Next Article in Journal / Special Issue
Nitric Oxide- and Sulfur-Mediated Reversal of Cadmium-Inhibited Photosynthetic Performance Involves Hydrogen Sulfide and Regulation of Nitrogen, Sulfur, and Antioxidant Metabolism in Mustard
Previous Article in Journal
Pollution of the Environment and Pollen: A Review
Previous Article in Special Issue
CRISPR-Cas Genome Editing for Insect Pest Stress Management in Crop Plants
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Mechanisms of Stress Tolerance in Cyanobacteria under Extreme Conditions

1
Laboratory of Algal Research, Centre of Advanced Study in Botany, Banaras Hindu University, Varanasi 221005, India
2
Department of Microbiology, C.C.S.U. Campus, Meerut 250005, India
3
Department of Biotechnology, Singhania University, Rajasthan 333515, India
4
Department of Zoology, C.M.B.College, Deorh, Ghoghardiha, Madhubani 847402, India
5
Department of Zoology, Maharaj Singh College, Maa Shakumbhari University, Saharanpur 247001, India
6
Department of postharvest Science, Agricultural Research Organization (A.R.O.)—Volcani Center, Rishon Lezion 7505101, Israel
*
Authors to whom correspondence should be addressed.
Stresses 2022, 2(4), 531-549; https://doi.org/10.3390/stresses2040036
Submission received: 17 October 2022 / Revised: 2 December 2022 / Accepted: 5 December 2022 / Published: 9 December 2022
(This article belongs to the Special Issue Physiological and Molecular Mechanisms of Plant Stress Tolerance)

Abstract

:
Cyanobacteria are oxygen-evolving photoautotrophs with worldwide distribution in every possible habitat, and they account for half of the global primary productivity. Because of their ability to thrive in a hostile environment, cyanobacteria are categorized as “extremophiles”. They have evolved a fascinating repository of distinct secondary metabolites and biomolecules to promote their development and survival in various habitats, including severe conditions. However, developing new proteins/enzymes and metabolites is mostly directed by an appropriate gene regulation system that results in stress adaptations. However, only few proteins have been characterized to date that have the potential to improve resistance against abiotic stresses. As a result, studying environmental stress responses to post-genomic analysis, such as proteome changes using latest structural proteomics and synthetic biology techniques, is critical. In this regard, scientists working on these topics will benefit greatly from the stress of proteomics research. Progress in these disciplines will aid in understanding cyanobacteria’s physiology, biochemical, and metabolic systems. This review summarizes the most recent key findings of cyanobacterial proteome study under various abiotic stresses and the application of secondary metabolites formed during different abiotic conditions.

1. Introduction

Cyanobacteria are a very diverse group photosynthetic prokaryotes that originated approximately 3.5 billion years ago and performed a critical role in shifting the earth’s environmental conditions from anaerobic to aerobic. The cyanobacteria are typically the first oxygen-evolving photoautotrophic colonizer of primary deglaciated environments, kicking off the first succession process. They also help to stabilize the soil by forming crusts on the top. Cyanobacteria are ubiquitous and can be found in aquatic and terrestrial ecosystems, including hot springs, deserts, and polar regions. Although these places have harsh climatic circumstances, cyanobacteria showed their ability to cope with the stresses of low temperatures and constant irradiance of high and low intensity of light, ultraviolet radiation (U.V.R.), freezing, and desiccation [1,2]
However, the stress conditions depend upon various factors, such as duration (U.V. exposure), periodicities (such as diel or annual temperature cycles), high irradiance, desiccation, and elevated and low temperatures. The cyanobacteria respond quickly after exposure to these stress factors and move under a state transition, which may be slow, with changes lasting months or years, which can lead to the emergence of a new strain [3].
The cyanobacterial strains can be used as a model to study the stress response because of a direct link with eukaryotic photosynthetic organisms or the plant [4]. The latest molecular techniques are employed at the molecular (D.N.A., R.N.A., and protein) levels to explore the physiological, biochemical, and metabolic activity of stressed organisms. However, the relationship between transcript and protein expressions still needs to be deeply studied [5] because it correlates the actual functions of genes and their transcribed products and can reflect the stress-specific cellular protein profile under defined stress circumstances.
In addition, proteomics offers advances in understanding cellular functions [6]. It can investigate the function and structure of proteins by acting as a link between the transcriptome and metabolomic profiles, allowing researchers to probe the cell’s physiological and metabolic states. Furthermore, in a stressful situation, the creation of “adaptation proteins” controls long-term cellular adaptation. These adaptation proteins are being studied to determine how the cells adapt to stressful situations. Identifying unregulated proteins with the latest new-generation techniques can be beneficial for studying stress proteomics.
The production of secondary metabolites, in response to the different biotic and abiotic stresses, also protects the cyanobacteria [7,8]. Depending on the stress situation, different suites of secondary metabolites are produced. Alkaloids, UV-absorbing polyketides, and terpenoids are frequently reported metabolites. As a result, they have a wide range of protective properties for the cells, including defense against grazers and predators, antioxidant, chemosensory, and photoprotection functions. These qualities can be used in cosmeceuticals, nutraceuticals, and medicines in industrial biotechnology.
This review article briefly discussed the different abiotic stresses and their impact on cyanobacterial biology. In addition, it summarizes the proteomics associated with cyanobacteria during exposure to different stresses; lastly, we have also summarized how the cyanobacterial inoculation enhances the growth and yields of the plant.

2. Occurrence of Cyanobacteria in the Environment

Cyanobacteria are ubiquitous and present in various shapes and sizes, including unicellular, colonial, filamentous, branching, and heterocystous, in the different possible environments. Cyanobacteria can switch between autotrophic, mixotrophic, and heterotrophic feeding modes. In addition, they play an essential part in global biogeochemical cycles by transforming organic molecules into proper forms [9].
Moreover, cyanobacteria are a substantial component of marine and freshwater ecosystems with global distribution. Some cyanobacterial species, when separated from coastal settings, tolerate high concentrations of salt water (halotolerant) [10]. However, some cyanobacterial strains can survive in the terrestrial habitats and play a vital role in nutrient recycling and sustainable environmental development. The presence of UV-absorbing sheath pigments and exopolysaccharide (EPS) make the cyanobacterial strains survive in the exposed terrestrial environment. However, some cyanobacterial strains are common colonizers of euryhaline environments and survive under 3–4 molar mass salt concentrations [11]. Cyanobacteria are remarkable colonizers of infertile substrates, such as desert sand, volcanic ash, and boulders [12]. They are incredible excavators who carve out hollows in the limestone and specific sandstone. Mountain streams [13], hot springs [14], and Antarctic and Arctic lakes [15] are all the homes of cyanobacterial species [16]. Additionally, they can be found in a wide variety of water types, from katharobic waters to polysaprobic zones [17].

Cyanobacterial Stress Tolerance

Cyanobacteria are known for their strong adaptability under different environmental conditions, such as exposure to Gamma or UV radiation, heat, salinity, desiccation, and heavy metals. However, some of the cyanobacterial strains are resistant to these stress conditions, as listed in Table 1. These environmental condition affect the morphology (cell wall alteration), biochemicals (production of screening pigments), or physiology (state transitions), or there may be some genetic changes that occur as a result of acclimatization under the environmental extremes [18].
In addition, the imposition of stress can lead to changes in the metabolic behaviors, such as the enhanced synthesis of polyunsaturated fatty acids [48] and enhanced concentration of photosystem unit in the thylakoid membrane, which is commonly observed [49]. However, the response time of short-term reactions are usually less than a minute, and rare structural modifications have been observed [50]. However, long-term reactions can span from several hours to days and frequently entail de novo protein formation. The changes in photosystem stoichiometry and alteration in the pigment quantity are examples of long-term responses [51].

3. General Mechanisms of Cyanobacteria to Survive under Extreme Habitat

The abiotic stresses, including the presence of salt, heavy metals, light, drought, and temperature extremes, primarily affect the physiology and metabolic behavior of the cyanobacteria [52]. However, in response to stress, a variety of stimuli and other defense systems have been activated [53,54]. Changes in the osmotic pressure have been observed during various stresses, such as the enhanced concentration of salt [55], freezing stress [56], and the stress of desiccation [57].
The generation of reactive oxygen species (ROS) has been observed after high P.A.R. [58] and U.V.R. [59], and these ROS have been neutralized by different enzymatics, such as superoxide dismutases (SOD), catalases (CAT), and guaiacol peroxidase (GPX), and non-enzymatic substances, such as carotenoids and glutathione reductase (G.R.) [60]. However, the excess generation of free radicals can lead to oxidative stress, which can affect the structural integrity and nucleic acids of the cells [61]. ROS also act as secondary signal molecules, regulating photosynthetic genes and inducing antioxidant enzymes through genetic regulation. However, the cyanobacteria can recover from the oxidative damages caused due to mild UV-B exposure, thanks to their efficient defense and repair system [62]. However, the prolonged exposure to UVR stress can cause the denaturation of DNA. The principal harm is the formation of photoproducts between adjacent bases; similar types of damage were found during desiccation stress [57,63].
The fluidity and integrity of the cell membranes are critical for normal life [64]. However, stresses such as temperature [65], salinity [66], desiccation [67], and photoinhibition generally cause alternation in lipid content of the cell [68]. Different strategies employed by cyanobacteria to cope with different abiotic stresses are described in Figure 1.
Heat-shock proteins (HSPs) were first reported in the context of high-temperature responses, but now have also been reported during different stress conditions [69,70]. However, in response to salinity stress cyanobacteria synthesizes nonreducing carbohydrates like trehalose [64]. Studies reported trehalose’s involvement in maintaining the integrity of membranes, proteins, and nucleic acids, also involved in their functions in osmotic stress tolerance [71].

3.1. Photosynthetic Active Radiation (P.A.R.)

All the photoautotrophic organisms rely on light to synthesize their food and as an energy resource. However, the availability of P.A.R. and its intensity change over time [23]. The intensity can range from as low as 0.1 mol m2 s1 to as high as 1500 mol m2 s1 in full sunlight [72,73]. As a result, polar cyanobacteria adapt to an oscillation regime and a wide range of P.A.R. [23]. However, the different cyanobacterial populations compete with each other for the radiation, light-scattering, and absorbing elements in the aquatic environment. This competition leads to vertical and horizontal species distribution in aquatic and terrestrial ecosystems [74,75].
Tiny cyanobacterial filaments or unicellular cyanobacterium may reduce the self-shading, which can use poor light quality for survival. However, the cyanobacterial group, having mat-forming characteristics, showed optical properties similar to picophytoplankton [23]. Significant levels of Chl-a and many thylakoids are signs of acclimatization to low intensities [76]. However, high irradiances are frequent on the exposed surfaces, such as soil crusts, rock, and glaciers, which leads to decreases in the surface density of thylakoid membranes, Chl-a content, and the buildup of stored chemicals [76] (highly reduced ratio of PS I to PS II, it could be below one) [77].

3.2. Light Quality

The absorption of particular wavelengths of light causes cyanobacteria to modify their pigment makeup results in a complementary chromatic adaptation to maximize the use of available light [78]. However, the absorption spectrum of light varies as a result of the adaptation to light quality [79]. The alteration in the pigment is directly associated with the ability to adapt during different intensities and duration of light. However, the upper surface protects the deeper layers, where the rate of photosynthesis is higher in a layered community [80]. In a previous study, Calothrix sp. and Synechococcus sp. have been evaluated for chromatic adaptations [81]. During the study, a higher concentration of phycoerythrin (P.E.) was generated when the Calothrix sp. PCC 7601 cultivated under green light, while an elevated level of phycocyanin (P.C.) was observed under the red-light condition [82].

