Skip to main content

Proteomics with a pinch of salt: A cyanobacterial perspective

Abstract

Cyanobacteria are ancient life forms and have adapted to a variety of extreme environments, including high salinity. Biochemical, physiological and genetic studies have contributed to uncovering their underlying survival mechanisms, and as recent studies demonstrate, proteomics has the potential to increase our overall understanding further. To date, most salt-related cyanobacterial proteomic studies have utilised gel electrophoresis with the model organism Synechocystis sp. PCC6803. Moreover, focus has been on 2–4% w/v NaCl concentrations within different cellular compartments. Under these conditions, Synechocystis sp. PCC6803 was found to respond and adapt to salt stress through synthesis of general and specific stress proteins, altering the protein composition of extracellular layers, and re-directing control of complex central intermediary pathways. Post-transcriptional control was also predicted through non-correlating transcript level data and identification of protein isoforms.

In this paper, we also review technical developments with emphasis on improving the quality and quantity of proteomic data and overcoming the detrimental effects of salt on sample preparation and analysis. Developments in gel-free methods include protein and peptide fractionation workflows, which can increase coverage of the proteome (20% in Synechocystis sp. PCC6803). Quantitative techniques have also improved in accuracy, resulting in confidence in quantitation approaching or even surpassing that seen in transcriptomic techniques (better than 1.5-fold in differential expression). Furthermore, in vivo metabolic labelling and de novo protein sequencing software have improved the ability to apply proteomics to unsequenced environmental isolates. The example used in this review is a cyanobacterium isolated from a Saharan salt lake.

Review

1.0 Introduction

Increasing salinity, believed to affect nearly one-fifth of the world's irrigated land, is a major factor impairing worldwide agricultural productivity [1, 2]. Furthermore, the problem is predicted to get considerably worse over the next 30 to 50 years [3]. Despite the importance of this problem, compounding effects of salinity and associated environmental stresses are not fully understood, particularly at the molecular level.

Cyanobacteria, formerly referred to as blue-green algae, are oxygenic phototrophic bacteria [4] which possess an internal membrane system, the photosynthetic thylakoid membrane. They are a multifarious group of organisms known to have colonised a wide range of ecosystems including soil, air, dry rock and aquatic systems [5]. Many species are capable of not only surviving, but thriving in conditions previously thought to be inhabitable, tolerating desiccation, high temperatures, extreme pH and high salinity, illustrating their capacity to acclimate to extreme environments [6]. Cyanobacteria have been classified into three groups relating to their salt tolerance, salt sensitive (or stenohaline), moderately halotolerant, and extremely halotolerant [7]. Example genera for each group include Anabaena, Synechocystis and Aphanothece, respectively.

Understanding how cyanobacteria acclimate to saline environments has been a source of interest, fuelled by two main drivers, agriculture and biotechnology. Firstly, comparable to higher plants, cyanobacteria perform oxygen evolving photosynthesis, and therefore they share many functional pathways. It is relatively easy to culture many cyanobacterial species and additional advantages such as rapid growth rate and low culture costs make them ideal models for investigating metabolic processes such as photosynthesis and respiration. Cyanobacteria therefore make suitable models for studying the physiology of salt tolerance and this has provided valuable insights into revealing the nature in which salinity prevents crop plant species from using aquatic resources [3, 8].

From a non-agricultural perspective, some of the osmotic compounds produced in response to high salt have the potential to play important roles in biotechnology and medicine. Osmotic compounds (also known as compatible solutes) are low-molecular mass, uncharged, hydrophilic molecules that do not interfere with cell metabolism. These compatible solutes help restore osmotic balance with surroundings and maintain membrane integrity and protein stability [9, 10]. Cyanobacteria are able to synthesise their own compatible solutes, as well as uptake them from the surroundings. They help stabilise and even enhance protein activity, for example, polyethyl glycols improving crystallization of proteins [11]. In medicine, trimethylamine N-oxide has the capability of in vitro rescue of cystic fibrosis (misfolded) proteins [12]. They are also used to protect cells from stress in the cosmetics industry, for example, ectoine has been shown to protect skin from UVA-induced damage [13, 14].

Biochemical, physiological and genetic studies have revealed how cyanobacteria are able to adapt to high salt environments, initially by active ion export followed by accumulation of compatible solutes [5, 15, 16]. However, the discovery of salt-regulated genes in cyanobacteria has been accelerated by recent progress in functional and structural genomics. The genomes of over 28 cyanobacteria are now fully sequenced (with at least 72 in progress) according to the Genomes On-Line Database [17] allowing more directed analysis on an individual gene scale as well as on a genome scale. The majority of these more global functional studies have been conducted on the model cyanobacterium Synechocystis sp. PCC6803 (henceforth referred to as Synechocystis). This unicellular and freshwater cyanobacterium has a fully sequenced and annotated genome [18]. It is also able to grow photoheterotrophically and is naturally transformable, all characteristics which enhance its suitability for combined systems level genetic, proteomic and metabolomic (metabolic flux) studies [18–20].

Salt studies in cyanobacteria have generated a large amount of information and are generally centred on four main thematic areas-

(1) Biochemical and physiology based studies [21–30].

(2) Salt intake and cell signalling [31, 32].

(3) Gene level responses- salt regulated genes [33, 34], microarrays [35, 36] and mutational analysis of salt tolerance determinants [37].

(4) Post-genomics [38–41].

This work has increased our understanding of the strategies cyanobacteria implement to adapt to high salinity, and a schematic overview is given in a review by Joset et al. [5]. The immediate, or shock, responses include the accumulation of compatible solutes and active export of inorganic ions, with these being regarded as largely protein synthesis independent [5]. However, in order to survive in high salt for extended periods of time, cells must adapt to the new conditions. Adaptation is a long-term process, and is therefore protein synthesis dependant. It is therefore imperative to study how proteins play functional roles in salt adaptation strategies, and this review will emphasise work specifically carried out in the proteomics field.

The proteome of an organism refers to the total set of proteins encoded by its genome (protein coding genes) [42], and therefore the field of proteomics encompasses studying these proteins, specifically the change in abundance (and status of post-translational modifications (PTM's)) in cells, tissues and organelles in response to changing environmental factors. Unlike previous protein based studies, proteomics involves the global examination of proteins. This field has the capability to reveal information on the level of protein expression, protein isoforms produced from each gene, the extent to which proteins are post-translationally modified and also the cellular and sub-cellular distribution of proteins.

In the 1990's, during the advent of genetic sequencing and bioinformatics, the ability to identify proteins increased through the use of mass spectrometry (MS). Now the technique is coupled to cross examination of databases, and has become an intrinsic part of proteomics research. With higher sensitivity and throughput, the number of proteins identified per experiment has increased. Proteins can be identified by peptide mass fingerprinting (PMF) or tandem MS, and together with the technical development of analysis software, the identification of proteins has become rapid and more accurate [43]. This provides a valuable tool for investigating protein profile changes in response to environmental stimuli, such as high salt. Furthermore, it has been demonstrated that expression patterns between mRNA and proteins levels may not correlate, and that protein levels are dependant on translational controls and regulated degradation, as well as the abundance of corresponding messages (mRNA) [44, 45]. These studies indicate gaps in our knowledge of how transcription, translation and post-translational modifications interact and ultimately effect protein abundance and function. By mapping which proteins interact together and which proteins play pivotal roles in certain conditions, a further understanding of survival traits, including modifications of specific cellular pathways, can be achieved. However, the task is not simple, and studies using nuclear magnetic resonance imaging have shown that proteins are more active and dynamic than initially predicted, adding complexity to understanding where each protein is located in a cell, when the protein is present and for how long, and with which other proteins it is interacting [46].

In this review, a general overview of current proteomic methodologies will be discussed, emphasising the aspect of using and analysing samples with potentially detrimental amounts of salt. Use of these technologies with the well studied model system Synechocystis, with associated findings will also be considered, highlighting current challenges. With the rapid advances in proteomic techniques, some of these goals will be addressed, with reference to more challenging systems.

2.0 Current proteomics methodologies

2.1 General overview

A proteomics experiment involves collaboration of several techniques, commencing with protein extraction and ending with accurate and reliable identification and quantitation. A typical proteomic workflow is presented in Figure 1.

Figure 1
figure 1

A general proteomic workflow. The main steps are, cell culture, protein extraction (liquid nitrogen freezing and cracking), sample fractionation (traditional 2DE and gel-free chromatography method), mass spectrometer analysis and protein quantitation and identification. Quantitation using 2DE is densitometry based and implemented with appropriate software (e.g. PD QUEST®, SameSpots®), but can also be undertaken on the mass spectrometer, relying on abundance of peptide peaks. Identification occurs with database search software (based on peptide mass fingerprint or sequence analysis). Proteins are commonly digested using trypsin. : indicate essential salt removal stages. RCDC® – protein quantitation assay kit, IEF- isoelectric focusing, HPLC- High Performance Liquid Chromatography, SCX- Strong Cation Exchange.

Cyanobacterial proteomes (in common with most proteomes) are very complex, and in some cases consist of several thousand proteins [18, 47]. Due to this complexity, two-dimensional polyacrylamide gel electrophoresis (2DE) has been widely utilised as the standard protein separation technique. Highly sensitive 2DE was developed in 1975, and the technique enabled fractionation of a complex mixture of proteins based on their isoelectric point (pI) and molecular weight [48, 49]. The method has been shown to have very impressive resolving power [50], and allows for a quantitative comparison of protein changes in multiple samples by using a variety of staining techniques. Another major advantage of 2DE is the ability to detect protein isoforms, for example, PTM's such as phosphorylation [51, 52]. However, 2DE has been criticised for a lack of automation and is often considered to be a laborious process. In addition, the success of the method is dependant on optimisation [53]. The ability to improve results by varying conditions and techniques, allows at least a satisfactory global overview of a proteome.

Recent advances in 2DE include the introduction of narrow range immobilised pH gradient (IPG) strips, which have significantly improved protein coverage [54]. IPG strips are prepared by co-polymerising acrylamide monomers with acrylamide derivatives which contain carboxylic and tertiary amino groups. The buffering groups form a fixed pH in the strip [55]. Automated spot excision systems and mass in-gel trypsin digestion techniques have also reduced the arduous nature of preparing samples from 2DE gels for MS analysis [56, 57], and software packages such as Samespots and Progenesis (Nonlinear Dynamics, Newcastle, U.K.) have enhanced gel analysis capabilities. These and additional methods are extensively reviewed elsewhere [58–61]. In addition to advances in protein identification and throughput efficiency of 2DE, quantitation methodology has also improved. Fluorescent and radioactive stains can be used prior to protein separation, as well as highly sensitive post-separation stains such as SYPRO ruby, coomassie and silver stains [62–64].

