan, C.; Khanna-Chopra, R. Heat shock proteins-Role in thermotolerance of crop plants. Curr. Sci. 1996, 71, 275–284. [Google Scholar]
  146. Morimoto, R.I. Cells in Stress: Transcriptional Activation of Heat Shock Genes. Science 1993, 259, 1409–1410. [Google Scholar] [CrossRef] [PubMed]
  147. Amin, J.; Ananthan, J.; Voellmy, R. Key features of heat shock regulatory elements. Mol. Cell. Biol. 1988, 8, 3761–3769. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  148. Grossman, A.D.; Erickson, J.W.; Gross, C.A. The htpR gene product of E. coli is a sigma factor for heat-shock promoters. Cell 1984, 38, 383–390. [Google Scholar] [CrossRef]
  149. Sorger, P.K.; Pelham, H.R. Yeast heat shock factor is an essential DNA-binding protein that exhibits temperature-dependent phosphorylation. Cell 1988, 54, 855–864. [Google Scholar] [CrossRef]
  150. Kong, X.-C.; Zhang, D.; Qian, C.; Liu, G.-T.; Bao, X.-Q. FLZ, a novel HSP27 and HSP70 inducer, protects SH-SY5Y cells from apoptosis caused by MPP+. Brain Res. 2011, 1383, 99–107. [Google Scholar] [CrossRef] [PubMed]
  151. Fang, H.; Wu, Y.; Huang, X.; Wang, W.; Ang, B.; Cao, X.; Wan, T. Toll-like Receptor 4 (TLR4) Is Essential for Hsp70-like Protein 1 (HSP70L1) to Activate Dendritic Cells and Induce Th1 Response. J. Biol. Chem. 2011, 286, 30393–30400. [Google Scholar] [CrossRef] [Green Version]
  152. Triantafilou, M.; Triantafilou, K. Heat-shock protein 70 and heat-shock protein 90 associate with Toll-like receptor 4 in response to bacterial lipopolysaccharide. Biochem. Soc. Trans. 2004, 32, 636–639. [Google Scholar] [CrossRef] [Green Version]
  153. Sato, S.; Fujita, N.; Tsuruo, T. Modulation of Akt kinase activity by binding to Hsp90. Proc. Natl. Acad. Sci. USA 2000, 97, 10832–10837. [Google Scholar] [CrossRef] [Green Version]
  154. Lin, X.; Söderhäll, I. Crustacean hematopoiesis and the astakine cytokines. Blood 2011, 117, 6417–6424. [Google Scholar] [CrossRef] [Green Version]
  155. Maningas, M.B.B.; Kondo, H.; Hirono, I.; Saito-Taki, T.; Aoki, T. Essential function of transglutaminase and clotting protein in shrimp immunity. Mol. Immunol. 2008, 45, 1269–1275. [Google Scholar] [CrossRef]
  156. Fagutao, F.F.; Maningas, M.B.B.; Kondo, H.; Aoki, T.; Hirono, I. Transglutaminase regulates immune-related genes in shrimp. Fish Shellfish Immunol. 2012, 32, 711–715. [Google Scholar] [CrossRef]
  157. Zhu, Y.-T.; Li, D.; Zhang, X.; Li, X.-J.; Li, W.-W.; Wang, Q. Role of transglutaminase in immune defense against bacterial pathogens via regulation of antimicrobial peptides. Dev. Comp. Immunol. 2016, 55, 39–50. [Google Scholar] [CrossRef] [PubMed]
  158. Yu, Z.; Geng, Y.; Huang, A.; Wang, K.; Huang, X.; Chen, D.; Ou, Y.; Wang, J. Molecular characterization of a p38 mitogen-activated protein kinase gene from Scylla paramamosain and its expression profiles during pathogenic challenge. J. Invertebr. Pathol. 2017, 144, 32–36. [Google Scholar] [CrossRef] [PubMed]
  159. He, S.; Qian, Z.; Yang, J.; Wang, X.; Mi, X.; Liu, Y.; Hou, F.; Liu, Q.; Liu, X. Molecular characterization of a p38 MAPK from Litopenaeus vannamei and its expression during the molt cycle and following pathogen infection. Dev. Comp. Immunol. 2013, 41, 217–221. [Google Scholar] [CrossRef] [PubMed]
  160. Yan, H.; Zhang, S.; Li, C.-Z.; Chen, Y.-H.; Chen, Y.-G.; Weng, S.-P.; He, J.-G. Molecular characterization and function of a p38 MAPK gene from Litopenaeus vannamei. Fish Shellfish Immunol. 2013, 34, 1421–1431. [Google Scholar] [CrossRef] [PubMed]
  161. Sanders, B.M. Stress Proteins in Aquatic Organisms: An Environmental Perspective. Crit. Rev. Toxicol. 1993, 23, 49–75. [Google Scholar] [CrossRef] [PubMed]
  162. Feder, M.E.; Hofmann, G.E. Heat-Shock Proteins, Molecular Chaperones, and The Stress Response: Evolutionary and Ecological Physiology. Annu. Rev. Physiol. 1999, 61, 243–282. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  163. Junprung, W.; Norouzitallab, P.; De Vos, S.; Tassanakajon, A.; Viet, D.N.; Van Stappen, G.; Bossier, P. Sequence and expression analysis of HSP70 family genes in Artemia franciscana. Sci. Rep. 2019, 9, 8391. [Google Scholar] [CrossRef] [Green Version]