3.3. Ultraviolet Radiation (U.V.R.)

The exposure of U.V. radiation significantly affects the morphology and physiology of cyanobacteria [83,84]. Cyanobacteria travel to areas with low UV-B flux to avoid the harm of UV-B and try to make a balance for efficient photosynthesis and pigment bleaching by optimizing their position in inland mats or the water column [85]. However, to cope with the stress of UV-B radiation and photooxidative damage caused by ROS, cyanobacteria synthesize metabolites, such as mycosporine-like amino acids (M.A.A.s) and scytonemin. These metabolites scavenge the reactive oxygen species and repair the D.N.A., which was damaged due to U.V. radiation [86]. The overview of the survivability of cyanobacteria under exposure to U.V. radiation has been summarized in Figure 2.
The U.V.R. sensitivity should be lower in the species that originated from the exposed areas. Although, resistant species have more giant cells or form coenobia, vast colonies, or mats that protect the interior cells/filaments [87,88]. In the previous study, it has been stated that the mats of resistant strains have been characterized by actively moving filaments, as well as the surface filaments, which contain modest concentration of pigments [89]. Surface cells could produce screening substances, such as sheath pigments, that have property to protect the consortium of inner cells [90]. U.V.R. at high levels slows growth and damages the photosynthetic system, including enzyme nitrogenase and membranes, but D.N.A. is the main target [91,92,93]. The study suggested exposure to UV-B suppresses the ATP production [94,95].
However, various changes have been observed at the protein and metabolite levels in the cyanobacteria after exposure of UV-B. Changes in the proteome during U.V. acclimation and UV-B shock, as indicated by subtractive high-resolution 2D gel electrophoresis, was investigated in the work on Nostoc commune [96]. However, the U.V. impact on protein synthesis has been classified into three categories, based on exposure time: (i) continuous decrease or increase of proteins after prolonged exposure, (ii) durable repression or induction of proteins after short-term exposure, and (iii) transient repression or short-term exposure [97].

3.4. Temperature

Rising global temperature is a severe threat to living organisms, and it has been estimated that temperature may rise by 0.2 °C per decade [98]. Temperature variation also influences a wide range of biological processes, including the physiology and metabolism of cyanobacteria [99]. However, the most crucial variation after temperature exposure is the variation in the membrane composition [100]. Based on temperature tolerance, cyanobacteria are divided into the psychrophilic, psychrotrophic, mesophilic, and thermophilic groups. Each of the cyanobacterial groups has a specific growth temperature. Psychotropics can withstand freezing temperatures (below 15 °C), mesophilic can withstand temperatures up to 50 °C, and thermophilic can withstand temperatures beyond 80 °C [101]. The duration and intensity of the temperature changes the components and metabolism of the cells [102]. Temperature-induced changes in membrane fluidity are one of the most immediate repercussions. The shift in membrane composition, which comprises changes in membrane fluidity, heat shock proteins (HSPs), fatty acid saturation, and protein integrity, are some common responses of temperature stress, as summarized in Figure 3.
Temperature variations have an impact on membrane lipid saturation and protein integrity. Therefore, 2-DE and MALDI-TOF analysis has been carried out to assess the influence of high temperature on the proteome of different cyanobacterial strains. For example, Anabaena doliolum, Synechocystis PCC 6803, and Spirulina platensis showed high denaturation and protein aggregation during analysis [103,104,105].

3.4.1. Low Temperature

It is well-known that cyanobacteria can survive at very low temperatures, even up to −20 °C [106]. However, exposure to low temperatures causes desaturation of fatty acids in the membrane and activates certain enzymes that boost transcription and translation efficiency [9,12,15,107]. In the previous study, it has been reported that exposure of cold stress can trigger more than 100 genes. In addition, the Desa, DesB, DesC, and DesD are cyanobacterial acyl-lipid desaturases that cause the unsaturation of fatty acids by forming a double bond in the 12, 15, 9, and 6 positions under the cold stress [108,109]. Respiration and photosynthesis, membrane and cell wall maintenance, transcription and translation, signal detection and transduction, and other cellular processes (nucleotide metabolism, cofactor biosynthesis) and unknown functions are all grouped into six groups. Other cold-temperature adaptation strategies include the formation of antifreeze and cold shock proteins, modifying critical protein kinetics, and building polyunsaturated fatty acids (PUFA) in membranes [107].

3.4.2. High Temperature

High temperatures cause excessive fluidity in the membrane and also change the nature of bonding (electrostatic and hydrogen) between different polar groups of proteins, resulting in structural changes and ion leakage [110]. However, the aggregation and denaturation of proteins are common phenomenon occurring during high-temperature stress, disturbing protein functions and trafficking. The HSPs then act as proteases and chaperonins and assist in protein refolding and causing high-temperature tolerance [105,111,112].

3.5. pH Stress

pH has an impact on cyanobacteria’s physiological and metabolic activities, as well as their growth. Changes in pH severally affect cellular physiology and metabolic functions, such as nutritional bioavailability and substance transfer across the cytoplasmic membranes. However, the ability of cyanobacteria for everyday functioning is precise under a fixed pH range. Cyanobacteria maintain a pH of 7.1–7.5 on the inside and nearly 5–10 on the outside of the cell. Extracellular pH reduction lowers the intracellular pH and affects several processes, such as cell wall biosynthesis and solute transfer, and it may limit the growth [113]. The enzymes carbonic anhydrase and oxalate decarboxylase are known to play a crucial role in pH regulation and activation. The upregulation of ATP-substrate binding proteins implies that the phosphate absorption mechanism is highly vulnerable to higher medium pH.

3.6. Salinity and Osmotic Stress

Salinity affects 20% of agriculture and 7% of the world’s land area and significantly limits crop productivity [114]. However, osmotic stress is a type of physical stress that occurs due to differences in the ion concentration of inside or outside of the cell. In general, both the salinity and osmotic stresses have comparable effects on cyanobacterial cells [115]. Salinity changes water potential and induces water loss and raises ionic concentration of the cell [116,117]. However, higher concentrations of inorganic ions can harm the cellular growth and metabolism of cyanobacteria [118], but in comparison to higher plants, cyanobacteria are much more resistant to salt stress [117]. The plasma membrane of the cell, which controls intracellular trafficking, is one of the critical aspects for the cyanobacteria to acclimatize under salty concentrations [119,120,121]
In general, cyanobacteria follow limited sodium absorption and active sodium efflux via Na+/H+ antiport as a control mechanism during accumulation of increased Na+ ions in the cytoplasm of cyanobacterial cells [122]. In addition, the cyanobacterial cell follows some other mechanisms during encounter Na+ toxicity: (1) collecting organic molecules to keep the osmoticum constant [123], (2) antioxidant defense system activation to detoxify R.O.S. [119], and (3) induction of salt-inducible proteins [124]. Nowadays, blue native SDS-PAGE, followed by MALDI-TOF/MS, analysis has been carried out to deem alteration in membrane dynamics at high salt concentrations in cyanobacteria The proteins involved in cell wall production have been discovered in Synechocystis [125]. These changes likely create a strong diffusion barrier, resulting in decreased inorganic ion flow into the periplasm [126].
Ionic stress has a significant impact on the cyanobacterial nitrogen fixation. The exposure of NaCl results in the inhibition of nitrogenase activity [127,128]. However, in the Microcoleus vaginitis, sucrose phosphate synthase (S.P.S.) activity increases with the elevated concentration of NaCl [129]. Differential stimulation of antioxidative systems in response to higher saline concentration has been observed in Synechocystis and Anabaena. In a study, the author reported a decrease in the S.O.D. enzyme in Anabaena doliolum after 24 hrs of exposure to 150 mM salt [119]. However, a higher concentration of saline conditions is also related to the production of glucosyl glycerol via glycolate metabolism and photorespiration; this is also a critical metabolic alteration observed during salinity stress conditions [130,131].

3.7. Desiccation

Desiccation/draught is one of the common stress factors of terrestrial and polar hydro-terrestrial ecosystems [132]. Cyanobacteria have been considered poikilohydric creatures, which means they can endure dehydration [133]. In the previous study, various authors have reported the desiccation potential of cyanobacteria. For example, the survival of dried Nostoc commune samples have been reported after 55 years [134] and 87 years [135], and also for more than a century [136]. Cyanobacteria follow numerous methods to decrease mechanical and osmotic stresses, similar to salt and freezing stress [137]. The desiccation adversely affected several physiological processes in consecutive order: firstly, stop the nitrogen fixation then slow rate of photosynthesis, and ultimately slow down the respiration although these physiological processes can be restored after rehydration. The ability of cyanobacteria to endure low water potential exemplifies their excellent desiccation tolerance (Figure 4). In a study, Chroococcus cryptoendolithic and Chroococcidiopsis have been reported to fix carbon dioxide at lower water potentials [138].

3.8. Freeze and Melting Cycles

In polar areas, cyanobacterial populations are frequently vulnerable to freeze/melt cycles [139,140]. Various cyanobacterial groups under low temperatures, especially in the Antarctica region, showed modifications in the structural characteristics [141]. The ice crystals can penetrate the intracellular structures, and the changing osmotic gradient can harm the cell [56,142]. However, the presence of anti-freezing proteins and other cryoprotectants, such as Dimethylsulfoxide (DMSO), inhibit the ice crystals’ formation [143].

3.9. Metals and Metalloid Stress

In general, cyanobacteria require significantly higher concentration of metals than non-photosynthetic organisms, since they perform oxygenic photosynthesis. Metal ions, such as Mg and Mn, play various roles in the photosynthetic complex. However, copper (Cu) is present in the cytochrome oxidase and plastocyanin of cyanobacteria. [144]. Cyanobacteria safeguard metallic homeostasis by various methods, including membrane alterations, transfer pump activation or inactivation, biotransformation, and sequestration [145]. However, the higher concentrations of Zn and Cu can lead to the generation of reactive oxygen species (R.O.S.), which affect normal physiology [146]. In addition, the excess Cu can disturb the photosynthesis and redox equilibrium of the cells, disrupting the cells’ ultrastructure and, finally, leading to death [147,148,149]. Metallothioneins (M.T.) are small cysteine-rich proteins associated with metals (such as Cd, As, Cu, Zn, and Hg) that were discovered for the first time in cyanobacteria that helps in surviving even under high Cd or Zn concentrations [150]. Similarly, rather than using proteomic research, most work on metal stress has relied on genomic methods, such as D.N.A. microarrays and targeted mutagenesis [151]. Only a few proteome-based studies have been documented [152,153].

4. Proteomics and Cyanobacteria

Proteomics reflects the cell’s functionality, bringing the ‘virtual life of genes to the actual life of proteins’ [5,154]. This represents an organism’s complete protein composition at a specific moment, under specific conditions. With recent discoveries in biotechnology, high proteomic techniques provide crucial improvements in detecting the integration, regulation, and function of the proteome [155]. In cyanobacteria, stress tolerance is regulated by significant gene expression, which changes the centre dogma’s downstream process. As a result, analyzing alterations in the metabolome and proteome is critical, since they are direct mediators of stress responses [4,156,157,158]. The elucidation of subcellular compartments, such as plasma periplasm [124], PSII complex [159], and thylakoid proteins, showed recent success in proteomic investigations [160]. Proteomics allows for a more precise knowledge of how environmental perturbations affect a system. Several proteomics investigations with different cyanobacterial species have been completed under various abiotic stresses (Table 2).