Protein mixtures can now be multiplexed and run on the same gel in a technique called differential in gel electrophoresis (DIGE) [65]. Through the use of multiple fluorescent dyes to label protein samples prior to 2DE, DIGE allows for multiple samples to be co-separated and visualized on one single gel. By using CyDyes (GE Healthcare, Buckinghamshire, U.K.) not only is the sensitivity and reproducibility enhanced, but also an internal standard is included increasing statistical confidence. Unlu et al. [65] initially reviewed this method comparing two Drosophila embryo extracts. DIGE was found to be reproducible and sensitive, detecting differential abundance of proteins at nanogram levels.

Quantitation can also be improved by using in vivo isotope labelling of proteins prior to 2DE separation. This method is particularly attractive because the isotope label is incorporated into the organisms' proteome, with no physiochemical or biological consequences. Potentially all proteins are labelled, and this technique is very applicable and compatible with accompanying techniques. The different peptide isotopes can be resolved by MS, and this is where quantitation occurs (based on peptide peak areas), in contrast to staining intensity analysis, which is only semi-quantitative, and requires many replicate gels for statistical relevance. Methods include stable isotope labelling with amino acids in cell culture (SILAC) [66] and elemental labelling [67]. In the SILAC method, cell lines are cultured in media lacking an essential amino acid, supplemented with an isotopically labelled form of that amino acid, for example, deuterated leucine. This process allows quantitation of potentially all leucine containing peptides [66]. For a more universal labelling method where all peptides can be labelled, elemental labelling with media containing 15N or 13C isotopes can be performed [67]. As demonstrated by Snijders et al. [68], combining metabolic labelling with gel electrophoresis means a further step is added to the traditional method. However, this produces a faster workflow to identify differentially expressed proteins, as the staining intensity is used as an initial pre-screen for differentially stained protein spots, and only these are selected for identification and accurate MS-based quantitation [68]. Two gels are run in parallel with each containing proteins sourced from two phenotypes, with one isotopically labelled reference phenotype (the reference phenotype is the same in both gels). A comparison of staining intensity can then highlight potentially differentially expressed proteins. In contrast to a traditional 2DE workflow, each spot potentially represents proteins from two phenotypes. Corresponding protein spots from both gels are excised and trypsin digested, and upon resolving peptides by MS, relative ion intensity from both phenotypes can be used for accurate quantitation. Another advantage of this method is that two phenotypes can be combined in one gel (labelled and un-labelled) reducing technical variation [68].

In recent years, developments have been made to replace gel-based fractionation to overcome their limitations. Complex protein mixtures can be separated based on physical and chemical properties, including size and hydrophobicity [69, 70]. It is also possible to digest proteins into peptides and separate on peptide properties [71–73]. Quantitation using non-gel-based (shotgun) methods has progressed rapidly, and usually involves comparison against an included standard. Quantitation can either be relative or absolute, and example methods are discussed further in section 4.

The common link in all the techniques discussed above is the mass spectrometer. It plays an important role in protein identification and can also be implemented for quantitation. Principally the mass spectrometer measures the mass-to-charge ratio of analytes (peptides). Analytes measured in the gas phase are ionised most popularly by electrospray ionisation (ESI) or matrix-assisted laser-desorption ionization time of flight mass spectrometry (MALDI-TOF MS) [74, 75]. MALDI-MS is usually coupled to time-of-flight (TOF) analysers and therefore intact peptide masses are measured, and these are subsequently matched to theoretical masses in a database to provide protein identification (known as PMF). ESI is mostly coupled to ion traps, triple quadrupole or quadrupole-TOF instruments, which fragment selected precursor ions (peptides) further and generate ion spectra. Utilising an array of different algorithms, this information is matched against protein sequence databases to provide protein identifications and information on peptide sequence is also provided. Aebersold et al. [76] presents a general insight into mass-spectrometry-based proteomics. Dealing with the extensive amount of data generated from these global scale studies is a challenge, and the role of bioinformatics in proteomics is fundamental. A review of bioinformatics in proteomics with particular reference to the development of novel algorithms and discussion of the integration of bioinformatic resources can be found in Blueggel et al. [77] and Huang et al. [78].

2.2 Working with 'Salty' samples

Salt poses a serious problem for resolving proteins by 2DE, but particularly in the first dimension where high voltage is required to focus proteins by isoelectric point. Removing salt is therefore imperative to allow sufficient focusing, minimise sample loss and avoid poor resolution due to streaking on gels. Salt removal can be performed during several stages of the general proteomic work-flow as summarised in Figure 1.

2.2.1 Salt removal before protein extraction

Cell harvesting requires centrifugation where the supernatant (media) is removed, leaving a cell pellet. To remove external salts, several further wash steps are easily performed by resuspending the pellet in a non-saline solution, for example growth media with no added salts, and repeating the centrifugation. In halophiles, cells lyse when the environmental salt concentration drops below ca. 10–15%. Therefore salt cannot be removed with a wash step here and instead suspension of cells in a low-salt solution can be used directly for protein extraction by inducing osmotic shock. However, after protein recovery from the supernatant by high speed centrifugation (250,000 × g), a wash step can be incorporated to remove excess salt.

2.2.2 Salt removal after protein extraction

At the Korea University, Seoul, the use of 3 kDa molecular weight (Mr) cut-off columns was successfully utilised to remove excess salts from proteins [79]. These filtration columns contain pores in which only small molecules, for example, water and salt can enter, excluding proteins. The nature in which the desalted protein sample is collected (i.e. to avoid any protein loss) means that extra elution buffer is present but this dilution of the protein sample is relatively small. Another method that has gained popularity is dialysis using microdialysis tubes and centrifugation. The result is a marked improvement in protein resolving power but the time added to the overall method can be two and twelve hours, a definite disadvantage as some proteases may still be active during this period [53]. This process requires a lot of solution, whereas small volumes of protein solution are more commonly used in proteomic methodologies. Another disadvantage is that some proteins can precipitate after dialysis [80].

Ultrafiltration is not a widely used technique, but has been successfully implemented to remove salt [81]. The ultrafiltration device is fitted with a filter membrane which usually allows molecules smaller than 5 kDa to pass through, but retain larger molecules in the cell such as proteins, when high pressure is applied. This method has the benefit of not increasing the sample volume.

Precipitation methods have been included in the proteomics process for several reasons, mostly to concentrate highly dilute samples, inhibit protease activity and again to remove disruptive material such as salt. Recently, Smith et al. [80] compared several precipitation methods using trichloroacetic acid (TCA), acetone, chloroform/methanol, ammonium sulphate as well as ultrafiltration. The most promising results were observed using TCA and ultrafiltration due to the advantageous desalting effects. The TCA process is very easily applied, consisting of two main steps, incubation at -20°C with TCA/acetone, and removal by washing with TCA and centrifugation [82]. However, it was reported that TCA precipitation results in integral membrane proteins loss, highlighting the need for alternative methods when resolving membrane proteins by electrophoresis [83]. Finally, precipitation with ammonium sulphate was shown to be useful when attempting to eliminate high abundance proteins [80].

2.2.3 Salt removal during fractionation

It is possible to adjust parameters or commonly used methods so as to eliminate interfering salt. Increasingly popular is the incorporation of a desalting step in the first dimension of a 2DE experiment, isoelectric focusing (IEF) [84]. By including an initial low-voltage step, the highly charged salt in the sample can migrate to the electrode, allowing the lesser charged proteins to migrate without interference. However, this may not be solely sufficient when working with hypersaline conditions, for example >6% w/v NaCl, and can add up to nine hours to the first dimension step. Cup-loading during IEF allows a concentration maximum of 50 mM salt, higher than passive rehydration of IPG strips [85]. In cup-loading, the protein sample is applied onto a small area of the strip during the focusing program, using a plastic cup, whereas in passive rehydration the sample and rehydration buffer are added directly to the entire strip and allowed to absorb for up to 12 hours. For passively rehydrated IPG strips, it has been shown that a simple regime of three, ten minute washes prior to focusing improves the quality of second dimension separation [86].

2.2.4 Online salt removal with instrumentation

More sophisticated methods of desalting have also been developed to remove small quantities of salt prior to MS. Particularly useful for high-throughput laboratories dealing with a large number of salt samples, is the coupling of a desalting chip with ESI-MS. Proteins are adsorbed on a hydrophobic poly(vinylidene difluoride) membrane, which allows the washing out of salts [87]. Furthermore, microspin columns with modified C18 membranes have been shown to be particularly rapid at salt removal [88].

Desorption electrospray ionisation (DESI) with MS analysis has been shown to tolerate higher levels of salt. Salt is particularly problematic when using ESI, and leads to ion suppression [89]. Non-volatile materials like salts alter the efficiency of droplet formation, which in turn negatively affects the number of ions reaching the mass spectrometer. Adding ammonium acetate to the buffer in ESI is known to reduce ion suppression effects in ESI, but the ability to optimise the spray solvent and use a variety of different surfaces for initial sample analysis using DESI, allows successful analysis of samples with 0.5 to 2% salt [90].

2.2.5 A simple salt removal strategy

Figure 2A schematically presents a salt removal strategy when using 2DE or one-dimensional (1DE) sodium doecyl sulphate polyacrylamide gel electrophoresis (SDS-PAGE) techniques, based on the current developments discussed above. Specifically, this method is easy to implement and relatively inexpensive. When used with halotolerant organisms, the washing buffer step is performed with a non-saline buffer at the cell harvesting stage, whereas with halophilic organisms it is applied at the protein extraction stage. To demonstrate the effectiveness of this procedure, Figure 2B shows a selection of small format (7 cm2) 2DE gels which were run for cyanobacterial protein extracts taken from cultures grown in 6% NaCl. Four gels were run with the elimination of different salt removal steps as shown in Figure 2A, and in the final gel the full procedure was followed.

Figure 2
figure 2

(A) A simple salt removal strategy for proteomic experiments using gel electrophoresis for protein fractionation. Desalting steps are in green boxes, example techniques in circular boxes and parameters in diamond boxes. N.B. For halophilic organisms the wash step with no salt buffer must be performed at the protein extraction stage. (B) Euhalothece desalting optimisation tests. 7 cm2 gels, pI 3–10. i) No desalting steps ii) TCA/acetone precipitation only iii) Desalting column not used iv) External salt not washed v) Complete desalting procedure followed.

3.0 Proteomic salt responses in a model system

The majority of proteomic studies on salt tolerance in cyanobacteria have been conducted on Synechocystis, using 2DE. Newly expressed or proteins with enhanced expression in salt stressed cells have been identified, and a group of these are termed 'stress proteins'. These studies have contributed to the construction of the Cyanobase database, containing over three thousand predicted protein coding genes [91]. Many of these proteins are annotated as 'hypothetical', as sequence similarity to proteins of known function is absent. Methods for isolating plasma, thylakoid, periplasmic and outer membrane proteins have been developed, as well as isolating pure plasma and periplasmic space proteins [41, 92–96]. This has allowed analysis of how proteins in different subcellular compartments assist in cellular adaptation to high salt.