  164. Ravi, V.; Kubofcik, J.; Bandopathyaya, S.; Geetha, M.; Narayanan, R.; Nutman, T.; Kaliraj, P. Wuchereria bancrofti: Cloning and characterization of heat shock protein 70 from the human lymphatic filarial parasite. Exp. Parasitol. 2004, 106, 1–10. [Google Scholar] [CrossRef]
  165. Luan, W.; Li, F.; Zhang, J.; Wen, R.; Li, Y.; Xiang, J. Identification of a novel inducible cytosolic Hsp70 gene in Chinese shrimp Fenneropenaeus chinensis and comparison of its expression with the cognate Hsc70 under different stresses. Cell Stress Chaperones 2009, 15, 83–93. [Google Scholar] [CrossRef] [Green Version]
  166. Gbotsyo, Y.A. The Effect of Cold Stress on Heat Shock Proteins in Nauplii (larvae) of the Brine Shrimp, Artemia franciscana. Honours Thesis, Saint Mary’s University, Halifax, NS, Canada, 2017; pp. 1–39. [Google Scholar]
  167. Chapple, J.; Smerdon, G.R.; Berry, R.; Hawkins, A.J. Seasonal changes in stress-70 protein levels reflect thermal tolerance in the marine bivalve Mytilus edulis L. J. Exp. Mar. Biol. Ecol. 1998, 229, 53–68. [Google Scholar] [CrossRef]
  168. Browne, R.; Bowen, S. Taxonomy and population genetics of Artemia. In Artemia Biology; Browne, R.A., Sorgeloos, P., Trotman, C.N.A., Eds.; CRC Press: Boca Raton, FL, USA, 1991; pp. 221–235. [Google Scholar]
  169. Triantaphyllidis, G.; Abatzopoulos, T.; Sorgeloos, P. Review of the biogeography of the genus Artemia (Crustacea, Anostraca). J. Biogeogr. 1998, 25, 213–226. [Google Scholar] [CrossRef]
  170. Sorgeloos, P.; Bossuyt, E.; Laviña, E.; Baeza-Mesa, M.; Persoone, G. Decapsulation of Artemia cysts: A simple technique for the improvement of the use of brine shrimp in aquaculture. Aquaculture 1977, 12, 311–315. [Google Scholar] [CrossRef]
  171. Kumar, V.; Das, B.K.; Swain, H.S.; Chowdhury, H.; Roy, S.; Bera, A.K.; Das, R.; Parida, S.N.; Dhar, S.; Jana, A.K.; et al. Outbreak of Ichthyophthirius multifiliis associated with Aeromonas hydrophila in Pangasianodon hypophthalmus: The role of turmeric oil in enhancing immunity and inducing resistance against co-infection. Front. Immunol. 2022, 13, 956478. [Google Scholar] [CrossRef]
  172. Clegg, J.S.; Drinkwater, L.E.; Sorgeloos, P. The Metabolic Status of Diapause Embryos of Artemia franciscana (SFB). Physiol. Zool. 1996, 69, 49–66. [Google Scholar] [CrossRef]
  173. Tran, P.T.N.; Kumar, V.; Bossier, P. Do acute hepatopancreatic necrosis disease-causing PirABVP toxins aggravate vibriosis? Emerg. Microbes Infect. 2020, 9, 1919–1932. [Google Scholar] [CrossRef]
  174. Criel, G.R.J.; Macrae, T.H. Artemia morphology and structure. In Artemia: Basic and Applied Biology. Biology of Aquatic Organisms; Abatzopoulos, T.J., Beardmore, J.A., Clegg, J.S., Sorgeloos, P., Eds.; Springer: Dordrecht, The Netherlands, 2002; pp. 1–37. ISBN 978-94-017-0791-6. [Google Scholar]
  175. Van Stappen, G. Zoogeography. In Artemia: Basic and Applied Biology. Biology of Aquatic Organisms; Abatzopoulos, T.J., Beardmore, J.A., Clegg, J.S., Sorgeloos, P., Eds.; Springer: Dordrecht, The Netherlands, 2002; pp. 171–224. [Google Scholar]
  176. Robbins, H.M.; Van Stappen, G.; Sorgeloos, P.; Sung, Y.Y.; MacRae, T.H.; Bossier, P. Diapause termination and development of encysted Artemia embryos: Roles for nitric oxide and hydrogen peroxide. J. Exp. Biol. 2010, 213, 1464–1470. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  177. Kumar, V.; De Bels, L.; Couck, L.; Baruah, K.; Bossier, P.; Broeck, W.V.D. PirABVP Toxin Binds to Epithelial Cells of the Digestive Tract and Produce Pathognomonic AHPND Lesions in Germ-Free Brine Shrimp. Toxins 2019, 11, 717. [Google Scholar] [CrossRef] [Green Version]
  178. Han, B.; Kaur, V.I.; Baruah, K.; Nguyen, V.D.; Bossier, P. High doses of sodium ascorbate act as a prooxidant and protect gnotobiotic brine shrimp larvae (Artemia franciscana) against Vibrio harveyi infection coinciding with heat shock protein 70 activation. Dev. Comp. Immunol. 2018, 92, 69–76. [Google Scholar] [CrossRef]