5. Application of Cyanobacterial Secondary Metabolites in Imparting Abiotic Stress Tolerance

The interaction of cyanobacteria with plant species have a long history, and they interact with each other in both positive and negative ways. The symbiotic relationships of rice with nitrogen fixing cyanobacteria in rice fields have been broadly studied. However, cyanobacteria synthesize a range of bioactive compounds that show plant growth promoting potential, and nowadays, cyanobacterial inoculants have been frequently utilized as a biofertilizers in the agricultural fields [171]. In previous studies, several authors reported constructive cyanobacterial functions, such as fixing atmospheric carbon and nitrogen, improving nutrient recycling, and producing various bioactive compounds [172]. They have much potential to store nitrogen and phosphorus in their protoplasm and can help in restoring soil fertility. In the previous study, several authors reported the enhancement in growth and yields after cyanobacteria filtrate. For example, Hashtroudi et al. [172] reported the growth of some horticultural crops by using water extracts of different cyanobacterial strains. The extract contained significant amount of auxin, which showed a stimulatory effect on root growth, dry weight, and plant height. Similar types of observation were reported by Haroun et al. [173] after applying Cylindrospermum muscicola and Anabaena oryzae filterate, which, after application, enhanced the chlorophyll, carbon, and nitrogen contents.
In cyanobacterial extracts, varying amounts of caffeic, polyphenolics gallic, vanillic, ferulic acids, kaempferol, flavonoids, rutin, and quercetin have been reported. The existence of these polyphenolics showed their ability to scavenge free radicals, chelate metals, and protect themselves from oxidative damage [174]. In addition, the predominance of polyphenols in cyanobacterial strains can lead to their better ecological adaptation under various stress conditions [174]. By generating and releasing a wide variety of bioactive chemicals, cyanobacteria can enhance plant growth and development [175]. In addition, they can also be used to impart abiotic stress tolerance in plants, as listed in Table 3.
Many abiotic variables (drought, salinity, and severe temperatures) manifest as osmotic stresses in plants, resulting in the formation of R.O.S. that damage carbohydrates, proteins, lipids, and D.N.A. and produce abnormal cell signaling [176]. However, soil inoculation with cyanobacterial-based biostimulants increased the antioxidant potential [177]. Similarly, Singh et al. [178] reported that soil inoculation with Plectonema boryanum and Oscillatoria acuta increased the enzymatic activity of phenylalanine ammonia-lyase and peroxidase in rice leaves, resulting in systemic tolerance against the stress.
Table 3. Role of cyanobacteria and its mechanism in imparting abiotic stress tolerance in plants.
Table 3. Role of cyanobacteria and its mechanism in imparting abiotic stress tolerance in plants.
StressCyanobacteriaPlantMechanismReferences
ColdAnabaena sp.Agrostissto lonifera L.Growth modulation in plants via transformation in flavodoxin gene[179]
Aphanothece halophyticPopulus albaWith the transformation in HSP70 gene[180]
Synechocystis sp. Synechococcus vulcanusNicotiana tabacumPlant transformation with a Δ9 and Δ12 acyl–lipid desaturase genes[181]
SalinityNostoc flagelliformeArabidopsis thalianaTransformation in plants using P-loop NTPase domain gene[182]
Anabaena vaginicola ISB42, Nostoc calcicola ISB43, Trichormus ellipsosporus ISB44, and Cylindrospermum michailovskoense ISB45Mentha piperitaGrowth promotion plant[183]
Nostoc carneum TUBT04 and N. carneum TUBT05Oryza sativaGrowth and yield enhancement in plants[184]
Lyngbya mucicola; Oscillatoria Princeps L.; Lyngbya phormidium
Gloeocapsa sp.; Cryophilic sp.
Oryza sativaEnhancement in soil organic matter and enhanced concentration of plant nutrients[185]
Anabaena oryzae, Anabaena doliolum, Phormidium fragile, Calothrix geitonos, Hapalosiphon intricatus, Aulosira fertilissima, Tolypothrix tenuis, Oscillatoria
Acuta, and Plectonema boryanum
Oryza sativaGrowth promotion and modulation of phytohormone concentration[186]
Scytonema hofmanniOryza sativaProduction of gibberellin-like hormone[187]
Nostoc kihlmani and Anabaena cylindricaTriticum aestivumEnhanced growth and improved soil structure[188]
Anabaena sphaerica Production of extracellular polysaccharides and proteins[189]
Nostoc flagelliformeArabidopsis thalianaTransformation in plants with P-loop NTPase domain gene[182]
Nostoc muscorum; Anabaena fertilissima, A. anomalaOryza sativaIncrease in the rhizosphere biology or phytohormone[190]
Nostoc calcicola, Nostoc spongiaeformae, Nostoc linckia, and Nostoc muscorumWheat, maize, and riceIncreased seed vigor index and germination, vigor index[191]
Anabaena variabilisMedicago truncatulaTransformation in plants with flavodoxin gene[192]
Anabaena sphaericaOryza sativaProduction of extracellular polysaccharides and proteins[193]
DroughtNostoc flagelliformeArabidopsis thalianaEnhance seed germination under salinity stress[182]
Nostoc sp. and Microcoleus sp.Senna notabilis and Acacia hillianaSeeds priming and enhanced germination[194]
Anabaena sp. Agrostis stolonifera L.Transformation in plants with a flavodoxin gene[179]
Leptolyngbya sp., Microcoleus sp., Nostoc sp., and Scytonema sp.Eucalyptus gamophylla and Grevillea wickhamiiEnhanced radicle initiation and improved shoot and root growth[195]
Anabaena sp.Nicotiana tabacumPlant transformation with a flavodoxin gene[196,197]
Osmotic stressAnabaena sp.Nicotiana tabacumTransformation in plants with flavodoxin and a ferredoxin–NADP+ reductase gene[198]
HeatAnabaena sp.Agrostis stolonifera L. Plant transformation with a flavodoxin gene[179]
Nostoc sp., Anabaena doliolum, Calothrix sp., Westiellopsis sp., and Phormidium papyraceumOryza sativaHeavy metal stress tolerance and improved plant growth[199]
Heavy
metal
Nostoc muscorumTrigonella foenumgracumEnhancement in the sugar, protein, and lipid content[200]
Synechococcus sp.A. thalianaTransformation in plants with a metallothionein gene[201]
Nostoc sp., Anabaena doliolum, Calothrix sp., Westiellopsis sp., and Phormidium papyraceumOryza sativaTolerance of heavy metal stress and enhancement in plant growth[202]
Spirulina platensisZea mays L.Activation of plant mechanisms[203]

6. Future Perspective

In the last few years, various authors have studied the impact of abiotic stresses on cyanobacterial biology and reported their observations. However, the proteins are the most essential ingredients of the cell, largely affected under stress conditions. Proteins are essential parts of the cell that are severally affected and the main effectors of metabolic disorders caused by abiotic stress. Stresses such as cold, salt, thirst, metals, and metalloids cause similar harm. Almost all stimuli have been linked to increased production of proteins linked to oxidative damage. Photorespiration is primarily induced by salt and, in some situations, UV-B induces D.N.A. damage directly. Temperature and pesticides primarily target membrane dynamics, while metals impact the system’s redox status.
However, certain stress conditions are used during the production of specific metabolites because cyanobacteria have been reported to acclimate to changing environmental conditions by modulating the concentration of metabolites. Detailed biochemical research could lead to the discovery of new biotechnologically essential substances. These molecules could be employed in food supplements, cosmetics, cryopreservation, medicine, and other applications. In addition, cyanobacteria could be used as part of regenerative waste disposal system in future aquatic and terrestrial environments.
Many of the proteins examined during cyanobacterial stress biology are linked with the higher plants, confirming their close evolutionary link. However, the proteomics in cyanobacteria is still in its early stages and needs an extensive study to understand stress responses. More research into the intricate web of metabolites and proteins under different combined stress situations is needed to address the types and ranges of proteins and their involvement in cyanobacterial survival and nourishment.

Author Contributions

R.K.G. and P.Y. designed study and investigation; P.Y., R.P.S., D.J., D.K. and N.B. wrote the manuscript; R.K.G. and A.K. supervised the study; S.R. and N.B. provided valuable feedback to this study. All authors have read and agreed to the published version of the manuscript.

Funding

This research received no external funding.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Not applicable.

Acknowledgments

The authors are thankful to the Head, Department of Botany, B.H.U., Varanasi, and DST-FIST for providing the infrastructure and necessary research facilities. The authors (P.Y. and R.P.S.) are also grateful to U.G.C., Government of India, New Delhi for providing junior/senior research fellowship (JRF/SRF).