3.1 Salt response in different cellular compartments

Figure 3 summarises the proteomics-based responses to salt stress in a unicellular cyanobacterial cell, and Table 1 defines the corresponding conditions under which cells were investigated. Cells were grown in BG11 media [97] supplemented with NaCl unless otherwise stated.

Figure 3
figure 3

Salt responses at the protein level. A summary of responses to varying degrees of salt stress (~2–4% w/v NaCl) in the main cellular compartments of a unicellular cyanobacterium cell, as revealed by proteomics studies [[38, 39, 41, 109]].

Table 1 A summary of proteomic studies in cyanobacteria.
3.1.1 Periplasmic Space

The periplasmic space (Figure 3) is situated between the plasma membrane and the outer membrane. No active ion transport mechanisms have been found in the outer membrane in Synechocystis cells, but transport proteins involved in the uptake of substrates are present in the periplasmic space, and these could be affected by high salt levels. This inspired Fulda et al. [41] to use a cold-osmotic shock technique [98] to isolate and compare the abundance of periplasmic proteins from Synechocystis grown in low salt (2 mM or 0.01% NaCl) and high salt (342 mM or 2% NaCl). The method of periplasmic protein isolation is unsuitable for examining adaptation to higher salinities, as the sample fraction becomes contaminated with proteins from other cellular compartments in these conditions. Extracted protein fractions were separated by SDS-PAGE, and two bands of interest were subjected to N-terminal sequencing. Unfortunately, both identified proteins were annotated as hypothetical [41].

By implementing 2DE for enhanced resolution with MALDI-TOF MS, the experiment was repeated [38]. pH 4–7 IPG strips were used with 23 × 23 cm 12.5% acrylamide gels. MALDI-TOF MS allows protein identification by soft ionisation methods with a TOF mass spectrometer [75]. Protein fragments generated by protease digestion are fixed in a solid phase matrix and ionised by laser. TOF analysis reveals experimental peptide masses which can then be matched to theoretical masses in a database. Six and three proteins were identified as enhanced and newly expressed, respectively, in the presence of elevated levels of salt [38]. Several of these proteins were thought to be involved in synthesising or modifying extracellular layers in Synechocystis, highlighting the importance of changes in cyanobacterial cell walls in adapting to high salinity. For example, a protein showed significant similarity to phosphoglycerate mutases from Streptococci, which are involved in polysaccharide synthesis in their capsules [38]. Furthermore, a salt enhanced protein was found to be homologous to cell surface lipoproteins of Mycobacteria spp. [38].

It is apparent from this study, that functional annotation of Synechocystis is still at an early stage, and characterisation of these salt enhanced proteins required searching for similar functional domains present in the databases for other organisms. A number of periplasmic proteins were also observed to have reduced expression in saline media, and these were attributed to the expected reduction in protein synthesis of non-stress proteins in salt stressed cells [38, 99].

3.1.2 Membrane Proteins

A combination of blue native/SDS-PAGE and MALDI-TOF was used to study the composition and dynamics of membrane protein complexes in Synechocystis, perturbed by three different growth modes and varying levels of CO2, iron, or salt [96]. Blue native is a proteomic technique implemented in order to separate and study intact protein complexes at high resolution [100, 101]. It enables further elucidation of protein function by analysing protein-protein interactions. A total of 53 protein spots were identified in this study, corresponding to 37 different Synechocystis genes [96]. A stress of 0.4 M (~2.4%) NaCl was applied to cells for 72 hours, but this did not induce any observable changes in the membrane protein complexes, which agreed with previous physiological studies (reviewed by Joset et al. [5]) that the plasma membrane and cytoplasm are likely to be the main subcellular compartments affected by salinity.

Huang et al. [102] screened for proteomic changes in the plasma membranes of Synechocystis in response to salt. The plasma membrane represents a barrier to the surrounding medium, and is likely to be responsive to environmental salinity changes [38]. These membranes were isolated by sucrose density gradient centrifugation and two-phase partitioning [92]. 106 proteins were identified, corresponding to 66 gene products, using 2DE and MALDI-TOF MS. 25 proteins changed significantly in abundance due to salt concentrations of 684 mM (~4.1%). A methanol/chloroform precipitation step was included after protein extraction to desalt and concentrate the sample [103]. An advantage of using 2DE for protein separation and visualisation is particularly apparent in this study, as isoform spots were clearly visible. Abundance changes in some of the isoforms imply PTM's may play a role in protein function in the stress conditions. Nearly one third of the salt-enhanced proteins identified in this study are substrate binding ABC transporters [102]. An increase in glucosylglycerol-binding protein, GgtB, was expected since Synechocystis cells can uptake this compatible solute from its environment. The evidence for an interrelation between salt and iron-stress in this organism was supported by the very high accumulation of iron binding lipoprotein, FutA1. It has been previously proposed that FutA1 plays a protective role in photosystem II under conditions of iron deficiency [104]. Increases in phosphate and nitrite/nitrate binding proteins have been hypothesised as a necessity for cells to overcome salt induced nutrient deficiency, a problem resulting from plasma membrane structural changes [105]. Moreover, an increase in the regulatory protein PII suggests a change in the carbon and nitrogen balance in salt-stressed cells.

Further salt-induced proteins, thought to play significant roles in stress, included vesicle-inducing protein in plastids (Vipp1), a protein involved in thylakoid membrane biogenesis in Arabidopsis [106], membrane-bound peptidyl-prolyl isomerase B, which could be involved in maintaining the integrity of proteins in the plasma membrane and CoxB, which supports recent work suggesting a role in managing photosynthesis in stressed cells [39]. Some of the hypothesised functions in salt tolerance require further characterisation, at the protein level through, for example, structural studies, or at the gene level via knock-out mutations from the protein-coding genes and assessment of phenotypic alterations. This is particularly true for hypothetical proteins which have no associated function. In addition to salt regulated proteins, 21 proteins were newly identified [102] and before further studies are undertaken it is possible to predict function to a certain limit. A plethora of open source software tools have been developed, with algorithms which help to predict cellular functions from conserved domains. Furthermore, characteristics such as hydrophobicity and attributes like lipoprotein or sec- signals, can be used to predict subcellular localisation [107, 108].

Many proteins from the plasma membrane were associated with salt stress in Synechocystis for the first time in this study, however, the limitations of using 2DE were recognized by virtue of the fact that no integral membrane proteins were identified [102].

3.1.3 Cytoplasm

Fulda et al. [39] used 2DE with MALDI-TOF to investigate the soluble proteome response in Synechocystis cells to approximately 4% salt, and 55 out of the 337 identified proteins were induced. In the salt-acclimated cells, induced proteins were organised into four groups; stress proteins which are salt specific, general stress proteins, enzymes involved in basic carbon metabolism and hypothetical proteins.

Specific salt stress proteins included the enzyme ADP-glucose pyrophosphorylase, responsible for synthesising the precursor ADP-glucose for the compatible solute glucosylglycerol. General stress proteins included molecular chaperones GroEL1 and elongation factor-Tu. Increased amounts of soluble electron carriers, flavodoxin and plastocyanin were also observed, and these are thought to play an important role in adjusting to stress-induced changes in electron transfer [39]. Several enzymes involved in carbohydrate metabolism also accumulated, including transketolase, glycogen phosphorylase and phosphoglycerate kinase [39]. The accumulation of the glucosylglycerol is proposed to cause this change in carbon metabolism.

3.1.4 Thylakoids

A method for isolating pure thylakoid membrane proteins suitable for 2DE analysis was developed nearly a decade ago, implementing aqueous polymer two-phase partitioning in combination with sucrose density centrifugation [92]. Despite this, a large scale proteomics assessment of the salt response of the photosynthetic membranes has not been performed. By combining traditional protein separation and quantitation methods SDS-PAGE and western blotting, Sudhir et al. [109] investigated the response of thylakoid membrane proteins in the cyanobacterium Spirulina platensis to 0.8 M (~4.8%) NaCl. S. platensis was isolated from sodium (150–200 mM Na+) rich lakes and ponds and is cultured in Zarrouk medium [109]. A 40% decrease in protein D1 from the PSII reaction centre was observed. Additional abundance changes in proteins were observed but these were not identified. The authors undertook additional physiological studies and postulated that a variety of effects occurred on photosynthetic electron transport activities due to the marked alterations in the composition of thylakoid membrane proteins [109].

Vipp1 protein, (see section 3.1.2), is located in the plasma membranes, but has been associated with thylakoid membrane formation in Arabidopsis [106]. It shares sequence similarity to stress related phage shock protein PspA [106], and a knockout mutant created in Synechocystis led to an inability to form thylakoid membranes. An increase in this protein in salt acclimated cells may imply a structural reorganisation in these membranes is necessary [102].

3.2 Confirmation methods

In these studies, a confirmation technique is often used in parallel with newly developed proteomics workflows with the aim of confirming their reliability and applicability. Complementary techniques for validation of protein abundance changes have recently been incorporated into many proteomic studies, with western blotting being the most widely applied [110]. This practice is normally encouraged for several proteins. For example, immunoblotting analysis confirmed the expression changes of several proteins in the study by Huang et al. [102].

3.3 Shock versus adaptation

Biological function is predominantly facilitated by proteins, highlighting the power of proteomics to identify the physiological changes which take place in cells in order to survive adverse conditions. However, transcriptomics is useful for assessing immediate gene expression changes, and therefore integrating both sets of data is necessary for a more complete systems level understanding. Nevertheless, proteomics can yield insight into immediate and long-term physiological responses. Fulda et al. [39] used 2DE in a proteomics study to address the salt shock (immediate) response versus salt adaptation (long-term) response in Synechocystis to approximately 4% NaCl. To produce a meaningful comparison, previously generated transcriptomic data and a survey of changes in physiological parameters were used to identify the most revealing times to harvest cells. Short-term shock and long-term adaptation times were at 2 hours and 5 days, respectively [36, 39]. A pulse 35S-methionine labelling method was used, allowing clear visualisation of newly induced proteins from radioactive 2DE-gels. In vitro metabolic labelling of cells with radioactively labelled amino acids is routinely performed, but despite advantages such as relatively high energy beta emitter potential and high specific activity, any proteins lacking methionine would be undetected [111, 112]. Despite the fewer induced proteins identified compared to genes induced in a microarray study, 90% of these identified proteins were also induced at the transcriptional level [36].

As expected, heat shock proteins (which protect or repair proteins), DNA and RNA-binding stress proteins and antioxidative enzymes all increased in expression in the short and long-term, in both transcriptomic and proteomic studies [35, 36, 39, 102]. However, only truncated versions of the chaperone GroEL1 and the elongation factor-Tu, were identified in the long term by proteomics, and whether this has functional significance remains unknown [39]. Seven enzymes involved in carbohydrate metabolism also accumulated after 5 days, whereas a similar pattern in gene expression was not observed. In fact, in this study, nearly half of the data for induced genes in acclimated cells (5 days) did not correlate with protein inductions [39]. Therefore, the authors proposed that PTM's may play a role in salt adaptation, particularly considering the multiple spots identified in gels corresponding to the same protein.