  179. Kumar, V.; Nguyen, D.V.; Baruah, K.; Bossier, P. Probing the mechanism of VPAHPND extracellular proteins toxicity purified from Vibrio parahaemolyticus AHPND strain in germ-free Artemia test system. Aquaculture 2019, 504, 414–419. [Google Scholar] [CrossRef]
  180. Kumar, V.; Baruah, K.; Bossier, P. Bamboo powder protects gnotobiotically-grown brine shrimp against AHPND-causing Vibrio parahaemolyticus strains by cessation of PirABVP toxin secretion. Aquaculture 2021, 539, 736624. [Google Scholar] [CrossRef]
  181. De Vos, S.; Bossier, P.; Vuylsteke, M. Genomic Tools and Sex Determination in the Extremophile Brine Shrimp Artemia franciscana; Ghent University: Ghent, Belgium, 2014. [Google Scholar]
  182. Defoirdt, T.; Crab, R.; Wood, T.K.; Sorgeloos, P.; Verstraete, W.; Bossier, P. Quorum Sensing-Disrupting Brominated Furanones Protect the Gnotobiotic Brine Shrimp Artemia franciscana from Pathogenic Vibrio harveyi, Vibrio campbellii, and Vibrio parahaemolyticus Isolates. Appl. Environ. Microbiol. 2006, 72, 6419–6423. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  183. Crab, R.; Lambert, A.; Defoirdt, T.; Bossier, P.; Verstraete, W. The application of bioflocs technology to protect brine shrimp (Artemia franciscana) from pathogenic Vibrio harveyi. J. Appl. Microbiol. 2010, 109, 1643–1649. [Google Scholar] [CrossRef] [PubMed]
  184. Miller, D.; McLennan, A.G. The heat shock response of the cryptobiotic brine Shrimp Artemia—I. A comparison of the thermotolerance of cysts and larvae. J. Therm. Biol. 1988, 13, 119–123. [Google Scholar] [CrossRef]
  185. Seebaugh, D.R.; Wallace, W. Importance of metal-binding proteins in the partitioning of Cd and Zn as trophically available metal (TAM) in the brine shrimp Artemia franciscana. Mar. Ecol. Prog. Ser. 2004, 272, 215–230. [Google Scholar] [CrossRef] [Green Version]
  186. Marques, A.; Ollevier, F.; Verstraete, W.; Sorgeloos, P.; Bossier, P. Gnotobiotically grown aquatic animals: Opportunities to investigate host-microbe interactions. J. Appl. Microbiol. 2006, 100, 903–918. [Google Scholar] [CrossRef]
  187. El-Magsodi, M.O.; Bossier, P.; Sorgeloos, P.; Van Stappen, G. Effect of Light Colour, Timing, and Duration of Light Exposure on the Hatchability of Artemia Spp. (Branchiopoda: Anostraca). Eggs. J. Crustac. Biol. 2016, 36, 515–524. [Google Scholar] [CrossRef] [Green Version]
  188. Clegg, J.S.; Jackson, S.A.; Van Hoa, N.; Sorgeloos, P. Thermal resistance, developmental rate and heat shock proteins in Artemia franciscana, from San Francisco Bay and southern Vietnam. J. Exp. Mar. Biol. Ecol. 2000, 252, 85–96. [Google Scholar] [CrossRef]
  189. Kumar, V.; Bossier, P. Importance of plant—Derived compounds and/or natural products in aquaculture. Aquafeed 2018, 10, 28–31. [Google Scholar]
  190. Kumar, V.; Roy, S. Aquaculture Drugs: Sources, Active Ingredients, Pharmaceutic Preparations and Methods of Administration. J. Aquac. Res. Dev. 2017, 8, 510. [Google Scholar] [CrossRef]
  191. Kumar, V.; Bossier, P. Novel plant-based compounds could be useful in protecting shrimp species against AHPND Vibrio parahaemolyticus. J. Inland Fish. Soc. India 2019, 51, 3–5. [Google Scholar]
  192. Kumar, V.; Wille, M.; Lourenço, T.M.; Bossier, P. Biofloc-Based Enhanced Survival of Litopenaeus vannamei Upon AHPND-Causing Vibrio parahaemolyticus Challenge Is Partially Mediated by Reduced Expression of Its Virulence Genes. Front. Microbiol. 2020, 11, 1270. [Google Scholar] [CrossRef] [PubMed]
  193. Baruah, K.; Norouzitallab, P.; Phong, H.P.P.D.; Smagghe, G.; Bossier, P. Enhanced resistance against Vibrio harveyi infection by carvacrol and its association with the induction of heat shock protein 72 in gnotobiotic Artemia franciscana. Cell Stress Chaperones 2017, 22, 377–387. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  194. Wandinger, S.K.; Richter, K.; Buchner, J. The Hsp90 Chaperone Machinery. J. Biol. Chem. 2008, 283, 18473–18477. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  195. Vos, M.; Hageman, J.; Carra, S.; Kampinga, H.H. Structural and Functional Diversities between Members of the Human HSPB, HSPH, HSPA, and DNAJ Chaperone Families. Biochemistry 2008, 47, 7001–7011. [Google Scholar] [CrossRef]