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Tashyreva, D.; Elster, J. Annual cycles of two cyanobacterial mat communities in hydro-terrestrial habitats of the high Arctic. Microb. Ecol. 2016, 71, 887–900. [Google Scholar] [CrossRef] [PubMed]
  2. Zakhia, F.; Jungblut, A.D.; Taton, A.; Vincent, W.F.; Wilmotte, A. Cyanobacteria in cold ecosystems. In Psychrophiles: From Biodiversity to Biotechnology; Springer: Berlin/Heidelberg, Germany, 2008; pp. 121–135. [Google Scholar]
  3. Larsson, J.; Nylander, J.A.; Bergman, B. Genome fluctuations in cyanobacteria reflect evolutionary, developmental and adaptive traits. B.M.C. Evol. Biol. 2011, 11, 1–21. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  4. Burnap, R.L. Systems and photosystems: Cellular limits of autotrophic productivity in cyanobacteria. Front. Bioeng. Biotechnol. 2015, 3, 1. [Google Scholar] [CrossRef] [PubMed]
  5. Gao, Y.; Xiong, W.; Li, X.B.; Gao, C.F.; Zhang, Y.L.; Li, H.; Wu, Q.Y. Identification of the proteomic changes in Synechocystis sp. PCC 6803 following prolonged UV-B irradiation. J. Exp. Bot. 2009, 60, 1141–1154. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  6. Li, T.; Yang, H.M.; Cui, S.X.; Suzuki, I.; Zhang, L.F.; Li, L.; Bo, T.T.; Wang, J.; Murata, N.; Huang, F. Proteomic study of the impact of Hik33 mutation in Synechocystis sp. PCC 6803 under normal and salt stress conditions. J. Proteome Res. 2012, 11, 502–514. [Google Scholar] [CrossRef]
  7. Burja, A.M.; Dhamwichukorn, S.; Wright, P.C. Cyanobacterial postgenomic research and systems biology. Trends Biotechnol. 2003, 21, 504–511. [Google Scholar] [CrossRef]
  8. Gupta, V.; Ratha, S.K.; Sood, A.; Chaudhary, V.; Prasanna, R. New insights into the biodiversity and applications of cyanobacteria (blue-green algae)—Prospects and challenges. Algal Res. 2013, 2, 79–97. [Google Scholar] [CrossRef]
  9. Yadav, P.; Gupta, R.K.; Singh, R.P.; Yadav, P.K.; Patel, A.K.; Pandey, K.D. Role of cyanobacteria in green remediation. In Sustainable Environmental Clean-Up; Elsevier: Amsterdam, The Netherlands, 2021; pp. 187–210. [Google Scholar]
  10. Gallon, J.R.; Jones, D.A.; Page, T.S. Trichodesmium, the paradoxical diazotroph. Arch. Hydrobiol. Suppl. Algol. Stud. 1996, 83, 215–243. [Google Scholar] [CrossRef]
  11. Reed, R.H.; Richardson, D.L.; Warr, S.R.; Stewart, W.D. Carbohydrate accumulation and osmotic stress in cyanobacteria. Microbiology 1984, 130, 1–4. [Google Scholar] [CrossRef] [Green Version]
  12. Dor, I.; Danin, A. Cyanobacterial desert crusts in the Dead Sea Valley, Israel. Arch. Hydrobiol. Algol. Stud. 1996, 83, 197–206. [Google Scholar] [CrossRef]
  13. Kann, E. Zur AutØkologie benthischer Cyanophyten in reinen europäischen Seen und Fliessgewässern. Algol. Stud. Arch. Hydrobiol. 1988, 50–53, 473–495. [Google Scholar]
  14. Castenholz, R.W. Species usage, concept, and evolution in the cyanobacteria (blue-green algae). J. Phycol. 1992, 28, 737–745. [Google Scholar] [CrossRef]
  15. Schulberg, O.M. Oscillatoialean cyanoprokaryotes and their application for algal culture technology. Arch. Hydrobiol. Algol. Stud. 1994, 75, 265–1278. [Google Scholar]
  16. Laamanen, M. Cyanoprokaryotes in the Baltic Sea ice and winter plankton. Arch. Hydrobiol. Suppl. Algol. Stud. 1996, 83, 423–433. [Google Scholar] [CrossRef]
  17. Van Landingham, S.L. Guide to the Identification, Environmental Requirements and Pollution Tolerance of Freshwater Blue-Green Algae (Cyanophyta); Environmental Monitoring and Support Laboratory, Office of Research and Development, U.S. Environmental Protection Agency: Las Vegas, NV, USA, 1982. [Google Scholar]
  18. Elster, J.; Lukesová, A.; Svoboda, J.; Kopecky, J.; Kanda, H. Diversity and abundance of soil algae in the polar desert, Sverdrup Pass, central Ellesmere Island. Polar Rec. 1999, 35, 231–254. [Google Scholar] [CrossRef]
  19. Marcozzi, C.; Camino, A.C.; Salerno, G.L. Role of NtcA, a cyanobacterial global nitrogen regulator, in the regulation of sucrose metabolism gene expression in Anabaena sp. PCC 7120. Arch. Microbiol. 2009, 191, 255–263. [Google Scholar] [CrossRef]
  20. Schwarz, R.; Forchhammer, K. Acclimation of unicellular cyanobacteria to macronutrient deficiency: Emergence of a complex network of cellular responses. Microbiology 2005, 151, 2503–2514. [Google Scholar] [CrossRef] [Green Version]
  21. Hawco, N.J.; Fu, F.; Yang, N.; Hutchins, D.A.; John, S.G. Independent iron and light limitation in a low-light-adapted Prochlorococcus from the deep chlorophyll maximum. ISME J. 2021, 15, 359–362. [Google Scholar] [CrossRef]
  22. Śliwińska-Wilczewska, S.; Konarzewska, Z.; Wiśniewska, K.; Konik, M. Photosynthetic pigments changes of three phenotypes of picocyanobacteria Synechococcus sp. under different light and temperature conditions. Cells 2020, 9, 2030. [Google Scholar] [CrossRef]
  23. Vincent, W.F. Cyanobacterial dominance in the polar regions. In The Ecology of Cyanobacteria; Whitton, B.A., Potts, M., Eds.; Kluwer Academic Publishers: Dordrecht, The Netherlands, 2000; pp. 321–340. [Google Scholar]
  24. El-Sheekh, M.M.; Alwaleed, E.A.; Ibrahim, A.; Saber, H. Detrimental effect of UV-B radiation on growth, photosynthetic pigments, metabolites and ultrastructure of some cyanobacteria and freshwater Chlorophyta. Int. J. Radiat. Biol. 2021, 97, 265–275. [Google Scholar] [CrossRef]
  25. Rastogi, R.P.; Sinha, R.P.; Moh, S.H.; Lee, T.K.; Kottuparambil, S.; Kim, Y.J.; Rhee, J.S.; Choi, E.M.; Brown, M.T.; Häder, D.P.; et al. Ultraviolet radiation and cyanobacteria. J. Photochem. Photobiol. B Biol. 2014, 141, 154–169. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  26. Orellana, G.; Gómez-Silva, B.; Urrutia, M.; Galetović, A. UV-A Irradiation Increases Scytonemin Biosynthesis in Cyanobacteria Inhabiting Halites at Salar Grande, Atacama Desert. Microorganisms 2020, 8, 1690. [Google Scholar] [CrossRef] [PubMed]
  27. Kraus, M.P. Resistance of blue-green algae to 60Co gamma radiation. Radiat. Bot. 1969, 9, 481–489. [Google Scholar] [CrossRef]
  28. Li, S.; Xu, M.; Su, Z. Computational analysis of LexA regulons in Cyanobacteria. BMC Genom. 2010, 11, 1–17. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  29. Singh, H.; Apte, S.K. Effect of 60Co-Gamma Ionizing Radiation and Desiccation Stress on Protein Profile of Anabaena 7120. Protein J. 2018, 37, 608–621. [Google Scholar] [CrossRef]
  30. Harel, Y.; Ohad, I.; Kaplan, A. Activation of photosynthesis and resistance to photoinhibition in cyanobacteria within biological desert crust. Plant Physiol. 2004, 136, 3070–3079. [Google Scholar] [CrossRef] [Green Version]
  31. Ismaiel, M.M.; Piercey-Normore, M.D. Gene transcription and antioxidants production in Arthrospira (Spirulina) platensis grown under temperature variation. J. Appl. Microbiol. 2021, 130, 891–900. [Google Scholar] [CrossRef]
  32. Gao, X.; Zhu, Z.; Xu, H.; Liu, L.; An, J.; Ji, B.; Ye, S. Cold adaptation in drylands: Transcriptomic insights into cold-stressed Nostoc flagelliforme and characterization of a hypothetical gene with cold and nitrogen stress tolerance. Environ. Microbiol. 2021, 23, 713–727. [Google Scholar] [CrossRef]
  33. Los, D.A.; Murata, N. Responses to cold shock in cyanobacteria. J. Mol. Microbiol. Biotechnol. 1999, 1, 221–230. [Google Scholar]
  34. Martin, R.M.; Moniruzzaman, M.; Stark, G.F.; Gann, E.R.; Derminio, D.S.; Wei, B.; Hellweger, F.L.; Pinto, A.; Boyer, G.L.; Wilhelm, S.W. Episodic decrease in temperature increases mcy gene transcription and cellular microcystin in continuous cultures of Microcystis aeruginosa PCC 7806. Front. Microbiol. 2020, 11, 601864. [Google Scholar] [CrossRef]
  35. Singh, R.P.; Yadav, P.; Kujur, R.; Pandey, K.D.; Gupta, R.K. Cyanobacteria and salinity stress tolerance. In Cyanobacterial Lifestyle and its Applications in Biotechnology; Academic Press: Cambridge, MA, USA, 2022; pp. 253–280. [Google Scholar]
  36. Asthana, R.K.; Nigam, S.; Maurya, A.; Kayastha, A.M.; Singh, S.P. Trehalose-producing enzymes MTSase and MTHase in Anabaena 7120 under NaCl stress. Curr. Microbiol. 2008, 56, 429–435. [Google Scholar] [CrossRef]
  37. Kumar Srivastava, A.; Bhargava, P.; Mishra, Y.; Shukla, B.; Chand Rai, L. Effect of pretreatment of salt, copper and temperature on ultraviolet-B-induced antioxidants in diazotrophic cyanobacterium Anabaena doliolum. J. Basic Microbiol. 2006, 46, 135–144. [Google Scholar] [CrossRef]
  38. Apte, S.K.; Fernandes, T.A.; Iyer, V.; Alahari, A. Molecular basis of tolerance to salinity and drought stresses in photosynthetic nitrogen-fixing cyanobacteria. In Plant Molecular Biology and Biotechnology; Narosa Publications: New Delhi, India, 1997; pp. 258–268. [Google Scholar]
  39. Fernandes, T.A.; Iyer, V.; Apte, S.K. Differential responses of nitrogen-fixing cyanobacteria to salinity and osmotic stresses. Appl. Environ. Microbiol. 1993, 59, 899–904. [Google Scholar] [CrossRef]
  40. Schwartz, S.H.; Black, T.A.; Jäger, K.; Panoff, J.M.; Wolk, C.P. Regulation of an osmoticum-responsive gene in Anabaena sp. strain PCC 7120. J. Bacteriol. 1998, 180, 6332–6337. [Google Scholar] [CrossRef]
  41. Kuritz, T.; Bocanera, L.V.; Rivera, N.S. Dechlorination of lindane by the cyanobacterium Anabaena sp. strain PCC7120 depends on the function of the nir operon. J. Bacteriol. 1997, 179, 3368–3370. [Google Scholar] [CrossRef] [Green Version]
  42. Srivastava, A.; Biswas, S.; Yadav, S.; Kumar, S.; Srivastava, V.; Mishra, Y. Acute cadmium toxicity and post-stress recovery: Insights into coordinated and integrated response/recovery strategies of Anabaena sp. PCC 7120. J. Hazard. Mater. 2021, 411, 124822. [Google Scholar] [CrossRef]
  43. Nagao, R.; Yokono, M.; Ueno, Y.; Suzuki, T.; Kato, K.; Kato, K.H.; Tsuboshita, N.; Jiang, T.Y.; Dohmae, N.; Shen, J.R.; et al. Molecular organizations and function of iron-stress-induced-A protein family in Anabaena sp. PCC 7120. Biochim. Biophys. Acta. 2021, 1862, 148327. [Google Scholar] [CrossRef]