4.0 Advances in proteomic methodologies

The studies discussed in section 3 lay the foundations for understanding global cellular salt responses. However, 2DE is the technique invariably used for protein separation and quantification. Limitations of 2DE are apparent when comparing proteomic data with transcriptomic data, most significantly, coverage of the full proteome is poor. Unfortunately, low abundance proteins, proteins with extreme pIs, membrane and membrane bound proteins (due to their hydrophobicity), cannot readily be resolved by 2DE [113]. For example, identifying and studying proteins with a pI of 4 to 7, is only part of the salt adaptation story. A protein pH range of 4 to 7 is popular for studying Synechocystis, as its predicted proteome is biased towards an acidic pH [17]. Furthermore, the presence of high abundance pigment proteins, for example, phycocyanin proteins in cyanobacteria, can cause smearing in gels and reduce the quality of resolution and quantitation [39]. In addition to proteome coverage issues, protein abundance changes based on staining intensity are only semi-quantitative. Consequently, this has led to the development of gel-free proteomics techniques, sometimes referred to as shotgun proteomics.

As mentioned previously, the rapid development in shotgun proteomics has led to an array of techniques to fractionate complex protein samples. For example, proteins can be separated by characteristics such as size, hydrophobicity or charge using size exclusion chromatography (SEC) [69], reverse-phase (RP) chromatography [70] and weak/strong anion exchange chromatography (WAX/SAX) [114, 115], respectively. Multi-dimensional separation schemes have also been applied at the peptide level, where proteins are digested with an appropriate enzyme, for example trypsin, and then separated by methods including strong cation exchange (SCX) [71], capillary isoelectric focusing [72] and liquid phase isoelectric focusing [73].

By combining an array of peptide and protein fractionation techniques prior to MS analysis, the proteome coverage can be increased significantly [116, 117]. Gan et al. [118] compared a combination of six fractionation workflows employing techniques which separate proteins via IEF, 1-D PAGE, WAX chromatography, as well as peptide separation using IEF or SCX, all prior to reverse phase multi-dimensional liquid chromatography and MS analysis [118]. 776 proteins were identified in Synechocystis, representing over 20% of the total predicted proteome. In conjunction with technical advances in this area, the need to increase the representation of an organism's proteome continues. One challenge is to increase the solubility of membrane proteins whilst maintaining compatibility with downstream processes [119, 120]. This can be achieved through the use of high percentage organic solvents and acids. Moreover, the use of strong zwitterionic and non-ionic detergents have been shown to improve the solubilisation of membrane proteins from cyanobacteria [121]. In a study conducted on the filamentous cyanobacterium, Anabaena variabilis, such techniques led to the identification of 646 (13%) proteins [121].

With the ability to identify hundreds of proteins in a single proteomics experiment, the need to quantify differential protein abundance efficiently and accurately, has been realised. It is now possible to perform a large scale proteome study by identifying and quantifying proteins from complex samples. Quantitation using shotgun methods has progressed at an impressive rate utilising isotope labelling strategies such as SILAC. Isotope tags can also be added enzymatically by transferring 18O to peptides from water [122]. In addition, chemical reactions are used where reagents are tagged to the protein or peptides, and these provide internal standards for quantitation. Examples include isotope coded affinity tags (ICAT) [123], mass coded abundance tagging (MCAT) [124], the AQUA strategy (for absolute quantitation) [125], and isobaric tags for relative and absolute quantification (iTRAQ) [126].

The iTRAQ procedure has recently gained in popularity [127–130]. It is usually based on a peptide labelling strategy, where four or eight amine reactive isobaric reagents are used, enabling the comparison of several phenotypes in one experiment (including technical and biological replicates). The quantification occurs at the MS/MS stage, and thus requires a tandem mass spectrometer [131]. MS/MS leads to fragmentation of peptides into individual amino acids allowing a far more accurate sample analyses.

iTRAQ has been demonstrated as a useful, robust and reliable tool for proteomics [131], with several advantages over densitometry and other labelling techniques including quantitation accuracy and overall workflow efficiency. A typical iTRAQ experiment takes no more than several hours over approximately 2 to 3 days to label and pre-process the samples, 2 to 3 days (based on 20 fractions × 135 minutes runs) to separate and analyse by MS, and 1 day to initially process the MS data (Figure 4). Moreover, these runs would include biological and technical replicates [129]. This time-scale and work load is considerably less than a 2DE work-flow which requires running 4 separate gels (in triplicate). Each gel requires 1st dimension separation (day 1), 2nd dimension separation (day 2), staining (overnight), spot matching and identifying spots of interest, cutting all spots (approximately 500–1000) and tryptically digesting (at least several days), peptide extractions (several days), and finally MS analysis (several days).

Figure 4
figure 4

An overview of a proteomic procedure with iTRAQ reagents. Proteins are extracted from four phenotypes which are to be compared and trypsin digested. The resulting peptides are labelled with isobaric tags and then pooled for fractionation and analysis at the MS/MS stage on the mass spectrometer. Isobaric tag intensity corresponds to peptide abundance and is used for protein quantitation.

Choe et al. [132] demonstrated the consistency of observations using 2DE and iTRAQ, as well as the reproducibility of the experiments. They reported that half of the quantified protein expression ratios have a coefficient of variation (CV) less than 0.31 using 2DE and less than 0.24 using isobaric tags; whereas 95% of the quantified protein expression ratios have a CV less than 0.81 using 2DE and less than 0.53 using isobaric tags [132]. Wu et al. [133] compared iTRAQ to DIGE and ICAT, and reported its superior quantitative sensitivity. A survey of technical, experimental and biological variations using the technique was also undertaken using Synechocystis as a model organism [129]. Furthermore, Chong et al. [128] used iTRAQ with Synechocystis, demonstrating the use of a multiple injections strategy to improve proteome coverage. An average of 218 proteins was identified and the issue of abundant pigment proteins was highlighted. There have been no studies to date, looking at salt stress or adaptation in cyanobacteria, using shotgun proteomic techniques.

5.0 Challenging systems

Synechocystis is a model organism for photosynthesis research and has therefore been investigated extensively. It is also the model organism for cyanobacterial and higher plant salt stress studies, and this is the case in proteomics. Its genome is fully sequenced which makes identification of proteins through MS and database searching, relatively simple. The expectations are that discoveries made in salt stress proteomics studies in Synechocystis, will provide insight into the adaptive mechanisms in other organisms. Uncovering the proteomic response of this model organism has made an invaluable contribution to understanding salt acclimation. However, cyanobacteria are ancient life forms and evolutionary pressures have created many diverse species, including those which can survive in higher salt concentrations. Unfortunately, the majority of these cyanobacteria do not have their genomes sequenced, so how is it possible to identify proteins and hence study the proteome of an unsequenced organism? This is an area known as cross-species proteomics, and is yet to be fully exploited [134, 135].

5.1 Cross species proteomics

Accompanying the rapid growth in the number of fully sequenced genomes, is the increasing number of protein sequences present in databases. This maximises the chance of identifying proteins from a non-sequenced organism using conventional protein identification software. Conventional software, for example Mascot (Matrix Science, London, UK) relies on matching of tryptic peptides to theoretical peptides present in databases. With increasing numbers of sequenced proteins, the chance of an unsequenced organism producing an exact same tryptic peptide is correspondingly enhanced. Developments in the area of MS and bioinformatics have also significantly expanded the applicability of cross-species proteomics [134–138]. Mass spectrometry-driven BLAST (MS BLAST) is a database search protocol which uses a list of peptide sequences generated by interpretation of MS/MS spectra using ProBLAST software [135, 139]. Significant hits in MS BLAST are colour-coded based on a high scoring pairs (HSP's) algorithm. More recently, open source software SPIDER uses an alternative algorithm for identification of novel proteins [140]. Collectively, they rely on the interpretation of MS/MS spectra using de novo sequencing, a technique which has also advanced with new powerful software, for example PEAKS and OLAV [141, 142].

The main challenge in cross-species proteomics is to produce high confidence protein identifications, and this requires meeting set criteria. Using probability calculations and cut-off scores in specific analysis software and fulfilling the guidelines generated for identification (I.D.) methods in proteomics journals [143], for example, requiring 2 or more peptides per protein, should be included. One approach is to design degenerate nucleotide primers for the protein coding gene by using peptide sequences detected in the MS. Alternatively, alignment files generated from nucleotide sequences from related organisms, which are present in nucleotide databases, can be used. Successful sequencing of the gene provides verification of the protein identification, and therefore extra confidence in the identification methods implemented. A method using metabolic labelling combined with gel electrophoresis [68, 144] is particularly advantageous. This method was initially designed to provide a fast and accurate method for protein quantification, however, it presents the opportunity to verify protein I.D.'s for non-sequenced organisms. In this method, proteins are metabolically labelled by culturing in the presence of a heavy isotope, for example, 13C or 15N. Cyanobacteria are often cultured in BG11 media [97], and the main source of nitrogen in this media is sodium nitrate, and therefore replacing this with an isotopic Na15NO3 is relatively simple, as long as a sufficient number of generation times are included for maximum incorporation efficiency (and it is confirmed that the organism cannot fix atmospheric nitrogen). These light (14N) and heavy (15N) peptides can be distinguished by mass-to-charge scale on the mass spectrometer, and the ratio in peak height or area between the two forms correlates with protein abundance. However, the difference in mass between labelled and unlabelled versions of the same peptide provides extra identification evidence, via imposition of an elemental constraint [145].

Protein fractionation is important to reduce sample complexity, and thereby allowing more successful interpretation of meaningful MS spectra, due to issues involved with ion suppression and duty cycle [89, 146]. 2DE can be used for protein separation which allows a further identification confirmation by matching gel positioning with predicted isoelectric point and molecular weight. Using highly saline samples with 2DE would require a salt removal workflow as illustrated in Figure 2. Similarly, SDS-PAGE allows confirmation by molecular weight, but has the added advantage of improved resolution of hydrophobic proteins [144]. A work-flow has been developed very recently to increase protein identifications in non-sequenced organisms using complex samples with high-throughput methods such as iTRAQ, in which case protein fractionation via gel electrophoresis is eliminated [147].

It has been demonstrated by comparative genomics [148] that the difference in salt tolerance between glycophytic and halophytic plants is quantitative, and not necessarily qualitative. Therefore, tolerance mechanisms may be conserved amongst all plant species. This highlights the importance in studying the salt response in alternative cyanobacteria despite the likelihood that 'new' proteins will not be discovered in non-sequenced isolates and there is more chance that a comparison to existing systems will be performed.

5.2 Proteomics of an unsequenced extreme halotolerant cyanobacterium

We have isolated a eurohaline cyanobacterium from a salt lake in southern Libya, in the heart of the Sahara. The salinity of the lake varies by sampling location and season, but can reach up to 16% w/v [149]. This microorganism can adapt to large variances in salt concentration, and unlike halophiles, does not require salt for survival. It therefore represents an ideal system to study cellular responses to adaptation in both high and low salt environments. This ability of adapting to a large range of salinities may actually be more difficult to achieve than the fixed salt requiring metabolism of halophiles.

Furthermore, it is an environmental isolate and not sequenced, presenting a challenge for protein identification. The method combining gel electrophoresis and metabolic labelling discussed above was used to study adaptation of this organism to the addition of 0, 3, 6 and 9% salt (NaCl) to BG-11 [144]. Proteins were identified using Mascot (relying on generating identical tryptic peptides to those present in the database) and MS BLAST, where extra proteins could be identified via de novo sequencing and homology searching. In total, 383 unique proteins were identified and 23 different organisms were required as identification sources (17 different cyanobacteria), allowing an in-depth analysis of its adaptive response to high and low salinity [144]. This particular cyanobacterium shares many similar salt-tolerance mechanisms to Synechocystis, albeit in higher salt concentrations. Euhalothece increases production of stress chaperones and antioxidative enzyme superoxide dismutase. There appear to be alterations to the cell wall, decreased pigment production as well as significant alterations in central intermediary metabolism, including an increase in the synthesis of compatible solutes. However, its proteome displays completely different patterns when grown in no salt (i.e. 0% NaCl added). An increase in stress related chaperones was accompanied by enhanced abundance of carbon uptake and fixation proteins and periplasmic iron binding protein. Further studies need to be undertaken to further understand this response to low salt. Proteins were also identified as salt responsive which play no noticeable role in Synechocystis tolerance, and these also require further characterisation.

Conclusion

With the increasing number of genome sequencing projects being undertaken and rapid advances in proteomics techniques, the opportunity for making greater discoveries in the post-genomics field has been recognised. In effect, genomic information can be decoded into a functional protein interpretation.

Understanding the salt response in cyanobacteria will make a relevant impact on ultimately understanding the detrimental effects of salinity on crops plants. Synechocystis is the best characterised cyanobacterium from a salt stress perspective and this stands true in the emerging field of proteomics. Gaining a systems level understanding of the salt stress response in this cyanobacterium requires generating and interpreting quantitative proteome data. Techniques are constantly and rapidly being developed and tested to overcome current limitations, and this includes the opportunity to study more difficult and challenging systems.

Synechocystis alters the protein composition of extracellular layers in response to salt stress, and in particular increases expression of ABC-transporters involved in nutrient acquisition in the plasma membrane. A reorganisation of the thylakoid membranes seems likely through the identification of salt-induced Vipp1, and the increase in energy capacity using PSI and respiration is important for acclimation. Specific salt stress proteins, as well as general stress proteins play an obvious role, and large changes in carbon metabolism are seen and these may be related to the long-term production of compatible solutes. Finally, a large array of hypothetical proteins were identified as playing a crucial part in adaptation, but only a small number had associated putative functions. These proteins require further investigation.

Important advances have been made in increasing proteome coverage through novel combinations of protein and peptide fractionation techniques. By combining reliable protein identifications with accurate and reproducible quantitation data, information on hundreds of proteins can be obtained in a single experimental workflow. The use of iTRAQ technology is gaining popularity here. Finally, proteomics is no longer confined to those organisms which have their genomes partially or fully sequenced. Efforts to develop techniques and interpretation software to analyse these proteomes, is beginning to reward the scientific community by producing interesting findings on the biological behaviour of fascinating, environmentally-significant organisms.

References

  1. Boyer JS: Plant productivity and environment. Science. 1982, 218 (4571): 443-448.

    CAS  Google Scholar 

  2. Nelson DE, Shen B, Bohnert HJ: Salinity tolerance mechanisms, models and the metabolic engineering of complex traits: Principles and Methods. Genet Eng (N Y). Edited by: Setlow J. 1998, New York , Plenum Press, 20: 153-176.

    Google Scholar 

  3. Hart BT, Lake PS, Webb JA, Grace MR: Ecological risk to aquatic systems from salinity increases. Aust J Bot. 2003, 51: 689-702.

    CAS  Google Scholar 

  4. Bryant DA: The cyanobacterial photosynthetic apparatus: comparisons to those of higher plants and photosynthetic bacteria. Photosynthetic picoplankton. Edited by: Li TPWKW. 1986, Can. Bull. Fish. Aquat. Sci., 214: 423-500.

    Google Scholar 

  5. Joset F, Jeanjean R, Hagemann M: Dynamics of response of cyanobacteria to salt stress: deciphering the molecular events. Physiol Plant. 1996, 96: 738-744.

    CAS  Google Scholar 

  6. Stal L: Cyanobacteria. Algae and cyanobacteria in extreme environments. 2007, 11: 659-680.

    Google Scholar 

  7. Reed RH, Stewart WDP: The responses of cyanobacteria to salt stress. Biochemistry of the algae and cyanobacteria. Edited by: L.J. Rogers JRG. 1988, Oxford , Oxford Science Publisher, 217-231.

    Google Scholar 

  8. James KR, Cant B, Ryan T: Responses of fresh water biota to rising salinity levels and implications for saline water management: A review. Aust J Bot. 2003, 51 (6): 703-713.

    CAS  Google Scholar 

  9. Galinski EA: Osmoadaptation in bacteria. Adv Microb Physiol. 1995, 37: 272-328.

    CAS  Google Scholar 

  10. Yancey PH: Water stress, osmolytes and proteins. Amer Zool. 2001, 41 (4): 699-709.

    CAS  Google Scholar 

  11. Jeruzalmi D, Steitz TA: Use of organic cosmotropic solutes to crystallize flexible proteins: application to T7 RNA polymerase and its complex with the inhibitor T7 lysozyme. J Mol Biol. 1997, 274 (5): 748-756.

    CAS  Google Scholar 

  12. Brown CR, Hong-Brown LQ, Biwersi J, Verkman AS, Welch WJ: Chemical chaperones correct the mutant phenotype of the delta F508 cystic fibrosis transmembrane conductance regulator protein. Cell Stress Chaperones. 1996, 1 (2): 117-125.

    CAS  Google Scholar 

  13. Roberts MF: Organic compatible solutes of halotolerant and halophilic microorganisms. Saline Systems. 2005, 1: 5-

    Google Scholar 

  14. Desmarais D, Jablonski PE, Fedarko NS, Roberts MF: 2-Sulfotrehalose, a novel osmolyte in haloalkaliphilic archaea. J Bacteriol. 1997, 179 (10): 3146-3153.

    CAS  Google Scholar 

  15. Gabbay-Azaria R, Tel-Or E: Mechanisms of salt tolerance in cyanobacteria. Plant Responses to the Environment. Edited by: Gresshoff PM. 1993, Boca Raton , CRC Press, 692-698.

    Google Scholar 

  16. Hagemann M, Zuther E: Selection and characterization of mutants of the cyanobacterium Synechocystis sp. PCC 6803 unable to tolerate high salt concentrations. Arch Microbiol. 1992, 158 (6): 429-434.

    CAS  Google Scholar 

  17. JVIRGEL. [http://www.jvirgel.de/]

  18. Kaneko T, Sato S, Kotani H, Tanaka A, Asamizu E, Nakamura Y, Miyajima N, Hirosawa M, Sugiura M, Sasamoto S, Kimura T, Hosouchi T, Matsuno A, Muraki A, Nakazaki N, Naruo K, Okumura S, Shimpo S, Takeuchi C, Wada T, Watanabe A, Yamada M, Yasuda M, Tabata S: Sequence analysis of the genome of the unicellular cyanobacterium Synechocystis sp. strain PCC6803. II. Sequence determination of the entire genome and assignment of potential protein-coding regions (supplement). DNA Res. 1996, 3 (3): 185-209.

    CAS  Google Scholar 

  19. Grigorieva G, Shestakov SV: Transformation in the cyanobacterium Synechocystis sp. PCC6803. FEMS Microbiol Lett. 1982, 13: 367-370.

    CAS  Google Scholar 

  20. Burja AM, Dhamwichukorn S, Wright PC: Cyanobacterial postgenomic research and systems biology. Trends Biotechnol. 2003, 21 (11): 504-511.

    CAS  Google Scholar 

  21. Ferjani A, Mustardy L, Sulpice R, Marin K, Suzuki I, Hagemann M, Murata N: Glucosylglycerol, a compatible solute, sustains cell division under salt stress. Plant Physiol. 2003, 131 (4): 1628-1637.

    CAS  Google Scholar 

  22. Hagemann M, Fulda S, Schubert H: DNA, RNA, and protein synthesis in the cyanobacterium Synechocystis sp. PCC 6803 adapted to different salt concentrations. Curr Microbiol. 1994, 28 (4): 201-207.

    CAS  Google Scholar 

  23. Hagemann M, Jeanjean R, Fulda S, Havaux M, Joset F, Erdmann N: Flavodoxin accumulation contributes to enhanced cyclic electron flow around photosystem I in salt-stressed cells of Synechocystis sp. strain PCC 6803. Physiol Plant. 1999, 105 (4): 670-678.

    CAS  Google Scholar 

  24. Jeanjean R, Matthijs HCP, Onana B, Havaux M, Joset F: Exposure of the cyanobacterium Synechocystis PCC6803 to salt stress induces concerted changes in respiration and photosynthesis . Plant Cell Physiol. 1993, 34: 1073-1079.

    CAS  Google Scholar 

  25. Marin K, Huckauf J, Fulda S, Hagemann M: Salt-dependent expression of glucosylglycerol-phosphate synthase, involved in osmolyte synthesis in the cyanobacterium Synechocystis sp. strain PCC 6803. J Bacteriol. 2002, 184 (11): 2870-2877.

    CAS  Google Scholar 

  26. Marin K, Suzuki I, Yamaguchi K, Ribbeck K, Yamamoto H, Kanesaki Y, Hagemann M, Murata N: Identification of histidine kinases that act as sensors in the perception of salt stress in Synechocystis sp. PCC 6803. PNAS. 2003, 100 (15): 9061-9066.

    CAS  Google Scholar 

  27. Reed RH, Borowitzka LJ, Mackay MA, Chudek JA, Foster R, Warr SRC, Moore DJ, Stewart WDP: Organic solute accumulation in osmotically stressed cyanobacteria. FEMS Microbiol Lett. 1986, 39 (1-2): 51-56.

    CAS  Google Scholar 

  28. Schubert H, Fulda S, Hagemann M: Effects of adaptation to different salt concentrations on photosynthesis and pigmentation of the cyanobacterium Synechocystis sp. PCC 6803. J Plant Physiol. 1993, 142: 291–295-

    Google Scholar 

  29. Singh SC, Sinha RP, Hader DP: Role of lipids and fatty acids in stress tolerance in cyanobacteria. Acta Protozoologica. 2002, 41: 297-308.

    CAS  Google Scholar 

  30. Tang D, Shi S, Li D, Hu C, Liu Y: Physiological and biochemical responses of Scytonema javanicum (cyanobacterium) to salt stress. Journal of Arid Environments. 2007, 71 (3): 312-320.

    Google Scholar 

  31. Elanskaya IV, Karandashova IV, Bogachev AV, M. H: Functional analysis of the Na+/H+ antiporter encoding genes of the cyanobacterium Synechocystis PCC 6803. Biochemistry (Mosc). 2002, 67 (4): 432-440.

    Google Scholar 

  32. Shoumskaya MA, Paithoonrangsarid K, Kanesaki Y, Los DA, Zinchenko VV, Tanticharoen M, Suzuki I, Murata N: Identical Hik-Rre systems are involved in perception and transduction of salt signals and hyperosmotic signals but regulate the expression of individual genes to different extents in Synechocystis. J Biol Chem. 2005, 280 (22): 21531-21538.

    CAS  Google Scholar 

  33. Vinnemeier J, Kunert A, Hagemann M: Transcriptional analysis of the isiAB operon in salt-stressed cells of the cyanobacterium Synechocystis sp. PCC6803. FEMS Microbiol Lett. 1998, 169: 323-330.

    CAS  Google Scholar 

  34. Bohnert HJ, Ayoubi P, Borchert C, Bressan RA, Burnap RL, Cushman JC, Cushman MA, Deyholos M, Galbraith DW, Hasegawa PM, Jenks M, Kawasaki S, Koiwa H, Kore-eda S, Lee BH, Michalowski CB, Misawa E, Nomura M, Ozturk M, Postier B, Prade R, Song CP, Tanaka Y, Wang H, Zhu JK: A genomics approach towards salt stress tolerance. Plant Physiol Biochem. 2001, 39: 295-311.

    CAS  Google Scholar 

  35. Kanesaki Y, Suzuki I, Allakhverdiev SI, Mikami K, Murata N: Salt Stress and hyperosmotic stress regulate the expression of different sets of genes in Synechocystis sp. PCC 6803. Biochem Biophys Res Commun. 2002, 290 (1): 339-348.

    CAS  Google Scholar 

  36. Marin K, Kanesaki Y, Los DA, Murata N, Suzuki I, Hagemann M: Gene expression profiling reflects physiological processes in salt acclimation of Synechocystis sp. strain PCC 6803. Plant Physiol. 2004, 136 (2): 3290-3300.

    CAS  Google Scholar 

  37. Miao X, Wu Q, Wu G, Zhao N: Sucrose accumulation in salt-stressed cells of agp gene deletion-mutant in cyanobacterium Synechocystis sp PCC 6803. FEMS Microbiol Lett. 2003, 218 (1): 71-77.

    CAS  Google Scholar 

  38. Fulda S, Huang F, Nilsson F, Hagemann M, Norling B: Proteomics of Synechocystis sp. strain PCC 6803: Identification of periplasmic proteins cells grown at low and high salt concentrations. Eur J Biochem. 2000, 267: 5900-5907.

    CAS  Google Scholar 

  39. Fulda S, Mikkat S, Huang F, Huckauf J, Marin K, Norling B, Hagemann M: Proteome analysis of salt stress response in the cyanobacterium Synechocystis sp. strain PCC 6803. Proteomics. 2006, 6 (9): 2733-2745.

    CAS  Google Scholar 

  40. Hagemann M, Techel D, Rensing L: Comparison of salt- and heat-induced alterations of protein synthesis in the cyanobacterium Synechocystis sp. PCC 6803. Arch Microbiol. 1991, 155 (6): 587-592.

    CAS  Google Scholar 

  41. Fulda S, Mikkat S, Schroder W, Hagemann M: Isolation of salt-induced periplasmic proteins from Synechocystis sp. strain PCC 6803. Arch Microbiol. 1999, 171 (3): 214-217.

    CAS  Google Scholar 

  42. Wasinger VC, Cordwell SJ, Cerpa-Poljak A, Yan JX, Gooley AA, Wilkins MR, Duncan MW, Harris R, Williams KL, Humphery-Smith I: Progress with gene-product mapping of the Mollicutes: Mycoplasma genitalium. Electrophoresis. 1995, 16 (7): 1090-1094.

    CAS  Google Scholar 

  43. Chalmers MJ, Gaskell SJ: Advances in mass spectrometry for proteome analysis. Curr Opin Biotechnol. 2000, 11 (4): 384-390.

    CAS  Google Scholar 

  44. Gygi SP, Rochon Y, Franza BR, Aebersold R: Correlation between protein and mRNA abundance in yeast. Mol Cell Biol. 1999, 19 (3): 1720-1730.

    CAS  Google Scholar 

  45. Lu P, Vogel C, Wang R, Yao X, Marcotte EM: Absolute protein expression profiling estimates the relative contributions of transcriptional and translational regulation. Nat Biotechnol. 2007, 25 (1): 117-124.

    CAS  Google Scholar 

  46. Wand AJ: Dynamic activation of protein function: a view emerging from NMR spectroscopy. Nat Struct Biol. 2001, 8 (11): 926-931.

    CAS  Google Scholar 

  47. Kaneko T, Nakamura Y, Wolk CP, Kuritz T, Sasamoto S, Watanabe A, Iriguchi M, Ishikawa A, Kawashima K, Kimura T, Kishida Y, Kohara M, Matsumoto M, Matsuno A, Muraki A, Nakazaki N, Shimpo S, Sugimoto M, Takazawa M, Yamada M, Yasuda M, Tabata S: Complete genomic sequence of the filamentous nitrogen-fixing cyanobacterium Anabaena sp. strain PCC 7120. DNA Res. 2001, 8 (5): 205-13; 227-53.

    CAS  Google Scholar 

  48. Klose J: Protein mapping by combined isoelectric focusing and electrophoresis of mouse tissues. A novel approach to testing for induced point mutations in mammals. Humangenetik. 1975, 26 (3): 231-243.

    CAS  Google Scholar 

  49. O'Farrell PH: High resolution two-dimensional electrophoresis of proteins. J Biol Chem. 1975, 250 (10): 4007-4021.

    Google Scholar 

  50. Choe LH, Lee KH: A comparison of three commercially available isoelectric focusing units for proteome analysis: The Multiphor, the IPGphor and the Protean IEF cell. Electrophoresis. 2000, 21 (5): 993-1000.

    CAS  Google Scholar 

  51. Packer NH, Pawlak A, Kett WC, Gooley AA, Redmond JW, Williams KL: Proteome analysis of glycoforms: a review of strategies for the microcharacterisation of glycoproteins separated by two-dimensional polyacrylamide gel electrophoresis. Electrophoresis. 1997, 18 (3-4): 452-460.

    CAS  Google Scholar 

  52. Robertson ES, Nicholson AW: Phosphorylation of Escherichia coli translation initiation factors by the bacteriophage T7 protein kinase. Biochemistry (Mosc). 1992, 31 (20): 4822-4827.

    CAS  Google Scholar 

  53. Westermeier R, naven T: Proteomics in Practice. 2002, Europe , Amersham Biosciences, 3rd

    Google Scholar 

  54. Simon WJ, Hall JJ, Suzuki I, Murata N, Slabas AR: Proteomic study of the soluble proteins from the unicellular cyanobacterium Synechocystis sp. PCC6803 using automated matrix-assisted laser desorption/ionization-time of flight peptide mass fingerprinting. Proteomics. 2002, 2 (12): 1735-1742.

    CAS  Google Scholar 

  55. Bjellqvist B, Ek K, Righetti PG, Gianazza E, Gorg A, Westermeier R, Postel W: Isoelectric focusing in immobilized pH gradients: principle, methodology and some applications. J Biochem Biophys Methods. 1982, 6 (4): 317-339.

    CAS  Google Scholar 

  56. Lopez MF: Better approaches to finding the needle in a haystack: Optimizing proteome analysis through automation. Electrophoresis. 2000, 21 (6): 1082-1093.

    CAS  Google Scholar 

  57. Libby K, Engelstein M: Automating and simplifying in-gel digestion, purification and MALDI spotting. Am Biotechnol Lab. 2003, 14:

    Google Scholar 

  58. Barry RC, Alsaker BL, Robison-Cox JF, Dratz EA: Quantitative evaluation of sample application methods for semipreparative separations of basic proteins by two-dimensional gel electrophoresis. Electrophoresis. 2003, 24 (19-20): 3390-3404.

    CAS  Google Scholar 

  59. Link AJ: Multi-dimensional peptide separtations in proteomics. Trends Biotechnol. 2002, 20: S8- S13.

    CAS  Google Scholar 

  60. Molloy MP: Two-dimensional electrophoresis of membrane proteins using immobilized pH gradients. Anal Biochem. 2000, 280 (1): 1-10.

    CAS  Google Scholar 

  61. Rabilloud T: Solubilization of proteins in 2-D electrophoresis: An outline. Methods Mol Biol. 1999, 112: 9-19.

    CAS  Google Scholar 

  62. Berggren K, Steinberg TH, Lauber WM, Carroll JA, Lopez MF, Chernokalskaya E, Zieske L, Diwu Z, Haugland RP, Patton WF: A luminescent ruthenium complex for ultrasensitive detection of proteins immobilized on membrane supports. Anal Biochem. 1999, 276 (2): 129-143.

    CAS  Google Scholar 

  63. Lopez MF, Berggren K, Chernokalskaya E, Lazarev A, Robinson M, Patton WF: A comparison of silver stain and SYPRO Ruby Protein Gel Stain with respect to protein detection in two-dimensional gels and identification by peptide mass profiling. Electrophoresis. 2000, 21 (17): 3673-3683.

    CAS  Google Scholar 

  64. Shevchenko A, Wilm M, Vorm O, Mann M: Mass spectrometric sequencing of proteins silver-stained polyacrylamide gels. Anal Chem. 1996, 68 (5): 850-858.

    CAS  Google Scholar 

  65. Unlu M, Morgan ME, Minden JS: Difference gel electrophoresis: a single gel method for detecting changes in protein extracts. Electrophoresis. 1997, 18 (11): 2071-2077.

    CAS  Google Scholar 

  66. Ong SE, Blagoev B, Kratchmarova I, Kristensen DB, Steen H, Pandey A, Mann M: Stable isotope labeling by amino acids in cell culture, SILAC, as a simple and accurate approach to expression proteomics. Mol Cell Proteomics. 2002, 1 (5): 376-386.

    CAS  Google Scholar 

  67. Krijgsveld J, Ketting RF, Mahmoudi T, Johansen J, Artal-Sanz M, Verrijzer CP, Plasterk RH, Heck AJ: Metabolic labeling of C. elegans and D. melanogaster for quantitative proteomics. Nat Biotechnol. 2003, 21 (8): 927-931.

    CAS  Google Scholar 

  68. Snijders AP, de Vos MG, de Koning B, Wright PC: A fast method for quantitative proteomics based on a combination between two-dimensional electrophoresis and 15N-metabolic labelling. Electrophoresis. 2005, 26 (16): 3191-3199.

    CAS  Google Scholar 

  69. Opiteck GJ, Jorgenson JW: Two-dimensional SEC/RPLC coupled to mass spectrometry for the analysis of peptides. Anal Chem. 1997, 69 (13): 2283-2291.

    CAS  Google Scholar 

  70. Swanson RV, Glazer AN: Separation of phycobiliprotein subunits by reverse-phase high-pressure liquid chromatography. Anal Biochem. 1990, 188 (2): 295-299.

    CAS  Google Scholar 

  71. Peng J, Elias JE, Thoreen CC, Licklider LJ, Gygi SP: Evaluation of multidimensional chromatography coupled with tandem mass spectrometry (LC/LC-MS/MS) for large-scale protein analysis: the yeast proteome. J Proteome Res. 2003, 2 (1): 43-50.

    CAS  Google Scholar 

  72. Chen J, Balgley BM, DeVoe DL, Lee CS: Capillary isoelectric focusing-based multidimensional concentration/separation platform for proteome analysis. Anal Chem. 2003, 75 (13): 3145-3152.

    CAS  Google Scholar 

  73. Xiao Z, Conrads TP, Lucas DA, Janini GM, Schaefer CF, Buetow KH, Issaq HJ, Veenstra TD: Direct ampholyte-free liquid-phase isoelectric peptide focusing: application to the human serum proteome. Electrophoresis. 2004, 25 (1): 128-133.

    CAS  Google Scholar 

  74. Fenn JB, Mann M, Meng CK, Wong SF, Whitehouse CM: Electrospray ionization for mass spectrometry of large biomolecules. Science. 1989, 246 (4926): 64-71.

    CAS  Google Scholar 

  75. Karas M, Hillenkamp F: Laser desorption ionization of proteins with molecular masses exceeding 10,000 daltons. Anal Chem. 1988, 60 (20): 2299-2301.

    CAS  Google Scholar 

  76. Aebersold R, Mann M: Mass spectrometry-based proteomics. Nature. 2003, 422 (6928): 198-207.

    CAS  Google Scholar 

  77. Blueggel M, Chamrad D, Meyer HE: Bioinformatics in proteomics. Curr Pharm Biotechnol. 2004, 5 (1): 79-88.

    CAS  Google Scholar 

  78. Huang H, Hu ZZ, Arighi CN, Wu CH: Integration of bioinformatics resources for functional analysis of gene expression and proteomic data. Front Biosci. 2007, 12: 5071-5088.

    CAS  Google Scholar 

  79. Cho CW, Lee SH, Choi J, Park SJ, Ha DJ, Kim HJ, Kim CW: Improvement of the two-dimensional gel electrophoresis analysis for the proteome study of Halobacterium salinarum. Proteomics. 2003, 3 (12): 2325-2329.

    CAS  Google Scholar 

  80. Jiang L, He L, Fountoulakis M: Comparison of protein precipitation methods for sample preparation prior to proteomic analysis. J Chromatogr A. 2004, 1023 (2): 317-320.

    CAS  Google Scholar 

  81. Yuan X, Desiderio DM: Proteomics analysis of human cerebrospinal fluid. J Chromatogr B Biomed Appl. 2005, 815 (1-2): 179-189.

    CAS  Google Scholar 

  82. Damerval C, DeVienne D, Zivy M, Thiellement H: Technical improvements in two-dimensional electrophoresis increase the level of genetic variation detected in wheat-seedling protein. Electrophoresis. 1986, 7: 53-54.

    Google Scholar 

  83. Wang Y, Sun J, Chitnis PR: Proteomic study of the peripheral proteins from thylakoid membranes of the cyanobacterium Synechocystis sp. PCC 6803. Electrophoresis. 2000, 21 (9): 1746-1754.

    CAS  Google Scholar 

  84. Gorg A: Advances in 2D gel techniques. Trends Biotechnol. 2000, 18: 3-6.

    Google Scholar 

  85. Berkelman T, Stenstedt T: 2D Electrophoresis: Principles and methods. 2001, Uppsala, Sweden , Amersham Pharmacia Biotech, 17-25.

    Google Scholar 

  86. Heppelmann C, Benson LM, Bergen HR: A simple method to remove contaminating salt from IPG strips prior to IEF. Electrophoresis. 2007, 28 (21): 3988-3991.

    CAS  Google Scholar 

  87. Lion N, Gellon JO, Jensen H, Girault HH: On-chip protein sample desalting and preparation for direct coupling with electrospray ionization mass spectrometry. J Chromatogr A. 2003, 1003: 11-19.

    CAS  Google Scholar 

  88. Naldrett MJ, Zeidler R, Wilson KE, Kocourek A: Concentration and desalting of peptide and protein samples with a newly developed C18 membrane in a microspin column format. J Biomol Tech. 2005, 16 (4): 423-428.

    Google Scholar 

  89. Annesley TM: Ion suppression in mass spectrometry. Clin Chem. 2003, 49 (7): 1041-1044.

    CAS  Google Scholar 

  90. Jackson AU, Talaty N, Cooks RG, Van Berkel GJ: Salt tolerance of desorption electrospray ionization (DESI). J Am Soc Mass Spectrom. 2007, 18 (12): 2218-2225.

    CAS  Google Scholar 

  91. Nakamura Y, Kaneko T, Tabata S: CyanoBase, the genome database for Synechocystis sp. strain PCC6803: status for the year 2000. Nucl Acids Res. 2000, 28 (1): 72-73.

    CAS  Google Scholar 

  92. Norling B, Zak E, Andersson B, Pakrasi H: 2D-isolation of pure plasma and thylakoid membranes from the cyanobacterium Synechocystis sp. PCC 6803. FEBS Lett. 1998, 436 (2): 189-192.

    CAS  Google Scholar 

  93. Huang F, Parmryd I, Nilsson F, Persson AL, Pakrasi HB, Andersson B, Norling B: Proteomics of Synechocystis sp. Strain PCC 6803: Identification of Plasma Membrane Proteins. Mol Cell Proteomics. 2002, 1 (12): 956-966.

    CAS  Google Scholar 

  94. Pisareva T, Shumskaya M, Maddalo G, Ilag L, Norling B: Proteomics of Synechocystis sp. PCC 6803. Identification of novel integral plasma membrane proteins. FEBS Journal. 2007, 274 (3): 791-804.

    CAS  Google Scholar 

  95. Kashino Y, Harayama T, Pakrasi HB, Satoh K: Preparation of membrane proteins for analysis by two-dimensional gel electrophoresis. J Chromatogr B Biomed Appl. 2007, 849 (1-2): 282-292.

    CAS  Google Scholar 

  96. Herranen M, Battchikova N, Zhang P, Graf A, Sirpio S, Paakkarinen V, Aro EM: Towards functional proteomics of membrane protein complexes in Synechocystis sp. PCC 6803. Plant Physiol. 2004, 134 (1): 470-481.

    CAS  Google Scholar 

  97. Rippka R, Deruelles J, Waterbury JB: Generic assignments, strain histories and properties of pure cultures of cyanobacteria. J Gen Microbiol. 1979, 111 (1): 1-61.

    Google Scholar 

  98. Neu HC, LA H: The release of enzymes from Escherichia coli by osmotic shock and during the formation of spheroplasts. J Biol Chem. 1965, 240: 3685–3692-

    Google Scholar 

  99. Hagemann M, Erdmann N: Activation and pathway of glucosylglycerol synthesis in the cyanobacterium Synechocystis sp. PCC 6803 . Microbiology. 1994, 140: 1427–1431-

    Google Scholar 

  100. Wittig I, Braun HP, Schagger H: Blue native PAGE. Nat Protocols. 2006, 1 (1): 418-428.

    CAS  Google Scholar 

  101. Camacho-Carvajal MM, Wollscheid B, Aebersold R, Steimle V, Schamel WWA: Two-dimensional Blue native/SDS gel electrophoresis of multi-protein complexes from whole cellular lysates: a proteomics approach. Mol Cell Proteomics. 2004, 3 (2): 176-182.

    CAS  Google Scholar 

  102. 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 (3): 910-920.

    CAS  Google Scholar 

  103. Wessel D, Flugge UI: A method for the quantitative recovery of protein in dilute solution in the presence of detergents and lipids. Anal Biochem. 1984, 138 (1): 141-143.

    CAS  Google Scholar 

  104. Tolle J, Michel KP, Kruip J, Kahmann U, Preisfeld A, Pistorius EK: Localization and function of the IdiA homologue Slr1295 in the cyanobacterium Synechocystis sp. strain PCC 6803. Microbiology. 2002, 148 (Pt 10): 3293-3305.

    CAS  Google Scholar 

  105. Wilkinson MJ, Northcote DH: Plasma membrane ultrastructure during plant protoplast plasmolysis, isolation and wall regeneration: a freeze-fracture study. J Cell Sci. 1980, 42 (1): 401-415.

    CAS  Google Scholar 

  106. Kroll D, Meierhoff K, Bechtold N, Kinoshita M, Westphal S, Vothknecht UC, Soll J, Westhoff P: VIPP1, a nuclear gene of Arabidopsis thaliana essential for thylakoid membrane formation. Proc Natl Acad Sci U S A. 2001, 98 (7): 4238-4242.

    CAS  Google Scholar 

  107. Gardy JL, Laird MR, Chen F, Rey S, Walsh CJ, Ester M, Brinkman FS: PSORTb v.2.0: expanded prediction of bacterial protein subcellular localization and insights gained from comparative proteome analysis. Bioinformatics. 2005, 21 (5): 617-623.

    CAS  Google Scholar 

  108. Juncker AS, Willenbrock H, Von Heijne G, Brunak S, Nielsen H, Krogh A: Prediction of lipoprotein signal peptides in Gram-negative bacteria. Protein Sci. 2003, 12 (8): 1652-1662.

    CAS  Google Scholar 

  109. Sudhir PR, Pogoryelov D, Kovacs L, Garab G, Murthy SD: The effects of salt stress on photosynthetic electron transport and thylakoid membrane proteins in the cyanobacterium Spirulina platensis. J Biochem Mol Biol. 2005, 38 (4): 481-485.

    CAS  Google Scholar 

  110. Burnette WN: "Western blotting": electrophoretic transfer of proteins from sodium dodecyl sulfate--polyacrylamide gels to unmodified nitrocellulose and radiographic detection with antibody and radioiodinated protein A. Anal Biochem. 1981, 112 (2): 195-203.

    CAS  Google Scholar 

  111. Chen Y, Drysdale J: Detection of iron binding proteins by a blotting technique. Anal Biochem. 1993, 212 (1): 47-49.

    CAS  Google Scholar 

  112. Wirth PJ, Romano A: Staining methods in gel electrophoresis, including the use of multiple detection methods. J Chromatogr A. 1995, 698: 123-143.

    CAS  Google Scholar 

  113. Gygi SP, Corthals GL, Zhang Y, Rochon Y, Aebersold R: Evaluation of two-dimensional gel electrophoresis-based proteome analysis technology. Proc Natl Acad Sci U S A. 2000, 97 (17): 9390-9395.

    CAS  Google Scholar 

  114. Kashino Y: Separation methods in the analysis of protein membrane complexes. J Chromatogr B Analyt Technol Biomed Life Sci. 2003, 797 (1-2): 191-216.

    CAS  Google Scholar 

  115. Kato Y, Nakamura K, Kitamura T, Tsuda T, Hasegawa M, Sasaki H: Effect of chromatographic conditions on resolution in high-performance ion-exchange chromatography of proteins on macroporous anion-exchange resin. J Chromatogr A. 2004, 1031 (1-2): 101-105.

    CAS  Google Scholar 

  116. Chan KC, Lucas DA, Hise D, Schaefer CF, Xiao Z, Janini GM, Buetow KH, Issaq HJ, Veenstra TD, Conrads TP: Analysis of the human serum proteome. Clinical Proteomics. 2004, 1: 101-226.

    Google Scholar 

  117. Ostrowski LE, Blackburn K, Radde KM, Moyer MB, Schlatzer DM, Moseley A, Boucher RC: A proteomic analysis of human cilia: identification of novel components. Mol Cell Proteomics. 2002, 1 (6): 451-465.

    CAS  Google Scholar 

  118. Gan CS, Reardon KF, Wright PC: Comparison of protein and peptide prefractionation methods for the shotgun proteomic analysis of Synechocystis sp. PCC 6803. Proteomics. 2005, 5 (9): 2468-2478.

    CAS  Google Scholar 

  119. Rabilloud T: Membrane proteins ride shotgun. Nat Biotechnol. 2003, 21 (5): 508-510.

    CAS  Google Scholar 

  120. Santoni V, Molloy M, Rabilloud T: Membrane proteins and proteomics: un amour impossible?. Electrophoresis. 2000, 21 (6): 1054-1070.

    CAS  Google Scholar 

  121. Barrios-Llerena ME, Chong PK, Gan CS, Snijders AP, Reardon KF, Wright PC: Shotgun proteomics of cyanobacteria--applications of experimental and data-mining techniques. Brief Funct Genomic Proteomic. 2006, 5 (2): 121-132.

    CAS  Google Scholar 

  122. Yao X, Freas A, Ramirez J, Demirev PA, Fenselau C: Proteolytic 18O labeling for comparative proteomics: model studies with two serotypes of adenovirus. Anal Chem. 2001, 73 (13): 2836-2842.

    CAS  Google Scholar 

  123. Gygi SP, Rist B, Gerber SA, Turecek F, Gelb MH, Aebersold R: Quantitative analysis of complex protein mixtures using isotope-coded affinity tags. Nat Biotechnol. 1999, 17 (10): 994-999.

    CAS  Google Scholar 

  124. Cagney G, Emili A: Mass-coded abundance tagging for protein identification and relative abundance determination in proteomic experiments. The Proteomics Protocols Handbook. 2005, 193-206.

    Google Scholar 

  125. Kirkpatrick DS, Gerber SA, Gygi SP: The absolute quantification strategy: a general procedure for the quantification of proteins and post-translational modifications. Methods. 2005, 35 (3): 265-273.

    CAS  Google Scholar 

  126. Ross PL, Huang YN, Marchese JN, Williamson B, Parker K, Hattan S, Khainovski N, Pillai S, Dey S, Daniels S, Purkayastha S, Juhasz P, Martin S, Bartlet-Jones M, He F, Jacobson A, Pappin DJ: Multiplexed protein quantitation in Saccharomyces cerevisiae using amine-reactive isobaric tagging reagents. Mol Cell Proteomics. 2004, 3 (12): 1154-1169.

    CAS  Google Scholar 

  127. Pandhal J, Wright PC, Biggs CA: A quantitative proteomic analysis of light adaptation in a globally significant marine cyanobacterium Prochlorococcus marinus MED4. J Proteome Res. 2007, 6 (3): 996-1005.

    CAS  Google Scholar 

  128. Chong PK, Gan CS, Pham TK, Wright PC: Isobaric tags for relative and absolute quantitation (iTRAQ) reproducibility: Implication of multiple injections. J Proteome Res. 2006, 5 (5): 1232-1240.

    CAS  Google Scholar 

  129. Gan CS, Chong PK, Pham TK, Wright PC: Technical, experimental, and biological variations in isobaric tags for relative and absolute quantitation (iTRAQ). J Proteome Res. 2007, 6 (2): 821-827.

    CAS  Google Scholar 

  130. Stensjo K, Ow SY, Barrios-Llerena ME, Lindblad P, Wright PC: An iTRAQ-based quantitative analysis to elaborate the proteomic response of Nostoc sp. PCC 7120 under N2 fixing conditions. J Proteome Res. 2007, 6 (2): 621-635.

    Google Scholar 

  131. Oda Y, Huang K, Cross FR, Cowburn D, Chait BT: Accurate quantitation of protein expression and site-specific phosphorylation. Proc Natl Acad Sci U S A. 1999, 96 (12): 6591-6596.

    CAS  Google Scholar 

  132. Choe LH, Franck KAZ, Lee KH: A comparison of the consistency of proteome quantitation using two-dimensional electrophoresis and shotgun isobaric tagging in Escherichia coli cells. Electrophoresis. 2005, 26 (12): 2437-2449.

    CAS  Google Scholar 

  133. Wu WW, Wang G, Baek SJ, Shen RF: Comparative study of three proteomic quantitative methods, DIGE, cICAT, and iTRAQ, using 2D Gel- or LC-MALDI TOF/TOF. J Proteome Res. 2006, 5 (3): 651-658.

    CAS  Google Scholar 

  134. Liska AJ, Shevchenko A: Expanding the organismal scope of proteomics: cross-species protein identification by mass spectrometry and its implications. Proteomics. 2003, 3 (1): 19-28.

    CAS  Google Scholar 

  135. Shevchenko A, Sunyaev S, Loboda A, Shevchenko A, Bork P, Ens W, Standing KG: Charting the proteomes of organisms with unsequenced genomes by MALDI-quadrupole time-of-flight mass spectrometry and BLAST homology searching. Anal Chem. 2001, 73 (9): 1917-1926.

    CAS  Google Scholar 

  136. Habermann B, Oegema J, Sunyaev S, Shevchenko A: The power and the limitations of cross-species protein identification by mass spectrometry-driven sequence similarity searches. Mol Cell Proteomics. 2004, 3 (3): 238-249.

    CAS  Google Scholar 

  137. Lester PJ, Hubbard SJ: Comparative bioinformatic analysis of complete proteomes and protein parameters for cross-species identification in proteomics. Proteomics. 2002, 2 (10): 1392-1405.

    CAS  Google Scholar 

  138. Liska AJ, Shevchenko A: Combining mass spectrometry with database interrogation strategies in proteomics. TrAC, Trends Anal Chem. 2003, 22 (5): 291-298.

    CAS  Google Scholar 

  139. MS BLAST. [http://www.dove.embl-heidelberg.de/Blast2/msblast.html]

  140. Han Y, Ma B, Zhang K: SPIDER: software for protein identification from sequence tags with de novo sequencing error. Proc IEEE Comput Syst Bioinform Conf. 2004, 206-215.

    Google Scholar 

  141. Ma B, Zhang K, Hendrie C, Liang C, Li M, Doherty-Kirby A, Lajoie G: PEAKS: powerful software for peptide de novo sequencing by tandem mass spectrometry. Rapid Commun Mass Spectrom. 2003, 17 (20): 2337-2342.

    CAS  Google Scholar 

  142. Colinge J, Masselot A, Giron M, Dessingy T, Magnin J: OLAV: towards high-throughput tandem mass spectrometry data identification. Proteomics. 2003, 3 (8): 1454-1463.

    CAS  Google Scholar 

  143. Carr S, Aebersold R, Baldwin M, Burlingame A, Clauser K, Nesvizhskii A: The need for guidelines in publication of peptide and protein identification data: Working group on publication guidelines for peptide and protein identification data. Mol Cell Proteomics. 2004, 3 (6): 531-533.

    CAS  Google Scholar 

  144. Pandhal J, Snijders APL, Biggs C, Wright PC: A cross-species quantitative proteomic study of salt adaptation in a halotolerant environmental isolate using 15N metabolic labelling . Proteomics. 2008, In Press:

    Google Scholar 

  145. Gao X, Ren Z, Zhao Y, Zhang H: Overexpression of SOD2 Increases Salt Tolerance of Arabidopsis. Plant Physiol. 2003, 133 (4): 1873-1881.

    CAS  Google Scholar 

  146. Cioni P, Bramanti E, Strambini GB: Effects of sucrose on the internal dynamics of azurin. Biophys J. 2005, 88 (6): 4213-4222.

    CAS  Google Scholar 

  147. Grossmann J, Fischer B, Baerenfaller K, Owiti J, Buhmann JM, Gruissem W, Baginsky S: A workflow to increase the detection rate of proteins from unsequenced organisms in high-throughput proteomics experiments. Proteomics. 2007, 7 (23): 4245-4254.

    CAS  Google Scholar 

  148. Taji T, Seki M, Satou M, Sakurai T, Kobayashi M, Ishiyama K, Narusaka Y, Narusaka M, Zhu JK, Shinozaki K: Comparative genomics in salt tolerance between Arabidopsis and Arabidopsis-related halophyte salt cress using Arabidopsis microarray. Plant Physiol. 2004, 135 (3): 1697-1709.

    CAS  Google Scholar 

  149. Ashhuby BA: Biofouling studies on reverse osmosis desalination of hypersaline waters PhD Thesis. Chemical and Process Engineering. 2007, Sheffield , University of Sheffield, PhD: 221-

    Google Scholar 

Download references

Acknowledgements

Funding provided by The University of Sheffield, Biology Research Division Devolved Funds, and an EPSRC studentship. PCW also wishes to thank the EPSRC for the provision of an Advanced Research Fellowship (GR/A11311/01) and funding (GR/S84347/01 and EP/E036252/1). CAB also acknowledges the EPSRC for the provision of an Advanced Research Fellowship (EP/E053556/01).

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Catherine A Biggs.

Additional information

Competing interests

The authors declare that they have no competing interests.

Authors’ original submitted files for images

Rights and permissions

Open Access This article is published under license to BioMed Central Ltd. This is an Open Access article is distributed under the terms of the Creative Commons Attribution License ( https://creativecommons.org/licenses/by/2.0 ), which permits unrestricted use, distribution, and reproduction in any medium, provided the original work is properly cited.

Reprints and permissions

About this article

Cite this article

Pandhal, J., Wright, P.C. & Biggs, C.A. Proteomics with a pinch of salt: A cyanobacterial perspective. Aquat. Biosyst. 4, 1 (2008). https://doi.org/10.1186/1746-1448-4-1

Download citation

  • Received:

  • Accepted:

  • Published:

  • DOI: https://doi.org/10.1186/1746-1448-4-1

Keywords