  196. Kumar, V.; Roy, S.; Behera, B.K.; Swain, H.S.; Das, B.K. Biofloc Microbiome With Bioremediation and Health Benefits. Front. Microbiol. 2021, 12, 3499. [Google Scholar] [CrossRef]
  197. Roy, S.; Baruah, K.; Bossier, P.; Vanrompay, D.; Norouzitallab, P. Induction of transgenerational innate immune memory against Vibrio infections in a brine shrimp (Artemia franciscana) model. Aquaculture 2022, 557, 738309. [Google Scholar] [CrossRef]
  198. Roy, S.; Kumar, V.; Behera, B.K.; Parhi, J.; Mohapatra, S.; Chakraborty, T.; Das, B.K. CRISPR/Cas Genome Edit-ing-Can It Become a Game Changer in Future Fisheries Sector? Front. Mar. Sci. 2022, 9, 924475. [Google Scholar] [CrossRef]
  199. Meimaridou, E.; Gooljar, S.B.; Chapple, J.P. From hatching to dispatching: The multiple cellular roles of the Hsp70 molecular chaperone machinery. J. Mol. Endocrinol. 2008, 42, 1–9. [Google Scholar] [CrossRef] [Green Version]
  200. Vabulas, R.M.; Raychaudhuri, S.; Hayer-Hartl, M.; Hartl, F.U. Protein Folding in the Cytoplasm and the Heat Shock Response. Cold Spring Harb. Perspect. Biol. 2010, 2, a004390. [Google Scholar] [CrossRef]
  201. Hartl, F.U.; Hayer-Hartl, M. Converging concepts of protein folding in vitro and in vivo. Nat. Struct. Mol. Biol. 2009, 16, 574–581. [Google Scholar] [CrossRef] [PubMed]
  202. Yam, A.Y.; Xia, Y.; Lin, H.-T.J.; Burlingame, A.; Gerstein, M.; Frydman, J. Defining the TRiC/CCT interactome links chaperonin function to stabilization of newly made proteins with complex topologies. Nat. Struct. Mol. Biol. 2008, 15, 1255–1262. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  203. Shi, J.; Fu, M.; Zhao, C.; Zhou, F.; Yang, Q.; Qiu, L. Characterization and function analysis of Hsp60 and Hsp10 under different acute stresses in black tiger shrimp, Penaeus monodon. Cell Stress Chaperones 2015, 21, 295–312. [Google Scholar] [CrossRef] [Green Version]
  204. Sung, Y.Y.; Ashame, M.F.; Chen, S.; MacRae, T.H.; Sorgeloos, P.; Bossier, P. Feeding Artemia franciscana (Kellogg) larvae with bacterial heat shock protein, protects from Vibrio campbellii infection. J. Fish Dis. 2009, 32, 675–685. [Google Scholar] [CrossRef] [PubMed]
  205. Kumar, V.; Roy, S.; Behera, B.K.; Das, B.K. RNA Interference and Its Potential Applications in Aquatic Animal Health Management. In Biotechnological Advances in Aquaculture Health Management; Springer: Singapore, 2021; pp. 25–41. [Google Scholar]
  206. Kumar, V.; Roy, S.; Behera, B.K.; Das, B.K. Disease Diagnostic Tools for Health Management in Aquaculture. In Advances in Fisheries Biotechnology; Springer: Singapore, 2021; pp. 363–382. [Google Scholar]
  207. Roy, S.; Kumar, V.; Behera, B.K.; Das, B.K. Epigenetics: Perspectives and Potential in Aquaculture. In Advances in Fisheries Biotechnology; Springer: Singapore, 2021; pp. 133–150. [Google Scholar]
  208. Buckley, B.A.; Owen, M.-E.; Hofmann, G.E. Adjusting the thermostat: The threshold induction temperature for the heat-shock response in intertidal mussels (genus Mytilus) changes as a function of thermal history. J. Exp. Biol. 2001, 204, 3571–3579. [Google Scholar] [CrossRef] [PubMed]
  209. Osovitz, C.J.; Hofmann, G.E. Thermal history-dependent expression of the hsp70 gene in purple sea urchins: Biogeographic patterns and the effect of temperature acclimation. J. Exp. Mar. Biol. Ecol. 2005, 327, 134–143. [Google Scholar] [CrossRef]
  210. Yokoyama, S.; Koshio, S.; Takakura, N.; Oshida, K.; Ishikawa, M.; Gallardo-Cigarroa, F.J.; Catacutan, M.R.; Teshima, S.-I. Effect of dietary bovine lactoferrin on growth response, tolerance to air exposure and low salinity stress conditions in orange spotted grouper Epinephelus coioides. Aquaculture 2006, 255, 507–513. [Google Scholar] [CrossRef]
  211. Smurov, A.O.; Podlipaeva, Y.I.; Goodkov, A.V. Heat shock protein of the Hsp70 family in the euryhaline cilate Paramecium nephridiatum and its role in adaptation to salinity changes. Cell Tissue Biol. 2007, 1, 244–247. [Google Scholar] [CrossRef]