  44. Stuart, R.K.; Dupont, C.L.; Johnson, D.A.; Paulsen, I.T.; Palenik, B. Coastal strains of marine Synechococcus species exhibit increased tolerance to copper shock and a distinctive transcriptional response relative to those of open-ocean strains. Appl. Environ. Microbiol. 2009, 75, 5047–5057. [Google Scholar] [CrossRef] [Green Version]
  45. Higo, A.; Ikeuchi, M.; Ohmori, M. cAMP regulates respiration and oxidative stress during rehydration in Anabaena sp. PCC 7120. FEBS Lett. 2008, 582, 1883–1888. [Google Scholar] [CrossRef] [Green Version]
  46. Potts, M. Desiccation tolerance of prokaryotes. Microbiol. storage. Soil Biol. Biochem. 1994, 23, 313–322. [Google Scholar]
  47. Singh, H.; Anurag, K.; Apte, S.K. High radiation and desiccation tolerance of nitrogen-fixing cultures of the cyanobacterium Anabaena sp. strain PCC 7120 emanates from genome/proteome repair capabilities. Photos. Res. 2013, 118, 71–81. [Google Scholar] [CrossRef] [PubMed]
  48. Shivaji, S.; Kiran, M.D.; Chintalapati, S. Perception and transduction of low temperature in bacteria. Phys. Biochem. Extrem. 2007, 194, 207. [Google Scholar]
  49. Singh, R.P.; Yadav, P.; Kumar, A.; Hashem, A.; Al-Ajani, A.F.; Abd Allah, E.F.; Gupta, R.K. Physiological and Biochemical Responses of Bicarbonate Supplementation on Biomass and Lipid Content of Green Algae Scenedesmus sp. BHU1 Isolated from Wastewater for Renewable Biofuel Feedstock. Front. Microbiol. 2022, 13, 839800. [Google Scholar] [CrossRef] [PubMed]
  50. Fujita, Y.; Murakami, A.; Aizawa, K. Short-term and long-term adaptation of the photosynthetic apparatus: Homeostatic properties of thylakoids. In The Molecular Biology of Cyanobacteria; Bryant, A.D., Ed.; Kluwer Academic Publishers: Dordrecht, The Netherlands, 1994; pp. 667–692. [Google Scholar]
  51. Bennett, A.; Bogorad, L. Complementary chromatic adaptation in a filamentous blue-green alga. J. Cell Biol. 1973, 58, 419–435. [Google Scholar] [CrossRef] [PubMed]
  52. Dhankher, O.P. Arsenic metabolism in plants: An inside story. New Phytol. 2005, 168, 503–505. [Google Scholar] [CrossRef]
  53. Rai, A.N.; Mishra, A.K.; Tiwari, D.N. Cyanobacteria: From Basic Science to Applications; Academic Press: Cambridge, MA, USA, 2018. [Google Scholar]
  54. Los, D.A.; Suzuki, I.; Zinchenko, V.V.; Murata, N. Stress responses in Synechocystis: Regulated genes and regulatory systems. In The Cyanobacteria: Molecular Biology, Genomics and Evolution; Academic Press: Cambridge, MA, USA, 2008; Volume 117, p. 157. [Google Scholar]
  55. Jose, F.; Jeanjean, R.; Hagemann, M. Dynamics of the response of cyanobacteria to salt stress: Deciphering the molecular events. Physiol. Plant 1996, 96, 738–744. [Google Scholar] [CrossRef]
  56. Tanghe, A.; Van Dijck, P.; Thevelein, J.M. Determinants of freeze tolerance in microorganisms, physiological importance, and biotechnological applications. Adv. Appl. Microbiol. 2003, 53, 129–176. [Google Scholar]
  57. Potts, M. Mechanisms of desiccation tolerance in cyanobacteria. Eur. J. Phycol. 1999, 34, 319–328. [Google Scholar] [CrossRef]
  58. Hihara, Y.; Kamei, A.; Kanehisa, M.; Kaplan, A.; Ikeuchi, M. D.N.A. microarray analysis of cyanobacterial gene expression during acclimation to high light. Plant Cell 2001, 13, 793–806. [Google Scholar] [CrossRef] [Green Version]
  59. He, Y.Y.; Häder, D.P. Reactive oxygen species and UV-B: Effect on cyanobacteria. Photochem. Photobiol. Sci. 2002, 1, 729–736. [Google Scholar] [CrossRef]
  60. Latifi, A.; Ruiz, M.; Zhang, C.C. Oxidative stress in cyanobacteria. FEMS Microbiol. 2009, 33, 258–278. [Google Scholar] [CrossRef] [Green Version]
  61. Xia, S.; Wang, K.; Wan, L.; Li, A.; Hu, Q.; Zhang, C. Production, characterization, and antioxidant activity of fucoxanthin from the marine diatom Odontella aurita. Marine Drugs 2013, 11, 2667–2681. [Google Scholar] [CrossRef]
  62. He, Y.Y.; Häder, D.P. UV-B-induced formation of reactive oxygen species and oxidative damage of the cyanobacterium Anabaena sp.: Protective effects of ascorbic acid and N-acetyl-L-cysteine. Photochem. Photobiol. B Biol. 2002, 66, 115–124. [Google Scholar] [CrossRef]
  63. Rothshield, L.J. Microbes and radiation. In Enigmatic Microorganisms and Life in Extreme Environments; Seckbach, J., Ed.; Kluwer Academic Publishers: Dordrecht, The Netherlands, 1999; pp. 551–562. [Google Scholar]
  64. Singh, S.C.; Sinha, R.P.; Hader, D.P. Role of lipids and fatty acids in stress tolerance in cyanobacteria. Acta Protozool. 2002, 41, 297–308. [Google Scholar]
  65. Řezanka, T.; Nedbalová, L.; Elster, J.; Cajthaml, T.; Sigler, K. Very-long-chain iso and anteiso branched fatty acids in N-acylphosphatidylethanolamines from a natural cyanobacterial mat of Calothrix sp. Phytochemistry 2009, 70, 655–663. [Google Scholar] [CrossRef]
  66. Allakhverdiev, S.I.; Kinoshita, M.; Inaba, M.; Suzuki, I.; Murata, N. Unsaturated fatty acids in membrane lipids protect the photosynthetic machinery against salt-induced damage in Synechococcus. Plant Physiol. 2001, 125, 1842–1853. [Google Scholar] [CrossRef] [Green Version]
  67. Olie, J.J.; Potts, M. Purification and biochemical analysis of the cytoplasmic membrane from the desiccation-tolerant cyanobacterium Nostoc commune UTEX 584. Appl. Environ. 1986, 52, 706–710. [Google Scholar] [CrossRef]
  68. Sakurai, I.; Hagio, M.; Gombos, Z.; Tyystjarvi, T.; Paakkarinen, V.; Aro, E.M.; Wada, H. Requirement of phosphatidylglycerol for maintenance of photosynthetic machinery. Plant Physiol. 2003, 133, 1376–1384. [Google Scholar] [CrossRef] [Green Version]
  69. Adhikary, S.P. Heat shock proteins in the terrestrial epiphytic cyanobacterium Tolypothrix byssoidea. Biol. Plant. 2003, 47, 125–128. [Google Scholar] [CrossRef]
  70. Webb, R.; Sherman, L.A. The cyanobacterial heat-shock response and the molecular chaperones. In The Molecular Biology of Cyanobacteria; Springer: Dordrecht, The Netherlands, 1994; pp. 751–767. [Google Scholar]
  71. Billi, D.; Potts, M. Life and death of dried prokaryotes. Res. Microbiol. 2002, 153, 7–12. [Google Scholar] [CrossRef]
  72. Nienow, J.A.; McKay, C.P.; Friedmann, E.I. The cryptoendolithic microbial environment in the Ross Desert of Antarctica: Light in the photosynthetically active region. Microb. Ecol. 1988, 16, 271–289. [Google Scholar] [CrossRef] [PubMed]
  73. Nienow, J.A.; McKay, C.P.; Friedmann, E.I. Cryptoendolithic microbial environment in the Ross Desert of Antarctica: Mathematical models of the thermal regime. Microb. Ecol. 1988, 16, 253–270. [Google Scholar] [CrossRef] [PubMed]
  74. Aigner, S.; Herburger, K.; Holzinger, A.; Karsten, U. Epilithic Chamaesiphon (Synechococcales, Cyanobacteria) species in mountain streams of the Alps—Interspecific differences in photo-physiological traits. J. Appl. Psychol. 2018, 30, 1125–1134. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  75. Bhaya, D.; Schwarz, R.; Grossman, A.R. Molecular responses to environmental stress. In The Ecology of Cyanobacteria; Springer: Dordrecht, The Netherlands, 2000; pp. 397–442. [Google Scholar]
  76. Berner, T.; Sukenik, A. Photoacclimation in photosynthetic microorganisms: An ultrastructural response. Isr. J. Plant Sci. 1998, 46, 141–146. [Google Scholar] [CrossRef]
  77. Kawamura, M.; Mimuro, M.; Fujita, Y. Quantitative relationship between two reaction centers in the photosynthetic system of blue-green algae. Plant Cell Physiol. 1976, 20, 697–705. [Google Scholar]
  78. Elster, J.; Benson, E.E. Life in the polar terrestrial environment with a focus on algae and cyanobacteria. In Life in the Frozen State; Fuller, B.J., Lane, N., Benson, E.E., Eds.; CRC Press: Boca Raton, FL, USA, 2004; pp. 111–150. [Google Scholar]
  79. Bonilla, S.; Villeneuve, V.; Vincent, W.F. Benthic and planktonic algal communities in a High Arctic Lake: Pigment structure and contrasting responses to nutrient enrichment. J. Phycol. 2005, 41, 1120–1130. [Google Scholar] [CrossRef]
  80. Sasaki, A.; Mizuno, A.N. Partitioning light spectra: Adaptive stratification of phytobenthic communities in Antarctic lakes. J. Theor. Biol. 2017, 424, 1–10. [Google Scholar] [CrossRef]
  81. Six, C.; Thomas, J.C.; Brahamsha, B.; Lemoine, Y.; Partensky, F. Photophysiology of the marine cyanobacterium Synechococcus sp. WH8102, a new model organism. Aquatic Micro. Eco. 2004, 35, 17–29. [Google Scholar] [CrossRef]
  82. Grossman, A.R.; Schaefer, M.R.; Chang, G.G.; Collier, J.L. The response of cyanobacteria to environmental conditions: Light and nutrients. In The Molecular Biology of Cyanobacteria; Bryant, A., Ed.; Kluwer Academic Publishers: Dordrecht, The Netherlands, 1994; pp. 641–675. [Google Scholar]
  83. Toro, R.; Araya, C.; Labra, F.; Morales, L.; Morales, R.G. Trend and recovery of the total ozone column in South America and Antarctica. Clim. Dyn. 2017, 49, 3735–3752. [Google Scholar] [CrossRef]
  84. Häder, D.P.; Kumar, H.D.; Smith, R.C.; Worrest, R.C. Effects of solar U.V. radiation on aquatic ecosystems and interactions with climate change. Photochem. Photobiol. Sci. 2007, 6, 267–285. [Google Scholar] [CrossRef]
  85. Donker, V.; Häder, D.P. Effects of solar and ultraviolet radiation on motility, photo movement and pigmentation in filamentous, gliding cyanobacteria. FEMS Microbiol. Lett. 1991, 86, 159–168. [Google Scholar] [CrossRef]
  86. Schulz, E.M.; Scherer, S. U.V. protection in cyanobacteria. Eur. J. Phycol. 1999, 34, 329. [Google Scholar] [CrossRef]
  87. Seckbach, J.; Oren, A. Oxygenic photosynthetic microorganisms in extreme environments: Possibilities and limitations. In Algae and Cyanobacteria in Extreme Environments; Seckbach, J., Ed.; Springer: Dordrecht, The Netherlands, 2007; pp. 3–25. [Google Scholar]
  88. Pattanaik, B.; Schumann, R.; Karsten, U. Effects of ultraviolet radiation on cyanobacteria and their protective mechanisms. In Algae and Cyanobacteria in Extreme Environments; Springer: Dordrecht, The Netherlands, 2007; pp. 29–45. [Google Scholar]
  89. Quesada, A.; Vincent, W.F.; Lean, D.R. Community and pigment structure of Arctic cyanobacterial assemblages: The occurrence and distribution of UV-absorbing compounds. FEMS Microbiol. Ecol. 1999, 28, 315–323. [Google Scholar] [CrossRef]