  212. Cohen, D.M.; Wasserman, J.C.; Gullans, S.R. Immediate early gene and HSP70 expression in hyperosmotic stress in MDCK cells. Am. J. Physiol. Physiol. 1991, 261, C594–C601. [Google Scholar] [CrossRef]
  213. Xiang, L.-X.; He, D.; Dong, W.-R.; Zhang, Y.-W.; Shao, J.-Z. Deep sequencing-based transcriptome profiling analysis of bacteria-challenged Lateolabrax japonicus reveals insight into the immune-relevant genes in marine fish. BMC Genom. 2010, 11, 472. [Google Scholar] [CrossRef] [Green Version]
  214. Yue-Chai, M.; Yu-jiao, Y.; Guo-liang, W. Effects of salinity challenge on the immune factors of Scylla serrata. Acta Agric. Zhejiangensis 2010, 22, 479–484. [Google Scholar]
  215. Xu, Q.; Liu, Y.; Liu, R. Expressed sequence tags from cDNA library prepared from gills of the swimming crab, Portunus trituberculatus. J. Exp. Mar. Biol. Ecol. 2010, 394, 105–115. [Google Scholar] [CrossRef]
  216. Karouna-Renier, N.K.; Zehr, J.P. Short-term exposures to chronically toxic copper concentrations induce HSP70 proteins in midge larvae (Chironomus tentans). Sci. Total Environ. 2003, 312, 267–272. [Google Scholar] [CrossRef]
  217. Arts, S.J.; Schill, R.O.; Knigge, T.; Eckwert, H. Stress Proteins (hsp70, hsp60) Induced in isopods and nematodes by field exposure to metals in a gradient near avonmouth, UK. Ecotoxicology 2004, 13, 739–755. [Google Scholar] [CrossRef]
  218. Sharp, V.; Miller, D.; Bythell, J.; Brown, B. Expression of low molecular weight HSP 70 related polypeptides from the symbiotic sea anemone Anemonia viridis forskall in response to heat shock. J. Exp. Mar. Biol. Ecol. 1994, 179, 179–193. [Google Scholar] [CrossRef]
  219. Eckwert, H.; Alberti, G.; Köhler, H.-R. The induction of stress proteins (hsp) in Oniscus asellus (Isopoda) as a molecular marker of multiple heavy metal exposure: I. Principles and toxicological assessment. Ecotoxicology 1997, 6, 249–262. [Google Scholar] [CrossRef]
  220. Staples, C.A.; Dome, P.B.; Klecka, G.M.; Oblock, S.T.; Harris, L.R. A review of the environmental fate, effects, and exposures of bisphenol A. Chemosphere 1998, 36, 2149–2173. [Google Scholar] [CrossRef]
  221. Purdom, C.E.; Hardiman, P.A.; Bye, V.V.J.; Eno, N.C.; Tyler, C.R.; Sumpter, J.P. Estrogenic Effects of Effluents from Sewage Treatment Works. Chem. Ecol. 1994, 8, 275–285. [Google Scholar] [CrossRef]
  222. Juhasz, A.L.; Naidu, R. Bioremediation of high molecular weight polycyclic aromatic hydrocarbons: A review of the microbial degradation of benzo[a]pyrene. Int. Biodeterior. Biodegrad. 2000, 45, 57–88. [Google Scholar] [CrossRef]
  223. Selvakumar, S.; Geraldine, P. Heat shock protein induction in the freshwater prawn Macrobrachium malcolmsonii: Acclimation-influenced variations in the induction temperatures for Hsp70. Comp. Biochem. Physiol. Part A Mol. Integr. Physiol. 2005, 140, 209–215. [Google Scholar] [CrossRef]
  224. Dorts, J.; Silvestre, F.; Tu, H.T.; Tyberghein, A.-E.; Phuong, N.T.; Kestemont, P. Oxidative stress, protein carbonylation and heat shock proteins in the black tiger shrimp, Penaeus monodon, following exposure to endosulfan and deltamethrin. Environ. Toxicol. Pharmacol. 2009, 28, 302–310. [Google Scholar] [CrossRef] [PubMed]
  225. Wang, Z.; Wu, Z.; Jian, J.; Lu, Y. Cloning and expression of heat shock protein 70 gene in the haemocytes of pearl oyster (Pinctada fucata, Gould 1850) responding to bacterial challenge. Fish Shellfish Immunol. 2009, 26, 639–645. [Google Scholar] [CrossRef] [PubMed]
  226. Janewanthanakul, S.; Supungul, P.; Tang, S.; Tassanakajon, A. Heat shock protein 70 from Litopenaeus vannamei (LvHSP70) is involved in the innate immune response against white spot syndrome virus (WSSV) infection. Dev. Comp. Immunol. 2019, 102, 103476. [Google Scholar] [CrossRef] [PubMed]
  227. Kumar, V.; Roy, S.; Behera, B.; Bossier, P.; Das, B. Acute Hepatopancreatic Necrosis Disease (AHPND): Virulence, Pathogenesis and Mitigation Strategies in Shrimp Aquaculture. Toxins 2021, 13, 524. [Google Scholar] [CrossRef]
  228. Cui, Y.-D.; Du, Y.-Z.; Lu, M.-X.; Qiang, C.-K. Cloning of the Heat Shock Protein 60 Gene from the Stem Borer, Chilo suppressalis, and Analysis of Expression. J. Insect Sci. 2010, 10, 100. [Google Scholar] [CrossRef] [Green Version]
  229. Gao, Q.; Zhao, J.; Song, L.; Qiu, L.; Yu, Y.; Zhang, H.; Ni, D. Molecular cloning, characterization and expression of heat shock protein 90 gene in the haemocytes of bay scallop Argopecten irradians. Fish Shellfish Immunol. 2008, 24, 379–385. [Google Scholar] [CrossRef]
  230. Almeida, M.B.; Nascimento, J.L.M.D.; Herculano, A.M.; Crespo-López, M.E. Molecular chaperones: Toward new therapeutic tools. Biomed. Pharmacother. 2011, 65, 239–243. [Google Scholar] [CrossRef]
  231. Carton, Y.; Poirié, M.; Nappi, A.J. Insect immune resistance to parasitoids. Insect Sci. 2008, 15, 67–87. [Google Scholar] [CrossRef]
  232. Niehaus, A.C.; Angilletta, M.J.; Sears, M.W.; Franklin, C.E.; Wilson, R.S. Predicting the physiological performance of ectotherms in fluctuating thermal environments. J. Exp. Biol. 2012, 215, 694–701. [Google Scholar] [CrossRef] [Green Version]
  233. Peck, L.S.; Morley, S.A.; Richard, J.; Clark, M. Acclimation and thermal tolerance in Antarctic marine ectotherms. J. Exp. Biol. 2014, 217, 16–22. [Google Scholar] [CrossRef] [Green Version]
  234. Peck, L.S. Organisms and responses to environmental change. Mar. Genom. 2011, 4, 237–243. [Google Scholar] [CrossRef] [PubMed]
  235. Somero, G.N. Proteins and Temperature. Annu. Rev. Physiol. 1995, 57, 43–68. [Google Scholar] [CrossRef] [PubMed]
  236. Morimoto, R.I.; Santoro, M.G. Stress–inducible responses and heat shock proteins: New pharmacologic targets for cytoprotection. Nat. Biotechnol. 1998, 16, 833–838. [Google Scholar] [CrossRef]
  237. Sørensen, J.G.; Kristensen, T.N.; Loeschcke, V. The evolutionary and ecological role of heat shock proteins. Ecol. Lett. 2003, 6, 1025–1037. [Google Scholar] [CrossRef]
  238. Clark, M.S.; Peck, L.S. Triggers of the HSP70 stress response: Environmental responses and laboratory manipulation in an Antarctic marine invertebrate (Nacella concinna). Cell Stress Chaperones 2009, 14, 649–660. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  239. Santoro, M. Heat shock factors and the control of the stress response. Biochem. Pharmacol. 2000, 59, 55–63. [Google Scholar] [CrossRef]
  240. Werner, I. The influence of salinity on the heat-shock protein response of Potamocorbula amurensis (Bivalvia). Mar. Environ. Res. 2004, 58, 803–807. [Google Scholar] [CrossRef]
  241. Sinnasamy, S.; Noordin, N.M.; MacRae, T.H.; bin Abdullah, M.I.; Bossier, P.; Wahid, M.E.B.A.; Noriaki, A.; Sung, Y.Y. Ingestion of food pellets containing Escherichia coli overexpressing the heat-shock protein DnaK protects Penaeus vannamei (Boone) against Vibrio harveyi (Baumann) infection. J. Fish Dis. 2016, 39, 577–584. [Google Scholar] [CrossRef]
  242. Ficke, A.D.; Myrick, C.A.; Hansen, L.J. Potential Impacts of Global Climate Change on Freshwater Fisheries; 2007; Volume 17. Available online: https://link.springer.com/article/10.1007/s11160-007-9059-5 (accessed on 11 November 2021).
  243. Lushchak, V.I.; Bagnyukova, T.V. Temperature increase results in oxidative stress in goldfish tissues. 2. Antioxidant and associated enzymes. Comp. Biochem. Physiol. Part C Toxicol. Pharmacol. 2006, 143, 36–41. [Google Scholar] [CrossRef]
  244. Lushchak, V.I.; Lushchak, L.P.; Mota, A.A.; Hermes-Lima, M. Oxidative stress and antioxidant defenses in goldfish Carassius auratus during anoxia and reoxygenation. Am. J. Physiol. Regul. Integr. Comp. Physiol. 2001, 280, R100–R107. [Google Scholar] [CrossRef] [Green Version]
  245. Costa-Pierce, B.A.; Bartley, D.M.; Hasan, M.; Yusoff, F.; Kaushik, S.J.; Rana, K.; Lemos, D.; Bueno, P.; Yakupitiyage, A. Responsible use of resources for sustainable aquaculture. Farming Waters People Food. Proc. Glob. Conf. Aquac. 2010, 113–147. Available online: https://www.fao.org/3/i2734e/i2734e.pdf (accessed on 27 September 2021).