  90. Vincent, W.F.; Mueller, D.R.; Bonilla, S. Ecosystems on ice: The microbial ecology of Markham Ice Shelf in the high Arctic. Cryobiology 2004, 48, 103–112. [Google Scholar] [CrossRef]
  91. Xue, L.; Zhang, Y.; Zhang, T.; An, L.; Wang, X. Effects of enhanced ultraviolet-B radiation on algae and cyanobacteria. Crit. Rev. Microbiol. 2005, 31, 79–89. [Google Scholar] [CrossRef]
  92. Donker, V.A.; Hader, D.P. Ultraviolet radiation effects on pigmentation in the cyanobacterium. Acta Protozool. 1997, 36, 49. [Google Scholar]
  93. Burnap, R.L.; Sherman, L.A. Deletion mutagenesis in Synechocystis sp. PCC6803 indicates that the manganese-stabilizing protein of photosystem II is not essential for oxygen evolution. Biochemistry 1991, 30, 440–446. [Google Scholar] [CrossRef]
  94. Kumar, A.; Sinha, R.P.; Häder, D.P. Effect of UV-B on enzymes of nitrogen metabolism in the cyanobacterium Nostoc calcicola. J. Plant Physiol. 1996, 148, 86–91. [Google Scholar] [CrossRef]
  95. Gao, X.; Ai, Y.F.; Qiu, B.S. Drought adaptation of a terrestrial macroscopic cyanobacterium, Nostoc flagelliforme, in arid areas: A review. Afr. J. Microbiol. Res. 2012, 6, 5728–5735. [Google Scholar]
  96. Schulz, M.E.; Schulz, S.; Wait, R.; Görg, A.; Scherer, S. The UV-B stimulon of the terrestrial cyanobacterium Nostoc commune comprises early shock proteins and late acclimation proteins. Mol. Microbiol. 2002, 46, 827–843. [Google Scholar] [CrossRef]
  97. Rai, S.; Pandey, S.; Shrivastava, A.K.; Singh, P.K.; Agrawal, C.; Rai, L.C. Understanding the mechanisms of abiotic stress management in cyanobacteria with special reference to proteomics. In Stress Biology of Cyanobacteria: Molecular Mechanisms to Cellular Responses; Srivastava, A.K., Rai, A.N., Neilan, B.A., Eds.; CRC Press: Boca Raton, FL, USA, 2013; Volume 93, p. 112. [Google Scholar]
  98. Pachauri, R.K.; Reisinger, A. (Eds.) Forth Assessment Report: Climate Change 2007; Intergovernmental Panel on Climate Change: Geneva, Switzerland, 2007; Volume 104. [Google Scholar]
  99. Sung, D.Y.; Kaplan, F.; Lee, K.J.; Guy, C.L. Acquired tolerance to temperature extremes. Trends Plant Sci. 2003, 8, 179–187. [Google Scholar] [CrossRef] [PubMed]
  100. Brock, D.T.; Madigan, T.D. Biology of Microorganisms, 7th ed.; Chapters 18 and 19; Prentice Hall: Englewood Cliffs, NJ, USA, 1994. [Google Scholar]
  101. Chintalapati, S.; Prakash, J.S.S.; Gupta, P.; Ohtani, S.; Suzuki, I.; Sakamoto, T.; Murata, N.; Shivaji, S. A novel Δ9 acyl-lipid desaturase, DesC2, from cyanobacteria acts on fatty acids esterified to the sn−2 position of glycerolipids. Biochem. J. 2006, 398, 207–214. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  102. Fort, D.G.; Herr, T.M.; Shaw, P.L.; Gutzman, K.E.; Starren, J.B. Mapping the evolving definitions of translational research. J. Clin. Transl. Sci. 2017, 1, 60–66. [Google Scholar] [CrossRef] [PubMed]
  103. Mishra, Y.; Bhargava, P.; Chaurasia, N.; Rai, L.C. Proteomic evaluation of the non-survival of Anabaena doliolum (Cyanophyta) at elevated temperatures. Eur. J. Phycol. 2009, 44, 551–565. [Google Scholar] [CrossRef] [Green Version]
  104. Slabs, A.R.; Suzuki, I.; Murata, N.; Simon, W.J.; Hall, J.J. Proteomic analysis of the heat shock response in Synechocystis PCC6803 and a thermally tolerant knockout strain lacking the histidine kinase 34 gene. Proteomics 2006, 6, 845–864. [Google Scholar] [CrossRef]
  105. Hongsthong, A.; Sirijuntarut, M.; Yutthanasirikul, R.; Senachak, J.; Kurdrid, P.; Cheevadhanarak, S.; Tanticharoen, M. Subcellular proteomic characterization of the high-temperature stress response of the cyanobacterium Spirulina platensis. Proteome Sci. 2009, 7, 1–19. [Google Scholar] [CrossRef]
  106. Psenner, R.; Sattler, B. Life at the freezing point. Science 1998, 280, 2073–2074. [Google Scholar] [CrossRef]
  107. Morgan-Kiss, R.M.; Priscu, J.C.; Pocock, T.; Gudynaite-Savitch, L.; Huner, N.P. Adaptation and acclimation of photosynthetic microorganisms to permanently cold environments. Microbiol. Mol. Biol. Rev. 2006, 70, 222–252. [Google Scholar] [CrossRef] [Green Version]
  108. Nishida, I.; Murata, N. Chilling sensitivity in plants and cyanobacteria: The crucial contribution of membrane lipids. Annu. Rev. Plant. Biol. 1996, 47, 541–568. [Google Scholar] [CrossRef] [Green Version]
  109. Sinetova, M.A.; Los, D.A. New insights in cyanobacterial cold stress responses: Genes, sensors, and molecular triggers. Biochim. Biophys. Acta 2016, 1860, 2391–2403. [Google Scholar] [CrossRef]
  110. Los, D.A.; Murata, N. Structure and expression of fatty acid desaturases. BBA-Lipids Lipid Metab. 1998, 1394, 3–15. [Google Scholar] [CrossRef]
  111. Inoue, N.; Taira, Y.; Emi, T.; Yamane, Y.; Kashino, Y.; Koike, H.; Satoh, K. Acclimation to the growth temperature and the high-temperature effects on photosystem II and plasma membranes in a mesophilic cyanobacterium, Synechocystis sp. PCC6803. Plant Cell Physiol. 2001, 42, 1140–1148. [Google Scholar] [CrossRef] [Green Version]
  112. Slabs, A.R.; Suzuki, I.; Simon, W.J. The heat shock response of Synechocystis sp. PCC 6803 analyzed by transcriptomics and proteomics. J. Exp. Bot. 2006, 57, 1573–1578. [Google Scholar]
  113. Giraldez-Ruiz, N.P.; Mateo, I.; Bonilla, I.; Fernandez-Pinas, F. The relationship between intracellular pH, growth characteristics and calcium in the cyanobacterium Anabaena sp. strain PCC7120 exposed to low pH. New Phytol. 1997, 137, 599–605. [Google Scholar] [CrossRef]
  114. Szabolcs, I. Soil and salinization. In Handbook of Plant and Crop Stress; Pessarakali, M., Ed.; Marcel Dekker: New York, NY, USA, 1994; p. 3. [Google Scholar]
  115. Csonka, L.N. Physiological and genetic responses of bacteria to osmotic stress. Microbiol. Rev. 1989, 53, 121–147. [Google Scholar] [CrossRef]
  116. Tijen, D.; Ismail, T. Exogenous glycinebetaine affects growth and proline accumulation and retards senescence in two rice cultivars under NaCl stress. Environ. Exp. Bot. 2006, 56, 72–79. [Google Scholar]
  117. Wei, Y.; Xu, X.; Tao, H.; Wang, P. Growth performance and physiological response in the halophyte Lycium barbarum grown at salt-affected soil. Ann. Appl. Biol. 2006, 149, 263–269. [Google Scholar] [CrossRef]
  118. Huang, F.; Fulda, S.; Hagemann, M.; Norling, B. Proteomic screening of salt-stress-induced changes in plasma membranes of Synechocystis sp. strain PCC 6803. Proteomics 2006, 6, 910–920. [Google Scholar] [CrossRef]
  119. Srivastava, A.K.; Bhargava, P.; Rai, L.C. Salinity and copper-induced oxidative damage and changes in the antioxidative defence systems of Anabaena doliolum. World J. Microbiol. Biotechnol. 2005, 21, 1291–1298. [Google Scholar] [CrossRef]
  120. Singh, R.N. Reclamation of ‘Usar’ lands in India through blue-green algae. Nature 1950, 165, 325–326. [Google Scholar] [CrossRef]
  121. Kanesaki, Y.; Suzuki, I.; Allakhverdiev, S.I.; Mikami, K.; Murata, N. Salt stress and hyperosmotic stress regulate the expression of different sets of genes in Synechocystis sp. PCC 6803. Biochem. Biophys. 2002, 290, 339–348. [Google Scholar] [CrossRef] [PubMed]
  122. Molitor, V.; Erber, W.; Peschek, G.A. Increased levels of cytochrome oxidase and sodium-proton antiporter in the plasma membrane of Anacystis nidulans after growth in sodium-enriched media. FEBS Lett. 1986, 204, 251–256. [Google Scholar] [CrossRef]
  123. Fulda, S.; Huckauf, J.; Schoor, A.; Hagemann, M. Analysis of stress responses in the cyanobacterial strains Synechococcus sp. PCC 7942, Synechocystis sp. PCC 6803, and Synechococcus sp. PCC 7418: Osmolyte accumulation and stress protein synthesis. J. Plant Physiol. 1999, 154, 240–249. [Google Scholar] [CrossRef]
  124. Fulda, S.; Huang, F.; Nilsson, F.; Hagemann, M.; Norling, B. Proteomics of Synechocystis sp. strain PCC 6803: Identification of periplasmic proteins in cells grown at low and high salt concentrations. Eur. J. Biochem. 2000, 267, 5900–5907. [Google Scholar] [CrossRef] [PubMed]
  125. Koch, M.; Orthwein, T.; Alford, J.T.; Forchhammer, K. The Slr0058 protein from Synechocystis sp. PCC 6803 is a novel regulatory protein involved in PHB granule formation. Front. Microbiol. 2020, 11, 809. [Google Scholar] [CrossRef]
  126. Allakhverdiev, S.I.; Nishiyama, Y.; Miyairi, S.; Yamamoto, H.; Inagaki, N.; Kanesaki, Y.; Murata, N. Salt stress inhibits the repair of photodamaged photosystem II by suppressing the transcription and translation of psbA genes in Synechocystis. Plant Physiol. 2002, 130, 1443–1453. [Google Scholar] [CrossRef] [Green Version]
  127. Apte, S.K.; Fernandes, T.; Badran, H.; Ballal, A. Expression and possible role of stress-responsive proteins in Anabaena. J. Biosci. 1998, 23, 399–406. [Google Scholar] [CrossRef]
  128. Apte, S.K.; Reddy, B.R.; Thomas, J. Relationship between sodium influx and salt tolerance of nitrogen-fixing cyanobacteria. Appl. Environ. Microbiol. 1987, 53, 1934–1939. [Google Scholar] [CrossRef] [Green Version]
  129. Chen, L.Z.; Li, D.H.; Song, L.R.; Hu, C.X.; Wang, G.H.; Liu, Y.D. Effects of salt stress on carbohydrate metabolism in desert soil alga Microcoleus vaginatus Gom. J. Integr. Plant Biol. 2006, 48, 914–919. [Google Scholar] [CrossRef]
  130. Eisenhut, M.; Kahlon, S.; Hasse, D.; Ewald, R.; Lieman-Hurwitz, J.; Ogawa, T.; Ruth, W.; Bauwe, H.; Kaplan, A.; Hagemann, M. The plant-like C2 glycolate cycle and the bacterial-like glycerate pathway cooperate in phosphoglycolate metabolism in cyanobacteria. Plant Physiol. 2006, 142, 333–342. [Google Scholar] [CrossRef] [Green Version]
  131. Srivastava, A.K.; Bhargava, P.; Thapar, R.; Rai, L.C. Salinity-induced physiological and proteomic changes in Anabaena doliolum. Environ. Exp. Bot. 2008, 64, 49–57. [Google Scholar] [CrossRef]
  132. Hejduková, E.; Elster, J.; Nedbalová, L. Annual cycle of freshwater diatoms in the High Arctic revealed by multiparameter fluorescent staining. Microb. Ecol. 2020, 80, 559–572. [Google Scholar] [CrossRef]
  133. Alpert, P. The limits and frontiers of desiccation-tolerant life. Integer. Comp. Biol. 2005, 45, 685–695. [Google Scholar] [CrossRef] [Green Version]
  134. Shirkey, B.; McMaster, N.J.; Smith, S.C.; Wright, D.J.; Rodriguez, H.; Jaruga, P.; Birincioglu, M.; Helm, R.F.; Potts, M. Genomic D.N.A. of Nostoc commune (Cyanobacteria) becomes covalently modified during long-term (decades) desiccation but is protected from oxidative damage and degradation. Nucleic Acids Res. 2003, 31, 2995–3005. [Google Scholar] [CrossRef] [Green Version]
  135. Lipman, C.B. The successful revival of Nostoc commune from a herbarium specimen eighty-seven years old. Bull. Torrey Bot. Club 1941, 68, 664–666. [Google Scholar] [CrossRef]
  136. Cameron, R.E. Species of Nostoc vaucher occurring in the Sonoran Desert in Arizona. Trans. Am. Microsc. Soc. 1962, 81, 379–384. [Google Scholar] [CrossRef]
  137. Scheibe, R.; Beck, E. Drought, desiccation, and oxidative stress. In Plant Desiccation Tolerance; Springer: Berlin/Heidelberg, Germany, 2011; pp. 209–231. [Google Scholar]
  138. Palmer, R.J.; Friedmann, E.I. Water relations and photosynthesis in the cryptoendolithic microbial habitat of hot and cold deserts. Microb. Ecol. 1990, 19, 111–118. [Google Scholar] [CrossRef]
  139. Borisov, V.B.; Siletsky, S.A.; Nastasi, M.R.; Forte, E. R.O.S. defense systems and terminal oxidases in bacteria. Antioxidants 2021, 10, 839. [Google Scholar] [CrossRef]
  140. Smith, M.G. Survival of E. coli and Salmonella after chilling and freezing in liquid media. J. Food Sci. 1995, 60, 509–512. [Google Scholar] [CrossRef]
  141. Prát, S. Ustoychivosttermalnykhsinezelenykhvodorosley k nizkimtemperaturam [Stability of thermal blue-green algae to low temperatures]. Bull. Int. Acad. Tcheque Sci. 1950, 20, 1–6. [Google Scholar]
  142. Ninagawa, T.; Eguchi, A.; Kawamura, Y.; Konishi, T.; Narumi, A. A study on ice crystal formation behavior at intracellular freezing of plant cells using a high-speed camera. Cryobiology 2016, 73, 20–29. [Google Scholar] [CrossRef] [PubMed]
  143. Cockell, C.S.; Stokes, M.D.; Korsmeyer, K.E. Overwintering strategies of Antarctic organisms. Environ. Rev. 2000, 8, 1–19. [Google Scholar] [CrossRef]
  144. Tchounwou, P.B.; Yedjou, C.G.; Patlolla, A.K.; Sutton, D.J. Heavy metal toxicity and the environment. Mol. Clin. Environ. Toxicol. 2012, 101, 133–164. [Google Scholar]
  145. Hantke, K. Bacterial zinc uptake and regulators. Curr. Opin. Microbiol 2005, 8, 196–202. [Google Scholar] [CrossRef] [PubMed]
  146. Xu, J.; Tian, Y.S.; Peng, R.H.; Xiong, A.S.; Zhu, B.; Hou, X.L.; Yao, Q.H. Cyanobacteria MT gene SmtA enhance zinc tolerance in Arabidopsis. Mol. Biol. Rep. 2010, 37, 1105–1110. [Google Scholar] [CrossRef]
  147. Nagalakshmi, N.; Prasad, M.N.V. Responses of glutathione cycle enzymes and glutathione metabolism to copper stress in Scenedesmus bijugatus. Plant Sci. 2001, 160, 291–299. [Google Scholar] [CrossRef]
  148. Yadav, P.; Singh, R.P.; Gupta, R.K. Role of cyanobacteria in germination and growth of paddy seedlings. Int. J. Phytol. Res. 2022, 2, 11–18. [Google Scholar]
  149. Wang, Z.; Li, D.; Li, G.; Liu, Y. Mechanism of photosynthetic response in Microcystis aeruginosa PCC7806 to low inorganic phosphorus. Harmful Algae. 2010, 9, 613–619. [Google Scholar] [CrossRef]
  150. Blindauer, C.A. Bacterial metallothioneins: Past, present, and questions for the future. J. Biol. Inorg. Chem. 2011, 16, 1011–1024. [Google Scholar] [CrossRef]
  151. Murata, N.; Suzuki, I. Exploitation of genomic sequences in a systematic analysis to access how cyanobacteria sense environmental stress. J. Exp. Bot. 2006, 57, 235–247. [Google Scholar] [CrossRef]
  152. Bhargava, P.; Mishra, Y.; Srivastava, A.K.; Narayan, O.P.; Rai, L.C. Excess copper induces anoxygenic photosynthesis in Anabaena doliolum: A homology-based proteomic assessment of its survival strategy. Photos. Res. 2008, 96, 61–74. [Google Scholar] [CrossRef]
  153. Pandey, S.; Rai, R.; Rai, L.C. Proteomics combines morphological, physiological and biochemical attributes to unravel the survival strategy of Anabaena sp. PCC7120 under arsenic stress. J. Proteom. 2012, 75, 921–937. [Google Scholar] [CrossRef]
  154. Yang, F.; Liao, D.; Wu, X.; Gao, R.; Fan, Y.; Raza, M.A.; Wang, X.; Yong, T.; Liu, W.; Liu, J.; et al. Effect of aboveground and belowground interactions on the intercrop yields in maize-soybean relay intercropping systems. Field Crops Res. 2017, 203, 16–23. [Google Scholar] [CrossRef]
  155. Chardonnet, S.; Sakr, S.; Cassier-Chauvat, C.; Le Maréchal, P.; Chauvat, F.; Lemaire, S.D.; Decottignies, P. First proteomic study of S-glutathionylation in cyanobacteria. J. Proteome Res. 2015, 14, 59–71. [Google Scholar] [CrossRef]
  156. Babele, P.K.; Kumar, J.; Chaturvedi, V. Proteomic deregulation in cyanobacteria in response to abiotic stresses. Front. Microbiol. 2019, 10, 1315. [Google Scholar] [CrossRef]
  157. Graves, P.R.; Haystead, T.A. Molecular biologist’s guide to proteomics. Microbiol. Mol. Biol. Rev. 2002, 66, 39–63. [Google Scholar] [CrossRef] [Green Version]
  158. Monteoliva, L.; Albar, J.P. Differential proteomics: An overview of gel and non-gel based approaches. Brief. Funct. Genom. 2004, 3, 220–239. [Google Scholar] [CrossRef] [Green Version]
  159. Kashino, Y.; Lauber, W.M.; Carroll, J.A.; Wang, Q.; Whitmarsh, J.; Satoh, K.; Pakrasi, H.B. Proteomic analysis of a highly active photosystem II preparation from the cyanobacterium Synechocystis sp. PCC 6803 reveals the presence of novel polypeptides. Biochemistry 2002, 41, 8004–8012. [Google Scholar] [CrossRef]
  160. Wang, Y.; Sun, J.; Chitnis, P.R. Proteomic study of the peripheral proteins from thylakoid membranes of the cyanobacterium Synechocystis sp. PCC 6803. Electrophoresis 2000, 21, 1746–1754. [Google Scholar] [CrossRef]
  161. Wang, H.; Yang, Y.; Chen, W.; Ding, L.; Li, P.; Zhao, X.; Wang, X.; Li, A.; Bao, Q. Identification of differentially expressed proteins of Arthrospira (Spirulina) plantensis-YZ under salt-stress conditions by proteomics and qRT-PCR analysis. Proteome Sci. 2013, 11, 6. [Google Scholar] [CrossRef] [Green Version]
  162. Panda, B.; Basu, B.; Rajaram, H.; Apte, S.K. Comparative proteomics of oxidative stress response in three cyanobacterial strains native to Indian paddy fields. J. Proteomics 2015, 127, 152–160. [Google Scholar] [CrossRef] [PubMed]
  163. Panda, B.; Basu, B.; Rajaram, H.; Kumar Apte, S. Methyl viologen responsive proteome dynamics of Anabaena sp. strain PCC7120. Proteomics 2014, 14, 1895–1904. [Google Scholar] [CrossRef] [PubMed]
  164. Hikari, J.; Österholm, J.; Kopf, M.; Battchikova, N.; Wahlsten, M.; Aro, E.M.; Hess, W.R.; Sivonen, K. Transcriptomic and proteomic profiling of Anabaena sp. strain 90 under inorganic phosphorus stress. Appl. Environ. Microbiol. 2015, 81, 5212–5222. [Google Scholar]
  165. Shrivastava, A.K.; Chatterjee, A.; Yadav, S.; Singh, P.K.; Singh, S.; Rai, L.C. UV-B stress-induced metabolic rearrangements explored with comparative proteomics in three Anabaena species. J. Proteom. 2015, 121, 122–133. [Google Scholar] [CrossRef] [PubMed]
  166. Babele, P.K.; Singh, G.; Kumar, A.; Tyagi, M.B. Induction and differential expression of certain novel proteins in Anabaena L31 under UV-B radiation stress. Front. Microbiol. 2015, 6, 133. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  167. Wang, Y.; Chen, L.; Zhang, W. Proteomic and metabolomic analyses reveal metabolic responses to 3-hydroxy propionic acid synthesized internally in cyanobacterium Synechocystis sp. PCC 6803. Biotechnol. Biofuels 2016, 9, 209. [Google Scholar] [CrossRef]
  168. Mehta, A.; López-Maury, L.; Florencio, F.J. Proteomic pattern alterations of the cyanobacterium Synechocystis sp. PCC 6803 in response to Cadmium, nickel and cobalt. J. Proteomics 2014, 102, 98–112. [Google Scholar] [CrossRef] [Green Version]
  169. Qiao, J.; Wang, J.; Chen, L.; Tian, X.; Huang, S.; Ren, X.; Zhang, W. Quantitative iTRAQ LC-MS/MS proteomics reveals metabolic responses to biofuel ethanol in cyanobacterial Synechocystis sp. PCC 6803. J. Proteome Res. 2012, 11, 5286–5300. [Google Scholar] [CrossRef]
  170. Tian, X.; Chen, L.; Wang, J.; Qiao, J.; Zhang, W. Quantitative proteomics reveals dynamic responses of Synechocystis sp. PCC 6803 to next generation biofuel butanol. J. Proteomics 2013, 78, 326–345. [Google Scholar]
  171. Singh, A.K.; Singh, P.P.; Tripathi, V.; Verma, H.; Singh, S.K.; Srivastava, A.K.; Kumar, A. Distribution of cyanobacteria and their interactions with pesticides in paddy field: A comprehensive review. J. Environ. Manag. 2018, 224, 361–375. [Google Scholar] [CrossRef]
  172. Hashtroudi, M.S.; Ghassempour, A.; Riahi, H.; Shariatmadari, Z.; Khanjir, M. Endogenous auxins in plant growth-promoting Cyanobacteria—Anabaena vaginicola and Nostoc calcicola. J. Appl. Phycol. 2013, 25, 379–386. [Google Scholar] [CrossRef]
  173. Haroun, S.A.; Hussein, M.H. The Promotive Effect of Algal Biofertilizers on Growth, Protein Pattern and Some Metabolic Activities of Lupinus termis Plants Grown in Siliceous Soil. Asian J. Plant Sci. 2003, 2, 944–951. [Google Scholar] [CrossRef] [Green Version]
  174. Singh, D.P.; Prabha, R.; Verma, S.; Meena, K.K.; Yandigeri, M. Antioxidant properties and polyphenolic content in terrestrial cyanobacteria. 3 Biotech. 2017, 7, 134. [Google Scholar] [CrossRef] [Green Version]
  175. Singh, S. A review on possible elicitor molecules of cyanobacteria: Their role in improving plant growth and providing tolerance against biotic or abiotic stress. J. Appl. Microbiol. 2014, 117, 1221–1244. [Google Scholar] [CrossRef]
  176. Panda, D.; Pramanik, K.; Nayak, B.R. Use of seaweed extracts as plant growth regulators for sustainable agriculture. Int. J. Bio-resour. Stress Manag. 2012, 3, 404–411. [Google Scholar]
  177. Errani, A.; Nardi, S.; Francioso, O.; Sanchez-Cortes, S.; Foggia, M.D.; Schiavon, M. Effects of two protein hydrolysates obtained from Chickpea (Cicer arietinum L.) and Spirulina platensis on Zea mays (L.) plants. Front. Plant Sci. 2019, 10, 954. [Google Scholar]