  246. Verdegem, M.C.J. Nutrient discharge from aquaculture operations in function of system design and production environment. Rev. Aquac. 2013, 5, 158–171. [Google Scholar] [CrossRef]
  247. Romero, J.; Gloria, C.; Navarrete, P. Antibiotics in Aquaculture—Use, Abuse and Alternatives. Health Environ. Aquac. 2012, 159, 159–198. [Google Scholar] [CrossRef] [Green Version]
  248. Defoirdt, T.; Boon, N.; Sorgeloos, P.; Verstraete, W.; Bossier, P. Alternatives to antibiotics to control bacterial infections: Luminescent vibriosis in aquaculture as an example. Trends Biotechnol. 2007, 25, 472–479. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  249. Cabello, F.C. Heavy use of prophylactic antibiotics in aquaculture: A growing problem for human and animal health and for the environment. Environ. Microbiol. 2006, 8, 1137–1144. [Google Scholar] [CrossRef] [PubMed]
  250. Sayes, C.; Leyton, Y.; Riquelme, C. Probiotic Bacteria as an Healthy Alternative for Fish Aquaculture. Intech Open 2018, 115–132. [Google Scholar]
Figure 1. Role of heat shock proteins (Hsps) in proteostasis and the host immune response.
Figure 1. Role of heat shock proteins (Hsps) in proteostasis and the host immune response.
Life 12 01777 g001
Figure 2. A schematic hypothetical illustration of the possible mechanism of action for Heat shock proteins (Hsps) in crustaceans.
Figure 2. A schematic hypothetical illustration of the possible mechanism of action for Heat shock proteins (Hsps) in crustaceans.
Life 12 01777 g002
Figure 3. Effect of abiotic and biotic stresses on host health and heat shock protein response.
Figure 3. Effect of abiotic and biotic stresses on host health and heat shock protein response.
Life 12 01777 g003
Figure 4. Schematic representation of the life cycle, physiological, growth, and molecular features of the A. franciscana model system.
Figure 4. Schematic representation of the life cycle, physiological, growth, and molecular features of the A. franciscana model system.
Life 12 01777 g004
Figure 5. Advantage of a gnotobiotic (germ free) A. franciscana model system.
Figure 5. Advantage of a gnotobiotic (germ free) A. franciscana model system.
Life 12 01777 g005
Table 1. Factors that modulate heat shock protein production and their functional significance in aquatic invertebrates.
Table 1. Factors that modulate heat shock protein production and their functional significance in aquatic invertebrates.
SpeciesHsp Inducing ConditionDosage & DurationHsp InducedTissue ExaminedImmune ResponseDisease/Stress ResistanceReferences
Artemia franciscana
(Brine shrimp)
Temperature shock 30 min sub-lethal heat shock at 37 °CHsp70Whole animal (adult)--High temperature (+)[94]
30 min heat shock from 21 °C to 37 °CHsp70Whole animal (adult)--High temperature (+)[95]
30 min NLHS at 37 °C, 6 h recoveryHsp70Whole animal (nauplii)--High temperature (+)[96]
30 min heat shock from 28 °C to 37 °C, 6 h recoveryHsp70Whole animal (nauplii)--Vibrio campbellii and V. proteolyticus (+)[24]
1 h cold shock from 28 °C to 4 °C and heat shock from 28 °C to 37 °C, 6 h recoveryHsp70Whole animal (nauplii)--High temperature and Vibrio campbellii (+)[21]
Feeding Hsp overproducing bacteriaFeeding Escherichia coli overproducing prokaryotic Hsp (Dnak)DnakWhole animal (nauplii)--Vibrio campbellii (+)[97]
Feeding Escherichia coli strain (YS2 and A native) overproducing Artemia Hsp 70 and DnakHsp70, DnakWhole animal (nauplii)proPO (+)Vibrio campbellii (+)[78]
Feeding HspsFeeding truncated portion of Hsp 70Hsp70, DnakWhole animal (nauplii)proPO (+)Vibrio campbellii (+)[98]
Plant-derived/natural compoundsFeeding Tex-OE® (Hspi compound) with 20 µL/L to 160 µL/L concentrationHsp70Whole animal (nauplii)---Vibrio campbellii (+) [28]
Pretreatment Pro-Tex® (Hspi compound) with 152 ppb for 1 hHsp70 Whole animal (nauplii)--High temperature and hypersalinity (+) [99]
Feeding Tex-OE® (Hspi compound) with 2.5 mg/L to 50 mg/L concentrationHsp70Whole animal (nauplii)proPO and TGase (+)Vibrio campbellii and V. harveyi (+)[27]
Feeding phenolic pyrogallol (Hspi compound) with 79 µM to 1185 µM concentrationHsp70 Whole animal (nauplii)proPO and TGase (+)Vibrio harveyi (+)[25]