  178. Singh, D.P.; Prabha, R.; Yandigeri, M.S.; Arora, D.K. Cyanobacteria-mediated phenylpropanoids and phytohormones in rice (Oryza sativa) enhance plant growth and stress tolerance. Antonie Van Leeuwenhoek 2011, 100, 557–568. [Google Scholar] [CrossRef]
  179. Li, Z.; Yuan, S.; Jia, H.; Gao, F.; Zhou, M.; Yuan, N.; Wu, P.; Hu, Q.; Sun, D.; Luo, H. Ectopic expression of a cyanobacterial flavodoxin in creeping bentgrass impacts plant development and confers broad abiotic stress tolerance. Plant Biotechnol. 2017, 15, 433–446. [Google Scholar] [CrossRef]
  180. Takabe, T.; Uchida, A.; Shinagawa, F.; Terada, Y.; Kajita, H.; Tanaka, Y.; Takabe, T.; Hayashi, T.; Kawai, T.; Takabe, T. Overexpression of DnaK from a halotolerant cyanobacterium Aphanothece halophytic enhances growth rate as well as abiotic stress tolerance of poplar plants. Plant Growth Regul. 2008, 56, 265–273. [Google Scholar] [CrossRef]
  181. Gerasymenko, I.M.; Sakhno, L.A.; Kyrpa, T.N.; Ostapchuk, A.M.; Hadjiev, T.A.; Goldenkova-Pavlova, I.V.; Sheludko, Y.V. Characterization of Nicotiana tabacum plants expressing hybrid genes of cyanobacterial Δ9 or Δ12 acyl-lipid desaturases and thermostable lichenase. Russ. J. Plant Physiol. 2015, 62, 283–291. [Google Scholar] [CrossRef]
  182. Cui, L.; Liu, Y.; Yang, Y.; Ye, S.; Luo, H.; Qiu, B.; Gao, X. The drnf1 gene from the drought-adapted cyanobacterium Nostoc flagelliforme improved salt tolerance in transgenic Synechocystis and Arabidopsis plant. Genes 2018, 9, 441. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  183. Shariatmadari, Z.; Riahi, H.; Abdi, M.; Hashtroudi, M.S.; Ghassempour, A.R. Impact of cyanobacterial extracts on the growth and oil content of the medicinal plant Mentha piperita L. J. Appl. Phycol. 2015, 27, 2279–2287. [Google Scholar] [CrossRef]
  184. Chittapun, S.; Limbipichai, S.; Amnuaysin, N.; Boonkerd, R.; Charoensook, M. Effects of using cyanobacteria and fertilizer on growth and yield of rice, Pathum Thani I: A pot experiment. J. Appl. Soc. Psychol. 2018, 30, 79–85. [Google Scholar] [CrossRef]
  185. Jan, Z.; Ali, S.; Sultan, T.; Ahmad, W. The role of cyanobacteria in the availability of major plant nutrients and soil organic matter to rice crop under saline soil conditions. Sarhad J. Agric. 2017, 33, 566–572. [Google Scholar] [CrossRef]
  186. Singh, N.K.; Dhar, D.W. Cyanobacterial reclamation of salt-affected soil. In Genetic Engineering, Biofertilization, Soil Quality and Organic Farming; Springer: Dordrecht, The Netherlands, 2010; pp. 243–275. [Google Scholar]
  187. Rodríguez, A.A.; Stella, A.M.; Storni, M.M.; Zulpa, G.; Zaccaro, M.C. Effects of cyanobacterial extracellular products and gibberellic acid on salinity tolerance in Oryza sativa L. Saline Syst. 2006, 2, 7. [Google Scholar] [CrossRef] [Green Version]
  188. Ghada, S.F.; Ahmed, D.A. Improved soil characteristics and wheat germination as influenced by inoculation of Nostoc kihlmani and Anabaena cylindrica. Rend. Lincei. 2015, 26, 121–131. [Google Scholar] [CrossRef]
  189. Manchanda, H. Influence of cyanobacterial filtrate on growth of rice seedlings under saline conditions. IJRAR 2018, 5, 19–22. [Google Scholar]
  190. Abbas, H.H.; Ali, M.E.; Ghazal, F.M.; El-Gaml, N.M. Impact of Cyanobacteria inoculation on rice (Oryza sativa) yield cultivated in saline soil. Am. J. Sci. 2015, 11, 13–19. [Google Scholar]
  191. Arora, M.; Kaushik, A.; Rani, N.; Kaushik, C.P. Effect of cyanobacterial exopolysaccharides on salt stress alleviation and seed germination. J. Environ. Biol. 2010, 31, 701–704. [Google Scholar]
  192. Coba de la Pena, T.; Redondo, F.J.; Manrique, E.; Lucas, M.M.; Pueyo, J.J. Nitrogen fixation persists under conditions of salt stress in transgenic Medicago truncatula plants expressing a cyanobacterial flavodoxin. Plant Biotechnol. J. 2010, 8, 954–965. [Google Scholar] [CrossRef]
  193. Manchanda, R.K. Significance of prognostic factor research in clinical verification. Indian J. Res. Homoeopathy 2018, 12, 1. [Google Scholar] [CrossRef]
  194. Muñoz-Rojas, M.; Chilton, A.; Liyanage, G.S.; Erickson, T.E.; Merritt, D.J.; Neilan, B.A.; Ooi, M.K.J. Effects of indigenous soil cyanobacteria on seed germination and seedling growth of arid species used in restoration. ISO4 2018, 429, 91–100. [Google Scholar] [CrossRef]
  195. Chua, M.; Erickson, T.E.; Merritt, D.J.; Chilton, A.M.; Ooi, M.K.; Muñoz-Rojas, M. Bio-priming seeds with cyanobacteria: Effects on native plant growth and soil properties. Restore Ecol. 2020, 28, S168–S176. [Google Scholar] [CrossRef]
  196. Tognetti, V.B.; Palatnik, J.F.; Fillat, M.F.; Melzer, M.; Hajirezaei, M.R.; Valle, E.M.; Carrillo, N. Functional replacement of ferredoxin by a cyanobacterial flavodoxin in tobacco confers broad-range stress tolerance. Plant Cell 2006, 18, 2035–2050. [Google Scholar] [CrossRef] [Green Version]
  197. Gharechahi, J.; Hajirezaei, M.R.; Salekdeh, G.H. Comparative proteomic analysis of tobacco expressing cyanobacterial flavodoxin and its wild type under drought stress. J. Plant Physiol. 2015, 175, 48–58. [Google Scholar] [CrossRef]
  198. Giró, M.; Ceccoli, R.D.; Poli, H.O.; Carrillo, N.; Lodeiro, A.F. An in vivo system involving co-expression of cyanobacterial flavodoxin and ferredoxin–NADP+ reductase confers increased tolerance to oxidative stress in plants. FEBS Open Bio 2011, 1, 7–13. [Google Scholar] [CrossRef] [Green Version]
  199. Mohsen, A.; Dowidar, S.; Abo-Hamad, S.; Khalaf, B. Role of cyanobacteria in amelioration of toxic effects of copper in ‘Trigonella foenumgracum’. Aust. J. Crop Sci. 2013, 7, 1488–1493. [Google Scholar]
  200. Tripathi, R.D.; Dwivedi, S.; Shukla, M.K.; Mishra, S.; Srivastava, S.; Singh, R.; Rai, U.N.; Gupta, D.K. Role of blue-green algae biofertilizer in ameliorating the nitrogen demand and fly-ash stress to the growth and yield of rice (Oryza sativa L.) plants. Chemosphere 2008, 70, 1919–1929. [Google Scholar] [CrossRef]
  201. Xu, J.; Xu, X.; Xie, S.Q. A comprehensive review on recent developments in quality function deployment. Int. J. Product. Qual. Manag. 2010, 6, 457–494. [Google Scholar] [CrossRef]
  202. Bhagwat, A.A.; Apte, S.K. Comparative analysis of proteins induced by heat shock, salinity, and osmotic stress in the nitrogen-fixing cyanobacterium Anabaena sp. strain L-31. J. Bacteriol. 1989, 171, 5187–5189. [Google Scholar] [CrossRef] [Green Version]
  203. Seifikalhor, M.; Hassani, S.B.; Aliniaeifard, S. Seed priming by cyanobacteria (Spirulina platensis) and salep gum enhances tolerance of maize plant against cadmium toxicity. J. Plant Growth Regul. 2020, 39, 1009–1021. [Google Scholar] [CrossRef]
Figure 1. The overview of strategies, followed by cyanobacterial strains to cope with different abiotic stresses.
Figure 1. The overview of strategies, followed by cyanobacterial strains to cope with different abiotic stresses.
Stresses 02 00036 g001
Figure 2. The overview of cyanobacterial survival strategies to cope with Ultraviolet radiation.
Figure 2. The overview of cyanobacterial survival strategies to cope with Ultraviolet radiation.
Stresses 02 00036 g002
Figure 3. The overview of cyanobacterial response after varying temperature conditions.
Figure 3. The overview of cyanobacterial response after varying temperature conditions.
Stresses 02 00036 g003
Figure 4. Schematic representation of the general mechanism of desiccation tolerance by cyanobacteria.
Figure 4. Schematic representation of the general mechanism of desiccation tolerance by cyanobacteria.
Stresses 02 00036 g004
Table 1. List of abiotic stress tolerance in cyanobacteria with their references.
Table 1. List of abiotic stress tolerance in cyanobacteria with their references.
S. No.Types of StressesReferences
1Nutrient deficiency[19,20]
2Light intensity[21,22,23]
3RadiationU.V.[24,25,26]
Gamma[27,28,29]
4TemperatureHigh[30,31]
Low[32,33,34]
5SaltSalinity[35,36,37]
Osmoticum[38,39,40]
6ChemicalsPesticides[41]
Heavy metals[42,43,44]
7Desiccation[45,46,47]
Table 2. Proteome analyses in response to different abiotic stresses in cyanobacteria.
Table 2. Proteome analyses in response to different abiotic stresses in cyanobacteria.
Cyanobacterial StrainStressMethodologiesProteins
Identified
References
Anabaena sp. PCC 7120Arsenic2-DE, MALDI-TOF/
MS, RT-PCR
45[153]
Arthrospira plantensis-YZSalt2DE, MALDI-TOF/
TOF, qRT-PCR
141[161]
Anabaena
Doliolum, Anabaena PCC7120, and Anabaena L-31
Methyl viologen2DE, MALDI-TOF
MS/MS
103, 92, and 41,
respectively
[162,163]
Anabaena sp. strain 90Inorganic phosphorus (Pi)2D-DIGE and
LC-MS/MS
43[164]
Anabaena L31,
Anabaena doliolum, and Anabaena sp. PCC 7120
UV-B stress2DE, MALDI-TOF
MS/MS
90, 91, and 98,
respectively
[165]
Anabaena sp. strain L31UV-B2DE, MALDI TOF
MS/MS,
Bioinformatics
21[166]
Synechocystis sp. PCC 68033-hydroxypro
pionic acid
(3-HP)
iTRAQ-LC
-MS/MS
L.C.–MS-based targeted
metabolomics
2264[167]
Synechocystis sp. PCC 6803Cobalt, cadmium, and nickel2-DE MALDI
TOF/MS RT-PCR
20 (Nickel) 26
(Cobalt) 13
(Cadmium)
[168]
Synechocystis sp. PCC 6803ButanoliTRAQ and
LC-MS/MS
303[169,170]
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Yadav, P.; Singh, R.P.; Rana, S.; Joshi, D.; Kumar, D.; Bhardwaj, N.; Gupta, R.K.; Kumar, A. Mechanisms of Stress Tolerance in Cyanobacteria under Extreme Conditions. Stresses 2022, 2, 531-549. https://doi.org/10.3390/stresses2040036

AMA Style

Yadav P, Singh RP, Rana S, Joshi D, Kumar D, Bhardwaj N, Gupta RK, Kumar A. Mechanisms of Stress Tolerance in Cyanobacteria under Extreme Conditions. Stresses. 2022; 2(4):531-549. https://doi.org/10.3390/stresses2040036

Chicago/Turabian Style

Yadav, Priya, Rahul Prasad Singh, Shashank Rana, Diksha Joshi, Dharmendra Kumar, Nikunj Bhardwaj, Rajan Kumar Gupta, and Ajay Kumar. 2022. "Mechanisms of Stress Tolerance in Cyanobacteria under Extreme Conditions" Stresses 2, no. 4: 531-549. https://doi.org/10.3390/stresses2040036

Article Metrics

Back to TopTop