Feeding poly-β-hydroxybutyrate (PHB) with 10 mg/L to 1000 mg/L concentrationHsp70Whole animal (nauplii)proPO and TGase (+)Vibrio campbellii (+)[26]
Phloroglucinol pretreatment (Hspi compound) with 30 µM concentrationHsp70Whole animal (nauplii)--Vibrio parahaemolyticus (AHPND strain) (+)[100]
Phloroglucinol treatment (Hspi compound) with 2 µM concentrationHsp70Whole animal (nauplii)DSCAM, proPO, PXN, Hsp90, Hsp70, and LGBP (+)V. parahaemolyticus (AHPND strain), V. harveyi and high temperature (+)[9]
Sodium ascorbate pretreatment (Hspi compound) with 200 ppm concentrationHsp70Whole animal (nauplii)SOD and GST (+)Vibrio harveyi (+)[101]
Penaeus monodon
(Tiger prawn)
Temperature shock4 days of exposure to either 0.1 µg L−1 or 1 µg L−1Hsp90Muscle Pesticide (endosulfan and deltamethrin) (+)[102]
1 h heat shock from 26 to 37 °C, 30 min recoveryHsc70Hemocytes--High temperature (+)[87]
24 h heat shock from 29 °C to 35 °CHsp70Tail muscle--Gill-associated virus (+)[71]
Litopenaeus vannamei (Pacific white shrimp)Administration of beneficial bacteriaInjection of 10 µL (1 × 107 cells mL−1) of live V. alginolyticusHsp60, Hsp70Hemocytes, muscle, stomach, heart, gill, and hepatopancreasLvHSP60 and LvHSP70 (+) --[103]
Macrobrachium rosenbergii
(Freshwater prawn)
Feeding plant-derived/natural compoundsPhloroglucinol treatment (Hspi compound) with 5–10 µM concentrationHsp70Whole animal (nauplii)--Vibrio parahaemolyticus (AHPND strain) (+) [104]
Portunus trituberculatus (Japanese blue crab)Osmotic stress25 ppt to 10 ppt and 40 ppt for 24 hHsp60Gill, gill muscle, ovary, antennal gland, abdominal muscle, hypodermis, heart, and intestinePtHsp60Osmotic stress (+)[76]
Scylla paramamosain
(Mud crab)
Administration of beneficial bacteriaCrab injected with 20 µL live V. alginolyticus (107 cells/mL)Hsp70Midgut, stomach, hepatopancreases, epidermis, thoracic ganglion, gill, eyestalk, heart, brain, muscles, and hemocytes--Vibrio alginolyticus, osmotic stress, and high temperature (+)[19]
Administration of beneficial bacteriaCrabs were injected with 50 µL S. aureus (5 × 107 CFU), 50 µL V. harveyi (5 × 107 CFU), and 50 µL WSSV supernatant (5 × 105 virus particles)Hsp90Hepatopancreases, stomach, gill, intestine, muscle, connective tissue, gonad, heart, and hemocytes Staphylococcus aureus, Vibrio harveyi, and WSSV (+)[20]
Osmotic stress15 ppt to 30 ppt for 96 hHsp70Midgut, stomach, hepatopancreases, epidermis, thoracic ganglion, gill, eyestalk, heart, brain, muscles, and hemocytes--Osmotic stress (+)[19]
Temperature shock 96 h of heat shock from 25 °C to 36 °C and cold shock from 25 °C to 10 °CHsp70Midgut, stomach, hepatopancreases, epidermis, thoracic ganglion, gill, eyestalk, heart, brain, muscles, and hemocytes--High temperature (+)[19]
(+): positive effect; (--): Not studied; NLHS: non-lethal heat shock; Hspi: heat shock protein inducing compound; NP: nonylphenol; BPA: bisphenol A diglycidyl ether; EE: 17a-ethynyl estradiol; DEHP: bis(2-ethylhexyl) phthalate; proPO: Prophenoloxidase; TGase: transglutaminase; DSCAM: Down syndrome cell adhesion molecule; PXN: peroxinectin; LGBP: lipopolysaccharide and β-1,3-glucan-binding protein; SOD: Superoxidase dismutase; GST: Glutathione S-transferase; LvHSP70: L. vannamei heat shock protein 70; LvHSP60: L. vannamei heat shock protein 60; PtHsp60: P. trituberculatus heat shock protein 60; and WSSV: white spot syndrome virus.
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Kumar, V.; Roy, S.; Behera, B.K.; Das, B.K. Heat Shock Proteins (Hsps) in Cellular Homeostasis: A Promising Tool for Health Management in Crustacean Aquaculture. Life 2022, 12, 1777. https://doi.org/10.3390/life12111777

AMA Style

Kumar V, Roy S, Behera BK, Das BK. Heat Shock Proteins (Hsps) in Cellular Homeostasis: A Promising Tool for Health Management in Crustacean Aquaculture. Life. 2022; 12(11):1777. https://doi.org/10.3390/life12111777

Chicago/Turabian Style

Kumar, Vikash, Suvra Roy, Bijay Kumar Behera, and Basanta Kumar Das. 2022. "Heat Shock Proteins (Hsps) in Cellular Homeostasis: A Promising Tool for Health Management in Crustacean Aquaculture" Life 12, no. 11: 1777. https://doi.org/10.3390/life12111777

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop