Skip to main content
Journal of Chemical Biology logoLink to Journal of Chemical Biology
. 2010 Jun 19;4(1):3–29. doi: 10.1007/s12154-010-0043-5

Quantitative proteomics: assessing the spectrum of in-gel protein detection methods

Victoria J Gauci 1, Elise P Wright 1, Jens R Coorssen 1,
PMCID: PMC3022124  PMID: 21686332

Abstract

Proteomics research relies heavily on visualization methods for detection of proteins separated by polyacrylamide gel electrophoresis. Commonly used staining approaches involve colorimetric dyes such as Coomassie Brilliant Blue, fluorescent dyes including Sypro Ruby, newly developed reactive fluorophores, as well as a plethora of others. The most desired characteristic in selecting one stain over another is sensitivity, but this is far from the only important parameter. This review evaluates protein detection methods in terms of their quantitative attributes, including limit of detection (i.e., sensitivity), linear dynamic range, inter-protein variability, capacity for spot detection after 2D gel electrophoresis, and compatibility with subsequent mass spectrometric analyses. Unfortunately, many of these quantitative criteria are not routinely or consistently addressed by most of the studies published to date. We would urge more rigorous routine characterization of stains and detection methodologies as a critical approach to systematically improving these critically important tools for quantitative proteomics. In addition, substantial improvements in detection technology, particularly over the last decade or so, emphasize the need to consider renewed characterization of existing stains; the quantitative stains we need, or at least the chemistries required for their future development, may well already exist.

Keywords: Protein staining, Fluorescence, Coomassie, Sypro Ruby, 2D gel electrophoresis, Analytical proteomics

Introduction

Proteins are the primary functional agents of biological systems; they underlie and regulate metabolic processes, signal transduction, small molecule/ion transport, cell replication, and apoptosis [1, 2]. Examining the proteome (the entire complement of proteins expressed by a genome in a given biological sample—whole organism, tissue, fluid, cell, or organelle) is thus a critical way to analyze how a cell responds to its environment [3]. The pursuit of this endeavor has spanned over half a century and involved innovations in a range of different methodologies that have made the study of organisms across an array of molecular levels possible. Such a breadth of ‘omics analyses (i.e., genomics, proteomics, lipidomics, metabolomics, etc) has led to a Systems Biology approach that is now gradually enabling the integrated understanding of cell and organismal physiology [4]. At the protein level, although newer technologies for resolution and identification are regularly introduced, polyacrylamide gel electrophoresis (PAGE) remains the most accepted, widespread, and successfully implemented technique for the quantitative, high-resolution separation, and characterization of these critical molecules [5].

Utilization of PAGE as a single (native or detergent based PAGE) or second dimension of separation following isoelectric focusing (2D PAGE), has been shown to deliver very high resolution and reproducibility for protein separation [3, 612]. In a single analysis, 2D electrophoresis (2DE) provides information on protein charge, abundance, localization, isoforms, and post-translational modifications. The interface between PAGE and downstream mass spectrometry (MS) has provided perhaps the most important modern innovation for protein identification as well as more detailed analysis of post-translational modifications [13]. Although it was initially thought that 2DE was not ideal for resolving some proteins (i.e., membrane, very acidic, low abundance, and so forth), modern methodological optimizations minimize such suggested limitations; this is the power and versatility of a well-characterized, mature technology [14, 15]. Examples of these technical enhancements include the use of (narrow range) immobilized pH gradient strips or (extra) large gel formats to accommodate additional resolving area [14]. The complexity of the protein milieu can also be reduced using pre-/sub- and/or post-fractionation of samples [3, 13, 16]. With these techniques, protein resolution has been substantially and routinely improved. Taken together, all of these developments indicate that the main limitation to the amount of proteomic information obtained from 2DE is unlikely to be its capacity to resolve proteins but rather to be one of protein detection (i.e., stain sensitivity) [17].

Even though 2DE separates proteins such that their individual isoelectric points (pI) and subsequent electrophoretic mobility act as physical coordinates on a gel, none of this information can be assessed until the protein map has been visualized. This is achieved through the use of protein stains which generally bind to proteins in situ, within the polyacrylamide gel matrix. As might be expected of such a mature technology, there exists a diversity of such reagents including densitometric stains such as (colloidal) Coomassie Brilliant Blue or Silver, fluorescent stains including Sypro Ruby (SR), Deep Purple (DP), and the reactive CyDyes and Alexa Fluors [differential gel electrophoresis (DIGE) dyes] [10, 1821]. Despite this variety of available stains, the greatest challenge with regard to protein detection is to identify a protein stain that has the following characteristics:

  1. a routine and reproducible lowest limit of detection with optimal signal-to-noise ratio (SNR);

  2. a wide dynamic range, with a linear relationship between protein quantity and staining intensity;

  3. compatibility with downstream microchemical characterization techniques;

  4. ease of use;

  5. inexpensive, high throughput rates of use.

There is no protein stain currently available that possesses all of these desired properties.

This paper will thus review and critically analyze some common stains used for quantitative protein detection in an effort to identify stains that may currently best satisfy the breadth of proteomic applications. Specific criteria chosen as relevant measures of effectiveness include: (1) a reported lowest limit of detection (LLD; standard error noted when given), defined as the lowest concentration that delivers a pixel volume three standard deviations greater than that of the measured background [22]—the smaller this value, the more sensitive the stain; and (2) linear dynamic range (LDR), a measure of the total capacity of a stain for accurate quantification—defining a strictly linear relationship between quantity and signal with minimal deviation. Whenever such information is available stains will also be evaluated in terms of (3) inter-protein variability; (4) total number of spots detected (i.e., after 2DE); and (5) MS compatibility. Evaluation of quantitative performance will focus on those studies presenting a minimum of three of the above defined criteria for 1D and/or 2DE analyses.

Densitometric stains

Coomassie Brilliant Blue

Prior to the 1960s, the separation of protein mixtures was commonly performed using filter paper, cellulose acetate strips, starch, or agarose support mediums. Polyacrylamide was introduced as an alternative due to its superior physical properties [2325]. The stain most commonly utilized at this time for mainly qualitative in-gel protein detection was amido black. It was not until the mid 1960s that researchers prioritized the detection sensitivity of protein stains and described Coomassie Brilliant Blue (CBB) as the preferred stain for quantitation [18, 26]. CBB is thought to bind to proteins through electrostatic interactions between the sulfonic groups of the dye and the basic side groups of amino acids [18, 2729]. It has also been suggested that CBB binds to proteins through interactions with aromatic residues as well as hydrophobic interactions [27, 29].

Over the years, there have been a remarkably large number of studies dedicated to improving solvent composition, changing dye concentration/type and developing strategies for staining protein/destaining gel matrix in 1D PAGE or isoelectric focusing (IEF) gels to potentially achieve higher levels of sensitivity [3042]. This continuous drive for improvement, however, was not able to enhance the detection sensitivity of CBB below ∼30 ng of bovine serum albumin (BSA) or actin [33, 4345]. The year 1981 saw mini-gel staining with CBB R-250 detect as little as 10 ng of protein (BSA, carbonic anhydrase (CA), α–lactalbumin) via microdensitometry [46]. This method also delivered a LDR between 10 and 200 ng for the aforementioned proteins. Inter-protein variability was reported as a percentage of protein weight successfully stained and was determined to be acceptable (percentages ranged from 11.5% to 26.7% with 5.4% standard deviation). Perhaps the most significant contribution to CBB staining was made by Neuhoff et al. [47] when the colloidal state of CBB (in particular the G form) was utilized to improve detection sensitivity from 10–30 ng to 1 ng of BSA. It was also claimed as little as 0.1 ng (BSA) could be detected [47]. This move from the traditional CBB formulation in organic solvents which produced high background staining to an alternate formulation marked a significant advance in CBB staining. The use of colloidal CBB permitted free coomassie molecules to penetrate the gel and bind to protein while the remainder was in large colloidal particles that were excluded from the gel [47, 48]. Additional advantages of this colloidal CBB (cCBB) formulation included reduced background staining of polyacrylamide gels and a simplified, lower cost procedure. Comparative evaluation of the available literature concerning CBB sensitivity yielded little published quantitative information that fulfilled the criteria for analysis in this review. Although, single aspects of CBB sensitivity have been previously addressed in specific studies, recent work clearly indicates that the sensitivity related characteristics of a stain, such as LLD and LDR, must both be analyzed to ensure confidence in the capacity of a stain to quantitatively represent the amount of protein in a given band or spot.

A well-designed study utilized a commercial formulation of cCBB (Pierce Chemical Company) and determined the LLD to be between 8 and 16 ng for broad-range molecular weight (MW) standards (Bio-Rad; nine proteins, 6.5–200 kDa) and a LDR for all proteins between 30 and 250 ng with high correlation (R = 0.9883) [20]. It was also shown that cCBB staining for four standards, in comparison to the fluorescent stain SR, produced low inter-protein variation. Furthermore, similar peptide mass profiles from matrix assisted laser desorption ionization–time of flight–mass spectrometry (MALDI–ToF–MS) were obtained following the use of either stain [20]. In comparison, the use of the Neuhoff CBB formulation [49] yielded a LLD of 4–8 ng protein (five standard proteins), LDR between 30 and 500 ng for BSA (R = 0.985), 8–500 ng for phosphorylase b (PhosB), ovalbumin (OVA) and peroxiredoxin (R = 0.987, 0.992, 0.983, respectively) and 15–500 ng for CA (R = 0.981); and there was similar MALDI–ToF–MS sequence coverage for all proteins and protein loads (4–125 ng/band) [49]. However, it was noted that sequence coverage was detrimentally affected when CBB stained gels were not destained.

While the staining characteristics of standard proteins are informative, the real application of a stain is the sensitive detection of proteins obtained from complex native samples. The Neuhoff formulation of cCBB was applied to a 2D separation of a total soluble protein extract derived from Arabidopsis thaliana to determine detection sensitivity [17]. Here, it was shown that the LLD for the total protein extract, consisting of a complex mixture of proteins with varying mobilities and abundance, was 2 μg of the protein loaded. The LDR for two proteins, one of high and the other of low abundance with relatively similar mobility, was linear (15.6 ng–4 μg) with a correlation value of 0.99. Resolution of this native extract by 2DE resulted in 250 detectable spots, the lowest total spot number obtained in comparison to all stains tested in the study [17].

Notably, it has recently been demonstrated that the imaging mode of cCBB stained gels may contribute to reduced sensitivity. Analysis of low MW protein standards (GE Healthcare) revealed that using the scanning transparency mode (as opposed to the standard reflection mode) for imaging cCBB stained gels improved the sensitivity of low MW proteins by as much as eight- to 29-fold [50]. This study reported a LDR from the individual protein LLD up to 1,000 ng, with no observed change in MS compatibility; the sequence coverage for all analyzed protein groups (three groups based on molecular weight, for a total of 24 proteins) ranged between 47% and 60% [50]. Therefore, it may be useful to explore alternative imaging modes before the densitometric use of CBB is completely dismissed as a tool for protein quantification.

As noted above, although several other published improvements to CBB staining protocols reportedly also deliver increased sensitivity, the assessments carried out in those studies unfortunately do not allow for their evaluation in this review [21, 5155]. Similarly, alternate staining strategies have also been developed to improve protein detection sensitivity by combining CBB with other stains. Limited quantitative characterization, and in some cases failure to offer a quantitative advantage have limited their use in proteomics [45, 56, 57]. Commercialization of CBB has become widespread and the stain is available as a stable and ready to use product at low cost. Most commercial CBB stains, usually the G form, are marketed for densitometric detection. However, CBB may be re-visited as a sensitive stain once again since recent literature has indicated that near-infrared fluorescence detection of proteins by CBB offers improved sensitivity [21, 54]. Even though the densitometric use of CBB is relatively insensitive, use of this stain may be reborn in proteomics for fluorescent protein detection.

Silver stain

The search for a new stain was warranted by the principle limitation to CBB staining: its apparently inadequate sensitivity for protein detection. In 1979, a silver stain procedure for in-gel protein detection was suggested to be 100-fold more sensitive than CBB [19]. This marked improvement in detection sensitivity heralded a period of studies aimed at perfecting the original silver stain method and/or developing alternate silver stains [5871]. Even though there are numerous methods for silver staining described in the literature, most comprise the same main steps—fixation, pretreatment/sensitization, impregnation (saturating the gel with silver ions), development (change in pH reduces silver ions to metallic silver) and cessation of development. In 1981, a rapid silver stain method was reported that still delivered reproducibility and high sensitivity, yielding a LLD of 0.1–0.2 ng and LDR between 1 and 30 ng (for three standard proteins) [46]. The authors described inter-protein variability in terms of percentage weight of proteins in the sample that were successfully stained. The stained fraction ranged from 5.2% to 28.2% (16.7 ± 9.7%; mean ± SD), suggesting that proteins such as PhosB, BSA, and OVA were not stained uniformly by this silver staining protocol. A different silver stain method was utilized to optimize quantitative detection of BSA, OVA, CA, soybean trypsin inhibitor, and α-lactalbumin detection in 2D gels [72]. While the rapid method above clearly indicated the temperatures at which staining should be undertaken [46], this second protocol was less attentive to temperature [72], focused on stain components and duration of exposure as well as employing a twofold stronger developer to deliver a LLD of 27 pg/mm3. The LDR values were markedly different for each of the proteins tested and ranged from the reported LLD up to 5–50 ± 16.3 ng/mm3 [72]. While numerous similar studies for the optimization of silver staining have been published, most do not meet the quantitative requirements for this review.

In 2000, it was shown that the acidic silver nitrate stain (Investigator Silver Stain Kit, Genomic Solutions) was more sensitive than CBB (and cCBB), detecting as little as 2–4 ng protein, but was limited by a narrow LDR (4–60 ng protein or 1 order of magnitude) [20]. This silver stain was susceptible to high inter-protein variability and could only be compatible with MS when glutaraldehyde was omitted from the process. Application of this silver stain kit (without glutaraldehyde fixation) to 2DE analysis of rat fibroblast whole soluble cell lysate revealed a small dynamic range—231.1 pixels within the spot perimeter (calculated as the difference between the most intense and the least intense matched spot) and a slightly lower total spot number (82%) relative to staining with SR [73]. While comparable between silver and SR stained protein at higher loads, MS coverage for two standard proteins (PhosB and OVA) was decreased at lower protein loads for silver stain (without glutaraldehyde). A silver nitrate staining method developed by Mortz et al. [74] was applied to A. thaliana total soluble protein extract separated by 1D and 2D PAGE [17]. The LLD was determined to be 1 ng (i.e., more sensitive than cCBB), but showed lower correlation for two unspecified protein bands arbitrarily selected by the authors (band 1, R = 0.82; band 2, R = 0.88). Also, total spot detection in 2D gels using silver nitrate was superior to that of cCBB but not equivalent to SR (cCBB ∼300; Silver ∼600; SR ∼800). Since the introduction of silver stain, there have been numerous attempts to enhance staining intensity. However, the addition of blue toning (i.e., incubation of a silver stained gel in a solution containing ferric chloride/potassium hexacyanoferrate III/oxalic acid which turns protein bands blue), Stains All (SA) or CBB staining prior to or following silver staining has demonstrated no substantive improvement in sensitivity of detection and/or quantitation [7582]. Thus, while it is generally accepted that silver staining provides slightly better sensitivity than CBB, one of its major limitations is its incompatibility with MS [83]. This incompatibility is believed to be due to the strong chemical reagents employed in the silver staining process which results in chemical modification or destruction of proteins. Chemical modification can involve glutaraldehyde cross-linking with proteins and blockade of trypsin digestion. This reduces the number of available peptides, sequence coverage and thus quality of protein identification [67, 84, 85].

MS compatibility is improved when glutaraldehyde/formaldehyde is omitted and/or ammoniacal silver staining is used, but this compromises detection sensitivity [68, 86, 87]. Due to this less-than-ideal compromise, there have been attempts to sustain sensitivity and increase MS compatibility by other means. It was shown that calconcarboxylic acid introduced as a silver ion sensitizer produced better sensitivity over the traditional silver stain method (0.05–0.2 ng/band) [88]. The LDR for standard proteins (SDS–6H, a mixture of the six standard proteins BSA, CA, PhosB, β-galactosidase, albumin, myosin; Sigma) covered a variety of ranges all with a narrow spread of linearity and exhibited high inter-protein variation (based on visual examination of the LDR plots shown) [88]. Application of this alternate silver stain to Escherichia coli BL-21 total soluble protein extract yielded a 2D protein map with a greater number of spots than seen with the traditional silver stain (as indicated by authors, and visual comparison of the images shown). Although not yet demonstrated, the assumption of MS compatibility is based on the fact that the stain does not appear to covalently modify protein. Another protocol utilizes erichrome black T (EBT) as the silver ion sensitizer [89]. Here, the LLD of EBT–Ag was 0.05–0.2 ng/band (SDS-6H; Sigma) and the stain was also capable of qualitatively detecting lower loads of total soluble protein extracted from E. coli in comparison to silver staining without glutaraldehyde. 2DE analysis of E. coli total soluble protein also supported the superior performance of EBT-Ag over traditional silver staining with 16% more detected spots [89]. The staining of standard proteins demonstrated that MS compatibility and sequence coverage were relatively similar for each protein tested between loads of 6–100 ng/band. Some proteins were also identified from as little as 3 ng/band [89]. Another MS compatible silver stain made use of the counter–ion dyes, ethyl violet and zincon (EZ). The LLD of standard proteins was 0.2 ng/band (SDS–6H; Sigma), the LDR was between 4 and 50 ng, and sequence coverage was relatively similar for each protein tested between loads of 6–100 ng/band and occasionally at 3 ng/band [90].

Although it is clear that silver staining can be altered to achieve specific requirements and high-detection sensitivity, there are various drawbacks to the use of this method. Silver staining requires various reagents to be prepared fresh with high quality water, can be laborious and tedious, tends to produce large inter-gel variation in intensities as it is without a staining endpoint and exhibits poor linear dynamic responses. Other potential drawbacks to silver staining that may also interfere with qualitative and quantitative analyses arise due to the fact that silver stain does not specifically stain proteins—it also detects nucleic acids and lipopolysaccharides [66, 9194] and is not sensitive for detection of all types of proteins [95]. Thus, a range of potential problems need to be carefully considered when choosing silver stain for in-gel protein detection and quantitation.

Negative stains

While various approaches to negative staining of proteins in-gel had previously been developed [96, 97], it was not until the introduction of heavy metal salts that wider interest in the application was piqued. Copper chloride was initially utilized for negative staining of protein bands [98]. Although copper chloride was found to be useful for negative staining, it was later shown that zinc chloride detected proteins with even greater sensitivity [99]. Alternative zinc chloride staining could also be achieved by exploiting the pH dependence of zinc chloride complex precipitation [100]. However, zinc staining by these methods was not homogeneous. The introduction of imidazole for zinc staining overcame this problem in SDS-PAGE applications [101] and was further modified for application to non-SDS gels [102]. These marked improvements to zinc staining resulted in a large number of studies aiming to further enhance the method and demonstrate that proteins within zinc stained gels were still biologically active, could be recovered with high yields and were compatible with MS technologies [103110]. Since this process leaves proteins presumably untouched, significant effort was invested in highlighting its qualitative potential and studies examining its quantitative capacity have been less prevalent. The quantification of protein using zinc staining is controversial in itself as the protein is not stained and protein concentration can only be quantified by pixel inversion. For this reason, few studies have characterized the potential for zinc staining as a quantitative in-gel protein detection method.

The zinc staining procedure developed by Fernandez-Patron et al. [111] showed the LLD of broad-range MW standards (Bio-Rad) to be between 1 and 2 ng protein, similar to that of SR [20]. Although the LDR ranged from 250 to 2,000 ng (standard proteins; R = 0.9843) it was obvious that the main limitation of zinc staining was the detection of low abundance proteins. Using a rapid zinc stain kit (Visual Protein, Taipei, Taiwan) it was shown that the LLD of low MW standards (GE Healthcare) ranged between 1.8 and 4.0 ng/band [50]. Although the use of scanning mode, rather than reflective mode, for imaging did not improve detection sensitivity it did, however, provide images with greater contrast. The LDR for this zinc procedure was very narrow (no greater than 140 ng) and may have been influenced by the rapid nature of the procedure used, as described by the manufacturer. MS analysis of soluble proteins from human hepatoma cell (hepG2) revealed that for 24 selected proteins (divided into three groups based on MW: (1) >45 kDa, (2) 30–45 kDa, and (3) <30 kDa), sequence coverage was roughly uniform (51.3–56.4%) [50].

The most significant disadvantage of negative staining is that protein quantification is only approximate given the nature of the staining. Also, zinc staining is not specific to protein and can also effectively detect nucleic acids and lipopolysaccharides [112, 113]. Since this stain has no definitive endpoint, its use also involves the risk of overdevelopment. Given these limitations and the sparse literature investigating the quantitative capacity of zinc staining, it seems questionable whether the application of this method will make a significant contribution to quantitative proteomics.

Reactive densitometric dyes

The application of Uniblue A capitalized on the staining capacity of amines. When reacted pre-electrophoretically at a ratio of 3 mg dye for every 10 mg protein (40°C for 3 h–pH 10.5), the 600 nm densitometrically measured LLD was reported to be 0.5 μg of protein [114]. While lysozyme staining was not consistent with the other standard proteins tested, the relationship between quantity and stain intensity for the majority remained linear from 0.5 to 25 μg (LDR) [114]. A high pH was chosen to accelerate the pre-labeling reaction; however, this may not be optimal in some cases. The authors advised that the reaction was feasible at lower pH levels and that this may be preferable if a universal approach to staining proteins of different pH lability was required. Another moiety specific stain, 2,2′-dihydroxy-6,6′-dinaphthyl disulfide in combination with fast black K, was used to densitometrically detect sulfhydryl groups. An LLD of 0.25 μg of lysozyme and 1 μg for all other tested proteins was reported [115]. However, since the sensitivity of this dye combination depends on sulfhydryl content, inter-protein variability is to be expected.

Additional densitometric stains

While CBB (colloidal or traditional) and silver staining have remained the strongest contenders for densitometrically detecting proteins, alternatives have been explored. However, the only other densitometric stain that has been quantitatively assessed and adheres to the criteria of this review, is based on the counter-ion dye couple of EZ. This EZ stain was first introduced for in-gel protein staining in 2002 and was slightly less sensitive than cCBB [116]. Two years later, the same research group used longer fixing times and EtOH instead of MeOH as the solvent to successfully demonstrate EZ staining comparable to cCBB [49]. For five standard proteins (PhosB, BSA, chicken OVA, bovine CA, and human peroxiredoxin I) the LLD was between 4 and 8 ng protein. The LDR for PhosB, OVA, and peroxiredoxin spanned between 8 and 1,000 ng (R = 0.997, R = 0.996, and R = 0.987, respectively); 4–1,000 ng for CA (R = 0.991) and 8–500 ng for BSA (R = 0.983) [49]. It was also shown that MS compatibility of this stain was comparable to cCBB for all loads between 4 and 125 ng/band [49]. Alternate densitometric stains have not gained popularity within the proteomic research community because these stains have been unable to compete with the performance of fluorescent staining and imaging. Also, most of these alternate densitometric stains have, at best, had limited characterization or have similar detection sensitivity to cCBB and hence have not been pursued for further application in proteomics [54, 117126].

Fluorescent dyes

Unlike densitometric stains which absorb light, fluorescent stains are detected by the light they emit. Emission is a result of excitation with a particular wavelength of light which elicits an energy shift in the fluorophore. When the fluorophore returns to its ground state, this energy can be emitted as a measurable photon, thereby enabling detection of protein when it is associated with the fluorophore [127]. New fluorescent stains have seen increasingly widespread use as they address some of the limitations of densitometric stains: they are sensitive, are measured using light emission rather than absorbance, have a broad LDR, produce minimal background interference and are compatible with MS analysis.

Reactive fluorescent dyes

The reactive dyes used in proteomics, so called as their labeling of proteins involves a chemical reaction, form permanent covalent bonds with proteins. Typically, this reaction targets specific moieties on an amino acid residue such as an amine, thiol or carboxyl side chain. In the presence of surplus dye (and/or permissive pH conditions), this reaction can become non-specific and label any susceptible amino acid. The majority of approaches using reactive dyes implement pre-labeling, whereby the reactive dye is attached to proteins in an extract prior to their resolution by electrophoresis. Some amine reactive dyes that have been tested include dansyl and dabsyl chloride and remazol; although these reactive dyes have been utilized for pre-labeling, their LLD, LDR, and degree of inter-protein variability have not yet been reported in great detail [128130].

Amine groups are a valid target choice for pre-labeling as they are present on almost every protein in the form of N-terminal, α- and ε amines. They are highly reactive and produce a strong amide bond [131]. In particular, labeling lysine (lys) residues facilitates near complete coverage of a proteome given the prevalence of lys in proteins [132]. This may, however, affect the efficiency of subsequent trypsin treatment if the reactive dye masks the lys residues [132, 133]; nonetheless, there are a range of alternative reagents available for the controlled digestion of proteins to defined peptides, as is required prior to MS analysis [134136].

Two-methoxy-2,4–diphenyl–3(2H)–furanone (MDPF) is a fluorescent alternative that also reacts with amines. When used to pre–label proteins, MDPF has been shown to yield a LDR between 1 and 500 ng for CA, myoglobin, catalase, and BSA [137]. An LLD of 1 ng was reported for this reactive dye and non-uniform staining for different proteins was identified [137]. For labeling, 20 μg of protein dissolved in 0.01 M borate buffer (pH 9.5) was mixed with 60 μg of MDPF in dimethyl sulfoxide (DMSO). In an independent study, MDPF was also applied to IEF, 1D and 2D post-electrophoretic gel staining [138]. For post-IEF staining, gels were shaken with 0.02 moles of MDPF in MeOH/0.2 M sodium borate buffer (pH 9.5) and washed in MeOH and water prior to second-dimension separation [138]. For post-SDS-PAGE staining proteins were fixed and stained with 3.8 mM MDPF for 1D gels and 0.95 mM MDPF with longer staining time for 2D gels [138]. An LDR between 50 and 300 ng (soluble human lymphoid cell line IM9 protein) was identified but this is clearly only a portion of that determined using the pre-electrophoretic MDPF method, and as such does not indicate a quantitative advantage over pre-labeling [138].

The quantitative potential of o-phthalaldehyde (OPA), a compound that reacts with primary amines, has also been reported [139]. Protein concentrations ranging from 0.1 to 50 μg/ml in 0.05 M sodium phosphate (pH 8.5) were dosed with 0.19 μmol OPA (in MeOH) and incubated in the dark. Also, the addition of β-mercaptoethanol prior to the labeling treatment was used to increase the fluorescent signal sevenfold; the reason for this was not clearly established but the authors suggested some augmentation of –SH group reactivity by β-mercaptoethanol as the cause. When used to detect transferrin, the LLD was ∼10 ng while the LDR was optimal for higher loads of protein, 0.1–50 μg/ml. These values were not consistent for all proteins tested, and the authors acknowledged that inter-protein variability was an impacting factor for detection [139].

Similar limits of sensitivity were achieved by pre-labeling with fluorescamine. Protein at concentrations up to 1.0 mg/ml in 0.04 M borate buffer (pH 9.0), were labeled with 0.5 mg of fluorescamine (in DMSO), however, no information regarding the conditions of incubation were provided. This amine reactive fluorophore could detect a minimum of 6 ng of myoglobin; and the LDR for myoglobin, chymotrypsinogen and OVA were 0.5–7, 0.5–9, and 0.5–12.5 μg, respectively [140]. Again, this stain was unable to deliver a uniform interaction with all protein standards as shown by the different LDR plots for the tested proteins [140]. Furthermore, it would be impossible to reproduce these results without details concerning the incubation conditions used. Fluorescamine has also been tested as a post-electrophoretic stain of myoglobin; while this application showed potential, only the LDR (1–7 μg), was reported [141].

The staining performance of the aforementioned reactive dyes applied prior to electrophoresis, however, was not optimal in terms of sensitivity, LDR, and inter-protein variability. A significant contribution to pre-labeling methods was achieved through the introduction of the mass and charge matched lys-targeting CyDyes. These dyes are used to label only a minimal number of lys residues (1–3%) at the ε-amine of each protein [132]. A proportional representation of proteins in a sample is attempted by maintaining low dye/protein ratios. DIGE is the primary technique that applies these dyes and involves tagging proteins pre-electrophoretically with fluorescently distinct labels known as cyanine 2 (Cy2), cyanine 3 (Cy3), and cyanine 5 (Cy5) [10, 142]. The use of these three spectrally distinct dyes allows for multiple samples to be resolved and then individually imaged on a single gel, although at substantially reduced total protein loads per sample. Alexa Fluor dyes 555 and 647 have been shown to be spectrally similar to Cy3 and Cy5, respectively, with an added advantage over the Cy dyes since they exhibit reduced photobleaching and self-quenching especially with more extensive labeling [143]. The application of these Alexa Fluor dyes for pre-labeling, however, has not been extensively tested and it remains to be seen whether they can deliver similar sensitivity.

Proteins to be labeled with CyDyes were prepared in a buffer containing 4–7 M urea/2 M thiourea/2% CHAPS (pH 8.5). Additional components include 30 mM Tris–HCl [10], 2% SB3–10 and 0.5% Triton X–100 [142]. Proteins were pre-labeled on ice for 30 min as recommended by GE Healthcare, the firm predominantly involved in marketing the DIGE technique [10, 142]. The fluorophores are used at doses between 200 and 400 pmol/50 μg of total protein and the reaction is incubated in the dark. Low rates of labeling were required to prevent reduced sample solubility and protein spot chains on gels that result from the labeling of multiple lys residues on any given protein. Pre-labeling is said to reduce inter-gel variability and improve protein detection with a LLD as low as 0.25 ng and a LDR of up to three to four orders of magnitude (determined using protein standards supplied by Sigma) [142]. Gels can also be imaged immediately following electrophoresis. While lys residues are almost ubiquitous in proteins, inter-protein variability can be expected as proteins do not have uniform amino acid content and lower abundance proteins are less likely to be labeled [10, 132]. In addition, it has also been reported that gel-to-gel variation still contributes most of the inherent variability to DIGE [144]. It is also important to consider the ramifications of selectively labeling a sample and subsequently loading only a fraction of this sample for electrophoresis; the total complement of detectable protein is thus reduced by the very design of the protocol used for detection.

When used to pre-label soluble mouse liver homogenate for 2DE separation, Cy2 detected 414 ± 0.21 spots, Cy3 detected 289 ± 1.09 spots, and Cy5 detected 398 ± 0.81 spots [142]. Pre-labeling soluble proteins from Pirellula sp. Strain 1 for 2DE separation as described by the manufacturer (GE Healthcare), showed that Cy2 detected 399 spots, Cy3 387 spots and Cy5 418 spots [10]. In comparison, SR was found to detect more spots (443, no error provided). Unfortunately, the pre-labeled proteins themselves cannot be identified using MS techniques. However, as only a fraction of the total amount of protein per sample is thought to be labeled, the unlabeled protein is, in theory, available for MS identification. These ‘unlabeled’ spots can be picked by estimating the shift in gel mobility that dye labeling causes and then selecting the unlabeled gel spot that is thought to correspond to the labeled fraction [10, 145]. It is, however, unclear how this estimation is done without knowing the number of potentially labeled lys residues in any given spot representing an unknown protein. Both studies reported that the CyDye pre-labeling approach is compatible with MS identification; however, the Pirellula sp. strain study did not present any data to support this claim [10]. The mouse liver homogenate protein spot identities were assigned using MALDI–ToF–MS and while sequence coverage was not quoted, the authors described their process for validating MS data and cross reference the calculated characteristics of candidate proteins with reported pIs and MW [142].

Another approach to reactive labeling is to target the thiol groups presented by cysteine (cys) residues. While cys is not as abundant or widespread in proteins as lys [132], these thiol groups are readily reactive and thus effective targets for labeling. For this reason, cys residues can also be labeled to saturation without the substantial loss of protein solubility encountered when lys residues are targeted [146]. Cys residues are also less likely to be at trypsin cleavage sites. As such, proteins labeled at cys residues can still be identified by MS following standard digestion protocols [133, 146, 147]. Focusing on thiols also allows the pIs of labeled proteins to be maintained since the dyes are neutrally charged. Recently, it has been demonstrated that saturation labeling of thiols can also be successfully carried out using the BODIPY dyes (FL–N–(2–aminoethyl) maleimide; FL C1–idoacetamide) by optimizing the labeling reaction conditions [148]. The BODIPY dyes demonstrate a LLD of 10 fmol (without a reported error) and a LDR of three orders of magnitude for yeast enolase I. Inter-protein variation was not referred to but can be inferred from the thiol specific nature of these dyes [148].

A recent study using monobromobimane (mBB) to label cysteine residues in proteins reported a LDR of 32–1,000 ng [20]. The LLD of this stain was between 4 and 32 ng (broad-range MW standards) and when compared to silver, cCBB and SR staining, mBB demonstrated the greatest inter-protein variability [20]. Also, mBB did not demonstrate strong MS compatibility since both BSA and soybean trypsin inhibitor could not be correctly identified by MS when labeled to saturation [20]. Proteins were pre-labeled at a concentration of 2 mg/ml and following denaturation, samples were cooled before the addition of mBB (6 mM final concentration). Samples were incubated in the dark, before excess cys was used to quench the reaction. Another older study used recombinant proteins to examine mBB quantification of proteins with published cys content [149]. The average LLD of protein was 86.2 ± 14 ng with an outlier of 212 ng for p21, a value much higher and thus representing lower sensitivity than the more recent study. Additionally, LDR maxima themselves ranged from 110 ng (recombinant protein p49) to 8.47 mg (recombinant protein p21); however, these are the extremes of detection across all of the tested recombinant proteins. LLD and LDR were also expressed in terms of cys content; mBB was capable of detecting as little as 6.3 ± 1.1 pmol of cys per band and delivered a linear response between 25 and 400 pmol [149]. A possible explanation for the increased sensitivity observed in the results of the more recent study was improved methods of detection. The advent of the cooled charge-coupled device (CCD) camera may have been a definitive factor in the difference in sensitivity observed between the two studies. Although mBB can be used to pre-label protein prior to IEF for 2DE, like other reactive dyes, it only detects a specific reactive group, so those proteins lacking thiols go undetected. Pre-labeling may also affect the mobility of some low MW proteins [150].

In 2003, a saturation approach to labeling protein cys residues with the commercially available CyDye maleimides (Amersham Biosciences/Invitrogen) reported a LLD of 0.1–5 ng of protein per band and LDR of three to four orders of magnitude [132]. Proteins were dissolved in lysis buffer, pH adjusted to 7.5, and reduced with Tris (2–carboxyethyl) phosphine hydrochloride (TCEP). Fluorophore was then added at a ratio 20 nmol/50 μg of protein before incubation, and quenched with 2 × 2D sample buffer prior to analysis by 2D gel electrophoresis. The main difference between minimal Labeling and this protocol is the final ratio of dye to protein. labeling of cys residues to saturation allows for the use of much smaller masses of tissue and is thus useful when working with human surgical, biopsy, post-mortem, and other limited tissue sources [151]. Application of saturation labeling to samples from bacteria, mouse liver, cancer cells, and feline brains have also been successful. Maleimide CyDyes have also been used to examine the 2D separation of pancreatic intra-epithelial neoplasia cell proteins [152]. This method was consistent with the protocol mentioned previously; however, the proteins were prepared from laser micro-dissected tissue and were quantified in terms of cell number [132]. The 1,000 cells used for optimal labeling yielded 2.3 μg of total protein; this was then labeled with 4 nmol of fluorophore, resulting in a much higher fluorophore to protein ratio than used previously [132]. This study detected ∼2,500 spots using micro-dissected cells and as a test-of-concept MALDI–ToF–MS was used to identify trypsin cleaved γ-actin with 38.7% sequence coverage [152].

Additional maleimides include the newly developed DY-680 and DY-780 dyes which are infrared thiol-reactive dyes; these were used to compare newborn and adult murine brains [153]. Proteins (2.5 mg) were solubilized, reduced and desalted before being treated with 200 μg of DY-680 or DY-780 in DMF. While this labeling procedure reduces the initial protein pool during the desalting and concentration steps, DY-680 was shown to be highly sensitive with a LLD of 10 fg (labeled tubulin, with no reported standard deviation) [153]. Once again, however, proteins without accessible cys residues were not detected by these maleimides, indicating a level of inter-protein variability that will be sample dependent. For 2D analysis of soluble mouse brain proteins, the LLD was 5 μg total protein (labeled and unlabeled). Although this labeling procedure has great potential sensitivity, the interference of these dyes with subsequent MS analysis will limit their implementation. This interference was described as a decrease in the MALDI–ToF–MS MASCOT scores of proteins labeled with DY-dyes compared with those stained with cCBB. Recently, the application of other fluorescently distinct unpatented DY-maleimides in protein staining has also been explored [154]. Pre-labeling for both 1D and 2D gels began with 20 μg total protein (human serum albumin (HSA) or Human keratinocytes for 2D) in 30 mM Tris–HCl (pH 7.5). Samples were reduced with TCEP before being treated with 8 nmol of a DY-maleimide dye and pre-labeled in the dark. While the 1D sample reaction was quenched with sample buffer in preparation for loading, 2D samples were quenched with stop buffer [154]. Using this method, DY-maleimides 505 and 635 made the detection of 0.13–1 ng HSA possible, and yielded a LDR of three orders of magnitude. 2D application of DY–maleimides 505 and 635 identified 1,212 ± 124 and 1,050 ± 28 protein spots, respectively, on 200 × 250 × 1.5 mm gels [154]. A subset of these spots was also submitted to MS analysis resulting in 22–60% sequence coverage for 17 DY-maleimide-labeled proteins.

Iodoacetylated cyanine dyes (ICy3 and ICy5) are a variation of the commercially available maleimides that were originally synthesized non-commercially [147]. The only reported LLD value was for BSA, a protein with 35 cys residues (5.8% of the total amino acid complement), at 2 ng [20]. While a LDR of three to four orders of magnitude is comparable with the current staining benchmark, SR, this range was only applicable to thiol containing proteins. An additional indication of inter–protein variability includes the 1–8% of proteins that are preferentially bound by one of the two iodoacetylated dyes [147]. Standard proteins, including BSA, chicken OVA and equine myoglobin, at 2 pmol final concentration were dosed with an ICy fluorophore. Soluble proteins extracted directly from cells (HMLEC line HB4a and its ErbB2-overexpressing derivative C3.6) were labeled concurrently with lysis to limit thiol modification during/after cellular disruption. Iodoacetylated fluorophores were used at 80 pmol/μg of total protein and reactions were incubated on ice for 60 min before being quenched with DTT [147]. The main goal of this method was to achieve saturation labeling of all thiol groups and this was reflected in the 20-fold increase in fluorophore concentration used by the authors compared with succinimidyl ester cyanine dyes discussed previously. These dyes were used in a comparative study of the detection delivered by ‘traditional’ and more contemporary stains in 2DE. When compared with lysine targeting CyDyes, silver and cCCB staining, ICy dyes detected 1,034 ± 245 spots, 85% of the total number of spots detected on the same gel with MS compatible silver stain. To further characterize the iodoacetylated cyanine dyes, soluble proteins extracted from mammary luminal epithelial samples were pre-labeled and the detected protein spots were tested for MS compatibility. Of these spots, 89 were submitted for identification by MALDI–ToF–MS and 51 proteins were identified. Identity assignment was made with a range of 25–68% sequence coverage [146, 147]. The IC–OSu ethyl–Cy3 and –Cy5 N–hydroxysuccinimide ester cyanine dyes (IC3 and IC5; Dojindo Laboratories) were also recently developed [155]. These dyes were directly compared with the commonly used CyDyes (GE Healthcare) by identifying protein spots displaying two- or five-fold differences in volume ratio and dividing this number by the total number of identified spots to give a percentage of similarity. Based on this, the authors judged the two labeling strategies to be equal in suitability for protein detection, multiplexing, and proteome quantitation [155].

Non-reactive fluorescent dyes

Post-electrophoretic staining with fluorescent non-covalent dyes is the most widely applied technique in proteomics for in-gel protein detection. These stains are not reactive and can commonly be removed from the gel and resolved proteins through extensive washing. SR is the most commonly employed fluorescent stain; not only is its use as simple as CBB staining, it is also reported to have a high level of sensitivity and wide LDR [20]. The manufacturer’s staining procedures used for 1D and 2D gels differ slightly. 2D gels are fixed in MeOH/HAc before SR staining whereas 1D gels are immersed in stain immediately following electrophoresis. A LLD between 1 and 2 ng (broad-range MW standards; Bio-Rad) was identified and quantification was linear from the LLD up to 1,000 ng of protein [20]. In comparison to the other stains tested (i.e., variations of silver, zinc/imidazole and CBB), SR was reported to have the lowest level of inter–protein variability but no quantitative measure was provided for this claim. SR also exhibited MS compatibility with sequence coverage greater than 36% for all proteins tested.

Recombinant proteins (rhuMAb, tPA, hGH) were serially diluted to characterize the staining potential of SR [156]. Following electrophoresis of these proteins, 1D gels were fixed in a solution of MeOH/HAc as were 2D gels. 2D gels were then incubated in stain overnight whereas 1D gels were stained for 3-4 h. 1D gel destaining was conducted using MeOH/HAc and a less concentrated solution was used for 2D gels. Subsequently, both gel types were rinsed with water. This report noted a LLD value between 0.5 and 5 ng of recombinant protein and the LDR of each of these three proteins spanned from 200- to 1,000-fold. This indicated a degree of inter-protein variation not suggested in the study mentioned previously [20] and implied that linearity is strongly dependent on the type of protein. 2DE analysis of a soluble E. coli protein fraction yielded an LDR of about a 1,000-fold [156]. This was determined using Progenesis or PD Quest generated histograms and plotting protein amount versus staining intensity.

Independent studies using the SR staining method detailed previously [20] were also carried out [157, 158]. It was reported that SR again yielded a LLD of 1–2 ng (SDS-6H marker proteins; Sigma) and a LDR of 2–500 ng [158]; little inter-protein variation was suggested by the LDR plot shown. Also, MALDI–ToF–MS analysis of standard proteins (BSA and CA) displayed sequence coverage between 29–34% and 48–58%, respectively (for protein loads between 4 and 64 ng) [157]. A comparison of this and the previous studies suggests that the LLD values are consistent. However, despite using similar protein standards, the LDRs reported in these studies show a twofold difference, the source of which may be the limited experimental range of protein concentrations tested by Cong et al. [158].

Using the manufacturer’s SR protocol, another study analyzed a total soluble protein extract from A. thaliana; the LLD was determined to be 500 ng and a linear relationship was revealed for both high and low abundance proteins (R = 0.96 and 0.97, respectively) [17]. In this study, SR provided the best overall protein spot detection in comparison to all other tested stains (i.e., DP, cCBB, silver nitrate and C16-F) and the highest staining reproducibility across triplicate 2D gels [17]. The significant potential of SR for 2D analysis was further highlighted using rat fibroblast-soluble lysate (50 μg protein per gel) [73]. SR detected 1,290 ± 34 spots, the greatest number in comparison to all other tested stains (including silver stain and cCBB) and was shown to have a dynamic range 680-fold greater than cCBB detected densitometrically, demonstrating superior capacity for protein detection. MALDI–ToF–MS analysis of PhosB and OVA (1D–PAGE) reinforced the advantages of using SR since reasonable sequence coverage was obtained even from loads below 9.3 ng/band [73]. 2D separation of pre-reduced standard proteins (Bio-Rad 161–0320) stained with SR according to the manufacturer’s staining instructions revealed the presence of 34 spots [detected by eye from an image captured using a 532 nm excitation and a 610 ± 30 nm band-pass emission filter (Typhoon 9200; GE Healthcare)] [159]. Among these, soybean trypsin inhibitor and myoglobin were only minimally detected due to an apparent ineffective staining capacity of these proteins. While this indicates a level of inter-protein variability, the majority of resolved spots were identifiable by MALDI–ToF–MS with minimum sequence coverage of 40% [159]. It has also been suggested that to overcome the high cost of SR for proteomic application, dilution or re-use of the stain could be considered; however, such use of SR was not optimal [160, 161]. Reported LLD values for BSA were detrimentally affected by SR re-use or dilution with water. Either treatment produced a twofold reduction in LLD from 1–2 to 2–4 ng. LDR values were also affected by up to sixfold with any attempt at economizing SR use [161].

Since the introduction of SR at its relatively high cost, there has been interest in developing equivalently performing yet inexpensive fluorescent ruthenium based alternatives. Ruthenium (II) tris (bathophenanthroline disulfonate). RuBPS (also known as RuBPSA and RuBTS) was introduced as an economical alternative to SR in 2000 [162]. Although it was suggested that RuBPS delivered superior sensitivity in comparison to SR [163], it was later demonstrated that there was no quantitative advantage over the original or optimized SR formulation [164]. It is, however, definitely more cost-effective and has thus been used in various proteomic investigations [165170]. It has been shown that the limit of sensitivity of RuBPS is approximately 10 ng protein/band (broad-range MW standards, Bio-Rad), as reported previously [164] and determined qualitatively from published data [163]; however, a more detailed evaluation of RuBPS staining indicated this sensitivity threshold to be much lower, at ∼2 ng protein/band (CA, soybean trypsin inhibitor, OVA, albumin, PhosB) [168]. Similarly to SR, the LDR for RuBPS was between 2 and 2,500 ng for BSA [168]. Another fluorescent ruthenium based stain, ASCO_Ru—commercially available from Sigma as bis(2,2′-bipyridine)-4′-methyl-4-carboxybipyridine-ruthenium-N-succinimidyl ester-bis(hexafluorophosphate), was explored and shown to detect and quantify as little as 80 pg of glutamate dehydrogenase [171]. It was also noted that approximately 12% more spots were detected by ASCO_Ru in comparison to SR (total soluble protein extracted from human colon carcinoma cells HCT 116) [171]. Rubeo fluorescent protein stain (G Biosceinces) was shown to not only detect fewer spots (mouse liver total soluble protein) but was subject to high inter-gel variation in comparison to SR and other fluorescent protein stains [21]. Thus, it seems that the search will continue for a fluorescent protein stain that outperforms SR, and is without its substantial expense.

Deep Purple (DP) is a fluorescent dye based upon the natural compound epicocconone, originally isolated from the fungus Epicoccum nigrum [172]. In 2003, a study comparing Lightning Fast (later renamed DP) to SR showed that the LLD of DP (∼64 pg protein/band; no error given) was eightfold superior to that of SR [173]. Additionally, 18–19% more spots in a 2DE separated sample of soluble rat microsomal proteins were detected using this alternate fluorescent stain in comparison with SR. The staining protocol consisted of agitated fixation in HAc followed by two water washes and a light protected 0.5% (v/v) DP staining step. Background was reduced by performing three short water washes [173]. Most proteins tested had a wide LDR of ∼102 pg–1,024 ng when detected with DP, and subsequent sequence analysis by MALDI–ToF–MS (>23%) was similar to that obtained following CBB or SR staining. It was also noted that the number of spots detected was greater when using a non-linear immobilized pH gradient strip regardless of the stain used [173]. Efforts to streamline the manufacturer’s protocol by consolidating washes with the pH changes necessary for effective staining were also assessed for their effect on the quantitative capacity of DP [174]. Here, they determined that pH played a critical role in DP staining and the enhanced method maintained sensitivity and LDR with fewer steps and less handling time [174]. Application of DP to a native soluble protein sample derived from A. thaliana demonstrated a LLD of 0.5 μg total loaded protein and a linear staining relationship was found for both high and low abundance proteins (R = 0.84) [17]. However, use of the manufacturer’s protocol resulted in only 75% of the spots being detected by DP in comparison to SR.

The performance of DP (a.k.a. LavaPurple), using the manufacturer’s protocol, was also assessed using 2D protein standards (Bio–Rad 161–0320); 41 protein spots were detected, comparable to the number seen using SR (38) [159]. These proteins were amenable to MALDI–TOF–MS analysis, with identifications being made with as much as 58.5% sequence coverage [159]. DP is suggested to have a slight advantage over SR when applied to peptide mass fingerprinting (PMF) as it is less likely to result in tested spot identification failure and provides consistent PMF for lower abundance spots [175]. It is also compatible with MALDI–ToF–MS and Liquid chromatography (LC)-MS methods of protein sequencing [176, 177]. One known disadvantage to DP staining is its photoinstability. After 6 min of light exposure, DP signal suffered a 50% reduction while 19 min of exposure to the same light intensity was required to produce a 44% reduction in SR signal [178]. Even with these disadvantages, DP and SR are available at similar prices. Direct comparison of DP and SR using fractions of total soluble and total membrane proteins extracted from mouse liver, revealed that the SNR of DP was closer to that of densitometrically detected CBB than SR, as well as demonstrating inferior detection of acidic, membrane, and low MW proteins [21]. Unfortunately, the separate testing of soluble and membrane protein fractions is not otherwise routinely carried out in the field. This raises the question of how many stains, largely characterized with soluble protein extracts, might be found to underperform in full proteomic analyses that naturally include membrane proteins. Additionally, the DP staining protocol requires more hands-on time than that of SR, even using the consolidated protocol [174]. Although the binding mechanism for DP with proteins has not been clearly defined it has been noted that DP does undergo a unique reversible reaction with primary amines [179].

C-16 fluorescein (C16-F) has been recently explored as an alternative to SR. 1D gels were fixed and stained in 1 μM dye dissolved in EtOH/HAc (twice), followed by two water washes and a brief rinse in HAc [180]. The same procedure was applied to 2D gels without the water washes. Using this C16-F staining procedure, the LLD identified for BSA was 0.12 ng, the LDR was 7.8–125 ng protein/band (based on four standard proteins) and some degree of inter-protein variability was indicated [180]. The staining method for C16-F [180] was also applied to a native soluble protein sample derived from A. thaliana [17]. Here, the LLD using C16-F for detection was equivalent to that of silver stain and DP but not as sensitive as SR. The linear relationship of C16-F staining for a pair of high and low abundance proteins with similar MW was equivalent to that of SR (R = 0.99). Although 2D protein detection using C16-F was poorer than with all stains tested other than cCBB, staining reproducibility across multiple gels was high.

Other fluorescent stains available in the commercial market include Krypton and Krypton IR protein stains (Pierce) and the family of LUCY dyes—LUCY® 506, LUCY® 565, LUCY® 569 (Sigma-Aldrich). The characteristics for protein sensitivity, however, have not yet been independently tested by researchers other than the manufacturer and their collaborators. Authors examining the LUCY dyes reported LLDs for BSA at 2 ng for LUCY® 506 and 5–10 ng for the remaining LUCY dyes [176]. All of the LUCY dyes stain α1-acid glycoprotein poorly, indicating some inter-protein variability.

Additional fluorescent stains

Since the introduction of fluorescent stains for in-gel protein detection, there have been various developments for alternate fluorescent detection methods. Most, however, do not reach the level of sensitivity achieved by SR, have complicated protocols or have not been extensively characterized for widespread use in proteomic applications [163, 181190]. Recently, there have been a few candidate stains that show equivalent staining characteristics to SR. In 2008, palmatine was shown to have similar detection sensitivity to cCBB [190], but was subsequently optimized to achieve sensitivity comparable to that of SR [158]. Originally, the stain was dissolved in EtOH/HAc, however, an alternate solvent containing SDS and HAc enhanced the performance of palmatine [158]. The LLD was 2 ng for all marker proteins tested (SDS-H; Sigma); since the protein concentration range tested was narrow (2–500 ng), the resultant LDRs were the same for both palmatine and SR [158]. It was also shown that for five protein standards, sequence coverage achieved by MS analysis was similar for all loads between 4 and 125 ng/band. Additionally, sequence coverage for 12 palmatine-stained spots from neuroblastoma SH-SY5Y total soluble protein extract was comparable to that obtained by SR staining (ranging from 31% to 69%). The other advantages of using this stain are the low cost, reduced labor and environmental friendliness (i.e., harmful organic reagents are not necessary).

Another possible alternative to rival SR is 4,4′-dianilino-1,1′-binaphthyl-5,5′-disulfonic acid (BisANS). Significant changes to the original staining method for BisANS have recently achieved comparable protein detection sensitivity to that seen with SR [157]. The original BisANS staining protocol according to Horowitz and Bowman [184] washed gels in water, stained with 20 μm BisANS (in water), washed in 2 M KCl, and rinsed briefly in water. The method developed by Cong et al. [157] differed significantly; gels were first fixed in EtOH/HAc, washed in water, stained with 0.0002% BisANS in EtOH/HAc and rinsed briefly with water before imaging. The LLD of standard protein markers (SDS–6H2; Sigma) was 1 ng/band and the LDR between 1 and 250 ng; both values were comparable to those obtained using SR [157]. MS analysis (MALDI–ToF) showed similar sequence coverage for both BisANS and SR for two standard proteins. Although palmatine and BisANS are simple and inexpensive to use (almost 100 times cheaper than SR), more research will be required to further develop the full extent of their detection capacity relative to that of SR.

Unlike exogenous stains, native fluorescence relies on the amino acid composition of proteins to facilitate in-gel detection. The ultraviolet (UV)-induced fluorescence of amino acid residues, tryptophan (Trp) and tyrosine, is the basis of native fluorescence. Since the introduction of native protein detection in gels, there have been various attempts to capitalize on and improve this process [185, 191193]. Native fluorescence has demonstrated protein detection comparable to that achieved with silver staining [194]. The LLD of PhosB, CA and glyceraldehyde 3-phosphate dehydrogenase was 5 ng and 1 ng for BSA. The LDR was between 1 and 500 ng of protein and 2D analysis of EA.hy 926 soluble proteins showed that native fluorescence detected 86% of spots relative to silver staining [194]. This procedure for in-gel protein detection is, however, very lengthy and requires up to two weeks for completion. Yet with the addition of 2,2,2-trichloroethanol to the gel matrix, native fluorescence protein detection could be carried out immediately after electrophoresis at 300 nm [195]. The LLD for low MW protein standards (Pharmacia) was 200 ng (Trp content 0.8–2.3%) and 20 ng protein (4.5% Trp). The LDR was between 0.2 and 2 μg protein/band (R = 0.99) and between 0.7 and 100 ng of Trp mass/protein [195]. It was undeniable that sensitive protein detection was highly dependent on the Trp content of proteins and hence resulted in a high degree of inter-protein variability.

Detection by native fluorescence was improved again with microelectrophoresis, where proteins are resolved with 3.5 × 8 cm slab gels (unlike mini-gels which are usually 6 × 8 cm). Detection of standard proteins (Sigma) was generally possible with as little as 0.1 ng of protein and based on this it was determined by the authors that the absolute limit of detection was 0.04 ng protein [196]. The LDR was between 0.1 and 20 ng of protein but the fluorescence intensities of each protein differed greatly from one another. This was not unexpected considering the varied number of Trp and tyrosine residues across the spectrum of proteins. Analysis of E. coli soluble protein extract (commercial preparation) by 2D analysis revealed that ∼300 spots could be detected (10 μg total protein load) [196]. The application of UV laser side entry excitation for native fluorescence protein detection has been shown to be even more sensitive in mini-gel applications [197]. The average LLD of the six protein standards used in this study (Sigma) was 5 pg per band. For these six standards, the LDR was between 20 pg and 16 ng protein and notable variation in fluorescence intensity was again revealed [197]. 2D analysis of a soluble sample prepared from E. coli (Bio-Rad) showed that limiting total protein loads to 1 and 0.25 μg did not alter the number of spots detected (280), but decreasing protein loads did negatively impact detection success [197]. UV–laser side entry excitation may be a more effective detection method than standard UV excitation and CCD camera detection as comparable spot numbers could be detected even with lower protein loads [196, 197].

Specific protein stains

This section will constitute a general overview of stains that are specific for particular protein moieties; however, most of the following stains have not been characterized extensively enough to fulfill all of the criteria established here for review. Like some of the techniques reviewed above, inter-protein variability is high with these stains, and indeed that is the key to their functional success; stain performance should correlate with the specific amount of a given moiety per protein, and is thus not uniform across a proteome. The cationic carbocyanine dye, 1-ethyl-2-{3- [1-ethylnaphtho (1,2d) thiazolin-2-ylidene]-2-methyl-propenyl}-naptho (1,2d) thiazolium bromide, for example, commonly known as SA, was one of the first stains used to distinguish proteins from RNA in bacterial polyribosomes [198]. In addition, by 1973, SA had been shown to differentially stain phosphorylated proteins blue and non-phosphorylated proteins red [199]. However, as studies continued with the use of SA it was determined that specificity could only be achieved if the sample components were known since both glycosylated and calcium binding proteins also stained blue [200202].

Glycosylated proteins

For the detection of glycosylated proteins in acrylamide gels, the periodic acid-Schiff stain (PAS) method was first introduced in 1964 and by 1969 had been further refined [203, 204]. Other modifications made to the PAS staining method have included the capacity for quick detection [205] and applicability to proteins resolved by native PAGE [206]. Alternative applications of the PAS stain for detection of glycoproteins include combining the PAS reaction with alcian blue or dansyl chloride staining [207210]. These procedures have not become widespread since their staining success depends on carbohydrate content with a minimum sugar requirement of 1 μg [209]. Although the thymol–sulfuric acid glycoprotein detection method is twofold more sensitive than the PAS method (limit of detection 0.05 μg carbohydrate), it is not stable and stained protein zones have been shown to fade and diffuse within a few hours at room temperature [211].

In addition to the fundamental staining demonstrated by PAS methods, the recent prevalence of fluorescence-based detection methods has led to the development of a sensitive fluorescent dye, Pro-Q Emerald 300, which has since become the stain of choice for glycoprotein detection [212]. Not only can this stain be useful in studying this protein modification in a sample but it can also be applied to quantification. Broad-range MW standards (Bio-Rad) were serially diluted and separated via 1D PAGE in order to assess the performance of Pro-Q Emerald 300 [212]. The manufacturer’s staining protocol delivered LLD values at or below the nanogram range for α1-acid glycoprotein (300 pg), glucose oxidase (300 pg) and avidin (1–2 ng) [212]. The LDR for α1-acid glycoprotein and glucose oxidase was demonstrated to be between ∼9 and 600 ng/lane and all other glycoproteins between ∼1.2 or 2.3–1,200 ng/lane (as derived from LDR figures and information provided on standard dilutions). Pro-Q Emerald 300 does not bind to non–glycosylated proteins but can detect lipopolysaccharides at concentrations of 2–4 ng, which may complicate crude protein extract analysis [212]. The main limitation to the use of Pro–Q Emerald 300 is that it cannot be used with laser-based gel scanners. This disadvantage led to the development of a new fluorescent glycoprotein stain, Pro-Q Emerald 488 [213]. Detection sensitivity of glycosylated protein depends greatly on carbohydrate content (α2-macroglobulin (9–10% CHO), glucose oxidase (12–13% CHO) and fetuin (22% CHO)–9.4 ng; α1-acid glycoprotein (38–42% CHO)–4.7 ng; avidin (7% CHO) and OVA (3–4% CHO)–18.8 ng). All glycoproteins were readily quantified over a 128–225-fold linear range, except avidin and OVA (64-fold, due to their carbohydrate content). It must also be noted that non-specific staining is observed when gels are heavily loaded with proteins that are not glycosylated (i.e., 250–1,000 ng). Pro-Q Emerald 488 is compatible with 2DE, however, since total protein stains used subsequently, such as SR, quench its fluorescence, simultaneous visualization is not possible and the stains must be used and detected serially [213].

Commercial glycoprotein stains available from Sigma (Glycoprotein detection kit) and Pierce (GelCode Glycoprotein Stain) detect glycoproteins based on modified versions of the PAS method, and have been successfully employed to reveal the glycosylation state of proteins [214, 215]. Although these colorimetric stains are useful, fluorescent detection of proteins is more sensitive. Glycoprofile III fluorescent kit (Sigma) and Krypton Glycoprotein staining kit (Pierce) also employ the periodate–oxidate chemistry to react with glycoproteins. The stains mentioned above detect total glycoprotein profiles of protein samples; however it is possible to detect sub-categories of glycosylation. Invitrogen has developed the Click-It™ O-GlcNAc Enzymatic Labeling System for detection of O-linked N-acetylglucosamine (O-GlcNAc) residues on target proteins. This system utilizes the enzymatic labeling of O-GlcNAc residues to azido-modified galactose via β-1,4-galactosyltransferase [216, 217]. This azide-labeled protein can then be fluorescently labeled with any of the Click-It™ detection regents, tetramethylrhodamine–alkyne (TAMRA; 300 nm UV illumination or 532 nm laser) or dapoxyl-alkyne (300 or 365 nm UV illumination) dyes. Determination of which alkyne dye to utilize may depend on the intended proteomic application. Both dyes are compatible with SR, but TAMRA can also be multiplexed with Pro-Q Emerald 300 glycoprotein gel stain and western detection with anti–TAMRA antibody, and Dapoxyl only with the Pro-Q Diamond phosphoprotein gel stain. Information available in regard to sensitivity of the above stain has only been stated by the manufacturer but has yet to be validated independently.

Phosphorylated proteins

Historically, phosphoproteins have been detected on polyacrylamide gels by radioactive means but this method of detection relies on radioactive phosphate being incorporated into proteins metabolically (preferably to equilibrium) and thus requires living cells [218]. In 1973, however, it was demonstrated that phosphorylated proteins could also be detected specifically via entrapment of liberated phosphate [219]. This method initially relies on the hydrolysis of phosphoester bonds under alkaline conditions in the presence of calcium ions to produce an insoluble calcium phosphate complex. This trapped phosphate is then treated with ammonium molybdate in dilute nitric acid to produce an insoluble nitrophosphomolybdate complex. Detection of this blue complex is then enhanced by staining with methyl green. This method is specific toward phosphoproteins, with a LLD of ∼3 μg (1 nmol of protein-bound phosphate) [219]. The GelCode Phosphoprotein staining kit (Thermo Scientific) and Phosphoprotein Stain kit (Ameresco) detect phosphoproteins based on this densitometric method developed by Cutting and Roth [219].

An alternate method for visualizing phosphoproteins is based on trivalent metal chelation [220]. In this method, aluminum ions are added to an acidic CBB stain solution. This promotes the formation of metal–protein chelates in which the aluminum reduces the negative charge and acts as a bridge between the dye and phosphate residue. This method could detect as little as 40 ng of apo-phosvitin, equivalent to 0.13 nmol of phosphate [220]. Like glycoproteins, a fluorescent dye for phosphoprotein detection is also now available. Pro-Q Diamond preferentially binds to phosphate moieties of proteins (weak non-specific binding to unphosphorylated protein was noted), can be used in conjunction with total protein stains and is MS compatible [221]. Pro-Q Diamond detects phospho-serines, -tyrosines, and -threonines with similar sensitivity [222]. Staining with Pro-Q Diamond can detect 1–2 ng of β-casein (five phosphate residues) and 8 ng of pepsin (one phosphate residue) [221]. As might be expected, it was also shown that total phosphate content (i.e., the total number of phosphorylated residues) influenced the detection limit for a particular protein. Non-specific detection of sulfonated moieties and others can also result in background issues. An advantage to the use of Pro-Q Diamond is that it can be diluted threefold without compromising sensitivity, fluorescence intensity or the LDR, thus substantially lowering the cost of use [223]. A new range of fluorescent phosphoprotein stains have been manufactured by PerkinElmer; the Phos–tag™ phosphoprotein stains are based on a metal (II) ion chelator that is highly selective for phosphomonoester residues of phospho-serine, -tyrosine, -threonine, -histidine, and -aspartate [224226]. This stain is also available in two forms to enable detection with a variety of gel imagers—Phos-tag™ 540 with maximum excitation at 540 nm and Phos-tag™ 300/460 with dual excitation at 300 and 460 nm.

Other proteins (iron-bound proteins, lipoproteins, protein tags)

Stains to detect other components of protein have also been developed. Ferene S {3–(2–pyridyl)–5,6–bis [2–(5–furylsulfonic acid)]–1,2,4–triazine, disodium salt} was introduced to detect non–heme protein–bound iron [227]. Although this method of staining with 0.75 mM Ferene S and 15 mM thioglycolic acid in HAc was rapid, it was relatively insensitive and the detection limits for the three proteins tested (ferritin, hemoglobin, and cytochrome c) varied greatly. A slightly more sensitive method for detecting non-heme iron proteins was developed and relied on the reaction of potassium ferricyanide with protein bound iron atoms [228]. This method yielded a LLD of 1 μg of ferritin, 2 μg of cyanobacteria extracted ferredoxin, and 100 μg human transferrin. The most sensitive iron stain to date, however, is based on a chemical reaction in which iron catalyzes the H2O2 oxidation of diaminobenzoate to an insoluble colored complex [229]. The detection limit was based on the amount of iron present per band, which was approximately 5 ng iron in 0.5 μg ferredoxin protein (spinach) and 4.7 ng iron for 1.9 μg of iron-containing superoxide dismutase (E. coli). Although this stain was sensitive and specific for the detection of protein bound iron, it could not differentiate between heme and non-heme protein-associated iron.

The detection of lipoproteins has commonly been carried out with Sudan Black B [230233] or Oil Red O [234, 235], however, these dyes are relatively insensitive. In 1994, a new dye for lipoprotein identification after native PAGE separation was demonstrated [236]. Filipin, a fluorescent stain could detect approximately 5 ng of unesterified cholesterol/band (based on pure low density lipoprotein) after 12 h of staining. This stain was suggested as part of a dual staining approach whereby lipoproteins were detected initially with filipin followed by total protein detection using CBB.

Proteomic stains have ventured still further and can now also be used for the selective staining of protein tags. Invision™ His-Tag In-gel stain (Invitrogen) is based on a fluorescent dye conjugated to a Ni2+/nitriloacetic acid complex. The Ni2+ metal binds selectively to the oligohistidine domain of His-tagged fusion proteins and as expected, detection varies and depends on the individual protein. A similar commercial His-tag protein stain is also available from Thermo Scientific (6xHis GelCode Protein Tag Staining Kit) and has been successfully applied [237]. Proteins can also be modified with the addition of a tetracysteine peptide (Cys–Cys–X–X–Cys–Cys—where X is a noncysteine amino acid). This reporter probe can be identified in the presence of FlAsH (a small, synthetic, membrane–permeable biarsenical ligand) where the interaction of the arsenic compound with the pair of thiol groups results in fluorescence [238]. This tetracysteine motif was optimized revealing that Cys–Cys–Pro–Gly–Cys–Cys had enhanced stability, the highest affinity and most rapid binding to biarsenical compounds [239]. This detection system for fused tetracysteine peptides is the foundation of the Lumino™ Green detection kit (Invitrogen) and detection sensitivity has not yet been proven outside the manufacturing labs. An overview of the published LLD values of all stains evaluated throughout this review is provided in Table 1.

Table 1.

Comparison of protein stains

Stain Details 1D SDS-PAGE 2D SDS-PAGE Mass spectrometry compatibility Reference
LLD Protein LDR Spot numbera
Coomassie Brilliant Blue 30 ng BSA/actin 0.5-20 μg/cm Yes [33], [4345]
CBB R-250 Mini gel staining microdensitometry 10 ng BSA, CA, β-lactalbumin 10-200 ng [46]
Colloidal CBB 0.1-1 ng BSA 30-500 ng 250 (A. thaliana total protein) [17], [47]
Commercial cCBB 8-16 ng Broad-range MW standards 30-250 ng [20]
Neuhoff formulation 4-8 ng
Rapid Silver Stain with careful temp. control 0.1-0.2 ng Standard Proteins 1-30 ng No [46]
Silver staining with EtOH used as primary solvent 27 pg/mm3 Soybean trypsin inhibitor 27 pg/mm3-5-50 ng/mm3 1,800 (mouse liver cytosolic proteins) No [72]
Silver Nitrate 1 ng A. thaliana total soluble extract Yes [74]
Acidic silver nitrate 2-4 ng Broad-range MW Std 4-60 ng Not without compromising LLD [20]
Alkaline silver diamine 2-4 ng Broad-range MW Std 8-60 ng N0 [20]
Silver staining preceded by calconcarboxylic acid, sensitization 0.05-0.2 ng/band CA, Phos B, β-galactosidase, ovalbumin, myosin 0.8-100 ng (myosin, B-galactosidase, Phos B) and 0.2-25 ng (ovalbumin, CA) Yes [88]
Silver Stain with ethyl violet zincon 0.2 ng/band BSA, CA, Phos B, β-galactosidase, albumin, myosin 4-50 ng Yes [90]
Ethyl violet Zincon–MeOH solvent > 4-8 ng PhosB, BSA, chicken OVA, bovine CA and human peroxiredoxin I Between 8 and 1,000 ng [116]
Ethyl Zincon, longer fixation and EtOH solvent 4-8 ng Phos B, BSA, Chicken OVA, bovine CA 8-1,000 ng Yes [49]
MDPF 1 ng CA, myoglobin, catalase, BSA 1-500 ng [138]
60 μg MDPF: 20 μg protein
OPA ∼10 ng Transferrin 0.1-50 μg/ml [139]
0.19 μmoles OPA: up to 50 μg/ml protein
Fluorescamine 6 ng Myoglobin 0.5-12.5 μg [140]
0.5 mg fluorescamine: Up to 1 mg/ml protein
DIGE 0.25 ng MLH and Pirella Strain 3-4 orders of magnitude 367 (av.) (Pirellula sp. Culture) Yes [10], [142]
200-400 pmol/50 μg protein 401 (av.) (APAP exposed livers)
BODIPY dyes 10 fmol Yeast enolase 1 3 orders of magnitude Yes [148]
mBB 4-32 ng Broad range MW Standards 32-100 ng Weak (BSA, soybean trypsin inhibitor misidentified) [20]
mBB 86.2 ± 14 ng Recombinant Proteins 86.2 ng-8.47 mg (largest range) [149]
CyDye maleimides 0.1-5 ng Albumin 3-4 orders of magnitude ∼2,500 (microdissected pancreatic cancer cells) Yes [132], [152]
DY680 and DY780 Dyes 10 fg Labeled tubullin Yes [153]
200 μg fluorophore: 25 mg protein
DY505 and DY635 Dyes 0.13-1 ng Human serum albumin 3 orders of magnitude 1,212 ± 124 1,050 ± 28, respectively (keratinocyte cell lysate) Yes [154]
20 μg total protein
ICy3 and ICy5 2 ng BSA 3-4 orders of magnitude 1,034 ± 245 (Erb2 overexpressing cells) Yes [147]
80 pmol fluorophore: 1 μg protein
SR (manufacturer’s protocol) 1-2 ng Broad range MW Std 1-100 ng 1,290 ± 34 (rat fibroblast total cell lysates) Yes [20], [73], [157, 158]
500 ng A. thaliana total protein
SR, MeOH/HAc fixation, stain from 3 h to overnight, destain in MeOH/HAc, water washed. 0.5-5 ng Recombinant protein 200-1,000-fold 1,324 ± 142 (using E.coli total protein and PDQuest) Yes [156]
RuBPS 10-20 ng Broad range MW Std 2-2,500 ng ∼1,800 (thyroid cell extract) Yes [168]
ASCQ_Ru 80 pg Glutamate dehydrogenase 4 orders of magnitude 1,537 (Human colon carcinoma cells) Not fully with trypsin [171]
DP 64 pg/band Low MW Std 4 orders of magnitude 1,102 (mouse/rat liver tissue) Yes [173]
C-16 fluorescein 0.12 ng BSA 7.8-125 ng/band Yes [180]
LUCY® Dyes 2-10 ng BSA [176]
Palmatine 2 ng BSA, CA, PhosB, β–galactosidase, albumin, mysoin 2-500 ng Yes [158]
BisANS (new protocol) 1 ng/band BSA, CA, PhosB, β–galactosidase, albumin, mysoin 1-250 ng Yes [157]
Pro-Q Emerald (488) 300 pg α1-acid glycoprotein and glucose oxidase ∼9-600 ng/lane [212]
Pro-Q Diamond 1-2 ng α-casein 500-100-fold Yes [221]
Zinc staining 1-2 ng Broad-range MW Std 250-2,000 ng [20]
Rapid zinc stain kit 1.8-4 ng/band Low MW Std 1.8-140 ng/band ∼350 (TCA precipitated human hepatocytoma cells) Yes [50]
Trivalent metal chelation 40 ng Apo-phosvitin [220]
Ferene S 1 μg Ferritin [228]
Iron and diaminobenzoate oxidation of H2O2 0.5 μg Ferredoxin [229]
Filipin 5 ng/band Unesterified cholesterol [236]

av. Average of all three cyanine dyes

aOnly reported if a quantitative number is provided.

Equipment innovations

In addition to considering the physiochemical properties of a stain it is important to remember that detection will also be influenced by our ability to quantitatively assess the stain. Detecting stained protein depends as much on the instruments and equipment that are available as on the stain itself. When the innovations in CBB staining that fathered quantitative staining as it is known today were made, densitometry was measured on a recording strip scanner with a film attachment [18]. While this machine only had a 4.65-fold level of magnification, it also performed the integrations necessary to measure protein amount. For the detection of CBB, amido black and silver stain during this era, densitometers (with varying degrees of automation) were the main form of imaging equipment. In 1968, for example, a Joyce Loebl Chromoscan recording and integrating densitometer was used to quantify amido black [240]. Two years later, there was improved automation with the use of a plexiglass cartridge driven by a motor that passed below a Gilford Model 220 absorbance indicator [241]. This apparatus was interfaced with a Gilford Optical Density Converter and Healthkit Servo Recorder. The sensitivity of this system was based on the recorder having four known ratio settings, thereby enabling adjustment to different band intensities. A spectrophotometer was used to densitometrically detect Drimarene and Uniblue A stained proteins; this instrument required the excision of gel pieces for measurement, including an unstained gel slice of the same thickness to serve as a blank [114]. The use of the spectrometer for detection, transport for more rapid sample exchange, and a recorder (or similar equipment), dominated densitometric imaging techniques for most gel assessment until the mid-to-late 1970s [43, 76, 118, 199, 200, 205, 242].

For silver-stained proteins, it was more common to photograph gels and scan these images for analysis, although densitometers were used as well [59, 6264]. One study used a Cohu Model 7120, 525 line, black and white camera to image silver stained proteins, and pictures were digitized using a Colorado Video Model 270 digitizer with 512 × 480 pixels and a 0 to 255 gray density value scale [72]. The consensus for zinc–imidazole-stained gels was also to take images, against a black background in this case, usually using a Polaroid camera or an automated imager, and follow with analysis [20, 100, 101, 104, 107, 108, 243].

Fluorescence posed a different challenge in terms of detection given the specific excitation and emission requirements of different fluorophores. In 1969, the fluorescence of proteins pre-labeled with anilinonaphthalene sulfonate and separated on a polyacrylamide gel were detected in two stages; it was necessary to first excite the fluorophore with a long wave UV lamp before capturing the image with an everyday camera fitted with appropriate filters [181]. This dual imaging process was widespread. In one study, a desk lamp was adapted by lining the shade with foil and fitting it with a UV bulb to visualize dansyl chloride-tagged proteins [244]. Following fluorophore excitation, the gel image was once again recorded with a ‘filtered’ camera. This practice for detecting fluorescence continued throughout the 1970s and was used in conjunction with a variety of fluorophores, including OPA and fluorescamine [139, 140, 245]. Even in the mid-1970s, Gilford instruments still required adjustment to facilitate the measurement of fluorescence [137, 140]. Not only was a filter (Corning 051) required to remove excitation energy but since the Gilford was designed to read absorbance, it was necessary to calculate the antilogarithm of peaks to obtain a measurement of signal. It gradually became more common to transilluminate gels with UV light to detect fluorescently stained proteins [20, 157, 158, 182, 184, 186, 190, 191, 195, 246, 247]. Subsequently, photographs of transilluminated gels had to be scanned by a densitometer before quantitative analysis could be undertaken; again, all without the aid of local computers that are now considered standard lab equipment.

During this time, the use of CCD cameras began to be introduced in electrophoresis literature. In 1988, the use of the CCD 2200 Imaging System which boasted 385 × 578 pixels was reported [138]. The maximum signal detectable by this camera was 105 electrons while the read noise or minimum signal was 1 electron. The range of detection afforded by this system presented an important development in the routine quantification of proteins and the application of CCD based systems for imaging gels still remains one of the most prevalent technologies in use today [193, 194, 196, 197, 236, 247249].

In the 1990s, specialized equipment for automated densitometric and fluorescent detection of proteins was marketed. While the transilluminator/camera option remained in use due to its accessibility [149], an Elsie 5 computer analysis system was also available [250]. It could be described as the end product of some 30 years worth of multistep gel handling in pursuit of quantitative analysis. Elsie 5 was a system designed specifically for the analysis of 2D gels. Not only did the Elsie 5 image gels but it was also designed to facilitate quantification and image manipulation including comparison with other gel images. The major advantage to using such systems for detection, as they are the basis for most current analyses, is their ability to reduce human error and bias and increase throughput; this is the cornerstone of all large-scale analyses of specific molecules or ‘omics’.

By 2001, a CCD camera with 1,024 × 1,024 pixel resolution was being used to detect proteins stained with SR and tagged with the fluorescent reporter gene, β-glucuronidase [251]. Despite being able to identify signal by eye in this study, the authors recommended the use of a CCD camera for the most optimal and accurate analysis. A comparison of three different in-gel protein detection approaches using MDPF, mBB, and SR was also carried out [252]. The CCD camera used in this study was automated and fully accessorized with the necessary filters and UV transilluminator to visualize each of these stains. A similar instrument, the Typhoon 8600 was also in fairly widespread use during this decade [253, 254].

Although CCD cameras revolutionized the in-gel detection of stained proteins and are thus in extensive use in the proteomic field, new technologies are emerging which may displace their prevalence. Complementary-metal-oxide-semiconductor devices allow for ultrasensitive detection through signal amplification [255]. These devices can be applied to many different detection systems including gel imaging which may present great potential in terms of decreasing detection limit.

Just as the ability to image gels affects the capacity for analysis, so too do the programs used to quantify and determine the signal associated with stained protein. A number of programs are available for gel image analysis and these have developed from hardware intensive systems that lacked a visual user interface to sophisticated gel analysis software that can be run from a desktop computer [256]. Currently, Delta2D (DECODON) and Progenesis SamesSpots (Nonlinear Dynamics) utilize an approach that first warps the images and then matches spots. This minimizes the problems associated with matching spots from gels that may have slight variations. While this approach is faster, the approach used by PDQuest (Bio-Rad), Decyder 2D (GE Healthcare), Melanie III (Genebio) and Dymension (Syngene), of detecting spots first and then following with image warping remains valid. Each of these programs also offers a differing array of options for editing images, statistical analysis and user interaction depending on the users’ needs. A comparison of different analytical programs rated PDQuest (Bio-Rad) as best able to deal with spot overlap but described Melanie III (Genebio) and Progenesis SameSpots (Nonlinear Dynamics) as “all-rounders” in terms of accuracy and coping with low S/N ratios [133]. A comparison of Delta2D and Proteomweaver (Bio-Rad) showed that Delta2D outperformed the Bio-Rad analysis program in spot detection, automatic spot matching and manual correction after warping [257]. Further, the use of a consensus gel image compiled from each gel image in an experimental series economizes spot detection and editing steps in analysis using Delta2D.

The advent of MS was a huge leap forward in protein identification [258]; this was and remains especially powerful as it links and promotes analysis using the vast number of available databases, including SwissProt, TrEMBL and NCBInr. In this review it is clear that most of the cited papers used MALDI–ToF for protein identification. In recent years, however, electrospray ionization (ESI) has been widely applied with MS. As ESI ionizes proteins from solution it is easily amenable to liquid chromatography MS (LC-MS) [259]. Liquid chromatographic separations of tryspin digested samples have increased sensitivity since few peptides elute at any one time. This can be valuable in those instances in which multiple proteins resolve to a single spot after 2DE. Not only does LC-MS achieve greater protein sequence coverage (particularly when iterative analyses are used), it is also a useful technique for analyzing complex protein mixtures when additional electrophoretic separation would otherwise be required [259, 260].

Coupled with these advances has been the further development of mass analyzers. Together, specific ionization and the development of ion trapping mass analyzers were able to achieve greater sensitivity and sequence coverage. A continuation of advances in ionization includes the relatively new Fourier transform-ion cyclotron resonance (FT-ICR) MS which can now measure protein quantities at low to sub-ppm ranges [261]. The new orbitrap mass analyzer also demonstrates this level of sensitivity and resolution but separates ions in an oscillating electric field. Also, while traditional MS utilizes collision-induced dissociation ions, alternatives which produce more uniform fragmentation, such as electron capture dissociation and electron transfer dissociation have been developed. Higher levels of sensitivity in MS methods would allow for the successful analysis of low copy number proteins that have proven difficult to identify from 2DE gels without sample enrichment; although in current circumstances serendipity still plays a role here, and even if a suspected low abundance protein is found, identification more often than not relies on a single peptide.

Although MALDI–ToF and ESI-MS are among the leading techniques for protein identification, advances have been made towards new MS proteomic strategies. These include so-called ‘shotgun’ approaches in which whole protein extracts are digested prior to to MS analysis; advances in new mass analyzers (the performance of these new instruments have been summarized in detail by Han et al. [262]; Table 1) as well as new quantitative strategies employing metabolic amino acid labeling (stable isotope labeling with amino acids in cell culture, isotope-coded affinity tagging, and isobaric tags for relative and absolute quantification) have demonstrated sub-femtomole sensitivity [260, 262265]. Nonetheless, these approaches have their own inherent limitations in terms of extent and reproducibility of the labeling reactions or breadth of applicability; ongoing refinements of these tools and their application will almost certainly overcome most difficulties, as has been the case with the maturing of gel-based proteomics.

It is clear that when detection technology moved forward, not only was there an increase in sensitivity but also a more comprehensive integration of the multiple tools required for detecting and identifying proteins; we’ve gone from jury-rigged UV illuminators and hand held cameras to the fully integrated imager, such as the Fuji LAS–3000, with illumination, filters, and camera all inclusive in the design. Data is also more easily accessible as images are already formatted for quantitative analysis. Furthermore, these advances in detection and imaging technology may well mean that some previously characterized stains are worth re-evaluating. Given the decreasing protein masses that are now routinely detectable, with the equipment now available it is possible that the quantitative values (LLD and LDR) originally reported one-to-four decades ago no longer accurately reflect the true detection capacity of these stains. Examples of stains already demonstrating improved LLDs with modern detection technology include the IR fluorescent detection of cCBB and the 2.5-fold improvement in the sensitivity of mBB detection between the years 1994 and 2000. Furthermore, without MS and its integration with databases to aid protein identification, a significant proportion of proteomic investigation would be little more than large-scale exercises in electrophoresis. It is the convergence of these concepts and techniques in pursuit of comprehensive, large scale analyses of native proteins that has made modern proteomics possible. Indeed, quantitative image analysis and ‘hyphenated’ PAGE–MS are perhaps the most important developments and techniques in our modern pursuit of proteomics. A continued drive for greater sensitivity and thus better overall detection will impel the pursuit of additional improvements and thus further, ‘deeper’ dissection of proteomes.

Discussion

Despite a moderate amount of success, historically, in detecting and quantifying proteins in-gel, each staining revolution also brings forth its own limitations. The effectiveness of some of these staining techniques must, in some cases, be questioned. For example, pre-labeling with cyanine succinimidyl esters and maleimides is reported to deliver subnanogram detection sensitivity rapidly and with minimal background. Practitioners also claim that low abundance proteins are easily accessed for detection and that detection is linear over up to five orders of magnitude. However, there are a few inherent complications in the process. For instance, labeling alters the molecular weight and pI of proteins and both the protein itself and its labeled derivative (s) have to then be resolved. Two (or more) resolved protein spots for each protein species doubles the total number of spots and reduces the effective resolution of the gel [20]. Labeling can also disrupt the structure and physical characteristics of proteins [266]. This means that excising these proteins for MS application requires estimation of the difference in migration to locate the unlabeled protein. However, post-electrophoretic staining is not without its own limitations; these gels usually have higher background levels and none have been reported to have LDR values of more than four orders of magnitude [267]. Also, even the most rapid staining techniques require washing and fixing in addition to the staining step. This makes post staining a more time intensive process although one that does not alter the native proteome and can provide for quantitative analysis. Low molecular weight proteins of low abundance continue to be poorly detected by any method, and its limitation still plagues even the most innovative staining approaches. At this stage, according to the criteria used here, SR likely remains the benchmark post-electrophoretic stain in the compromise between performance, ease of use and cost. DIGE, however, is also somewhat popular due to its claims of high sensitivity and multiple sample comparisons. Nonetheless, it remains that this method relies on loading less sample in an effort to effect multiple separations within a single gel; it has still not been effectively established that this approach is in any way superior to resolving full protein loads on separate gels and using powerful, widely available imaging systems/packages to most comprehensively compare the resolved proteomes.

Prior to beginning this review, it was necessary to identify the characteristics of a stain that would indicate its protein quantification efficacy. What was its minimum detection limit (i.e., LLD)? How broad was the linear relationship between signal and protein quantity (i.e., LDR)? How uniform was staining between different proteins (i.e., inter-protein variability)? These were the most widely reported characteristics, with LLD and LDR commonly used as determinants of the staining sensitivity. One of the problems associated with LLD reporting is that between different papers, units are very rarely uniform and there exists no standard unit of measurement for sensitivity. This becomes a impediment to the unequivocal comparison of stain performance. In this review, these criteria were minimally required to facilitate quantitative and comparative evaluation of stains used in proteomics. Unfortunately, some definitive criteria were consistently omitted from studies. Inter–protein variability was rarely addressed. Another frequent omission was reference to lowest limit of quantification (LLQ). In addition to LDR, this characteristic describes the functional ‘window’ of a stain; the concentration range in which a stain can be used quantitatively. An LLD does not guarantee a quantifiable signal that relates to protein quantity. If anything, LLQ is a more useful reporter of stain performance as it indicates the limit of useful quantitative capacity. Very few papers reported LLQ, likely further emphasizing the critical and growing need for better and wider interactions between proteomics and physical chemistry (for example, see International Congress on Analytical Proteomics) [268].

In addition to regularly including supplementary quantitative criteria, such as LLQ and inter-protein variability, when evaluating stain performance we must also actively assess the equipment used in any given evaluation. When reviewing stains last evaluated decades ago, the question of how large an improvement contemporary technology might bring to sensitivity assessments cannot be avoided. Certainly the increased sensitivity demonstrated by CBB when imaged using IR light [21] suggests that other older stains may also benefit from examination with higher resolution imaging equipment. The quantitative stains we need, or at least the chemistries required for their future development, may well already exist.

Acknowledgements

The authors thank Dr. R.H. Butt for insightful discussions during the writing of this review.

Abbreviations

2DE

two-dimensional electrophoresis

BisANS

4,4′-dianilino-1,1′-binaphthyl-5,5′-disulfonic acid

BSA

bovine serum albumin

C16-F

C-16 fluorescein

CA

carbonic anhydrase

CBB

Coomassie Brilliant Blue

cCBB

colloidal Coomassie Brilliant Blue

CCD

cooled charge device

Cy2

Cyanine 2

Cy3

Cyanine 3

Cy5

Cyanine 5

cys

cysteine

DIGE

differential gel electrophoresis

DMSO

dimethyl sulfoxide

DP

deep purple

DTT

dithiothreitol

EBT

erichrome black T

EDTA

ethylene diamine tetraacetic acid

EtOH

ethanol

EZ

ethyl violet and zincon

HAc

acetic acid

HSA

human serum albumin

ICy

iodoacetylated cyanine dye

IEF

isoelectric focusing

LC

liquid chromatography

LDR

linear dynamic range

LLD

lowest limit of detection

lys

lysine

MALDI-ToF-MS

matrix-assisted laser desorption ionization-time of flight-mass spectrometry

mBB

monobromobimane

MDPF

two-methoxy-2,4-diphenyl-3(2H)-furanone

MeOH

methanol

MS

mass spectrometry

MW

molecular weight

OPA

o-phthalaldehyde

OVA

ovalbumin

PAGE

polyacrylamide gel electrophoresis

PAS

periodic acid-Schiff

PhosB

phosphorylase b

PMF

peptide mass fingerprinting

RuBPS

ruthenium (II) tris (bathophenanthroline disulfonate)

SA

Stains All (1-ethyl-2-{3- [1-ethylnaphtho (1,2d) thiazolin-2-ylidene]-2-methyl-propenyl}-naptho (1,2d) thiazolium bromide)

SDS

sodium dodecyl sulfate

SNR

signal-to-noise ratio

sp.

species

SR

SYPRO Ruby

TCEP

Tris (2-carboxyethyl) phosphine hydrochloride

Trp

tryptophan

UV

ultraviolet

Footnotes

Victoria J. Gauci and Elise P. Wright contributed equally to this study.

References

  • 1.Mathews CK, Holde KE, Ahern KG. Biochemistry. 3. Canada: Addison-Wesley Publishing Company; 2000. p. 1186. [Google Scholar]
  • 2.Lodish H, Berk A, Matsudaira P, Kaiser CA, Kreiger M, Scott MP, Zipursky SL, Darnell J. Molecular cell biology. 5. New York: W. H. Freeman Company; 2004. [Google Scholar]
  • 3.Wittmann-Liebold B, Graack H-R, Pohl T. Two-dimensional gel electrophoresis as tool for proteomics studies in combination with protein identification by mass spectrometry. Proteomics. 2006;6:4688–4703. doi: 10.1002/pmic.200500874. [DOI] [PubMed] [Google Scholar]
  • 4.Righetti PG. Electrophoresis: the march of pennies, the march of dimes. J Chromatogr A. 2005;1079:24–40. doi: 10.1016/j.chroma.2005.01.018. [DOI] [PubMed] [Google Scholar]
  • 5.Ong S-E, Pandey A. An evaluation of the use of two-dimensional gel electrophoresis in proteomics. Biomol Eng. 2001;18:195–205. doi: 10.1016/s1389-0344(01)00095-8. [DOI] [PubMed] [Google Scholar]
  • 6.O'Farrel PH. High resolution two-dimensional electrophoresis of proteins. J Biol Chem. 1975;250:4007–4021. [PMC free article] [PubMed] [Google Scholar]
  • 7.Klose J. Protein mapping by combined isoelectric focusing and electrophoresis of mouse tissue. Humangenetik. 1975;26:231–243. doi: 10.1007/BF00281458. [DOI] [PubMed] [Google Scholar]
  • 8.Anderson NL, Nance SL, Tollaksen SL, Giere FA, Anderson NG. Quantitative reproducibility of measurements from coomassie blue-stained two-dimensional gels: Analysis of mouse liver protein patterns and a comparison of BALB/c and C57 strains. Electrophor. 1985;6:592–599. [Google Scholar]
  • 9.Blomberg A, Blomberg L, Norbeck J, Fey SJ, Larsen PM, Larsen M, Roepstorff P, Degand H, Boutry M, Posch A, Görg A. Interlaboratory reproducibility of yeast protein patterns analyzed by immobilized pH gradient two-dimensional gel electrophoresis. Electrophor. 1995;16:1935–1945. doi: 10.1002/elps.11501601320. [DOI] [PubMed] [Google Scholar]
  • 10.Gade D, Thiermann J, Markowsky D, Rabus R. Evaluation of two-dimensional difference gel electrophoresis for protein profiling. J Mol Microbiol Biotechnol. 2003;5:240–251. doi: 10.1159/000071076. [DOI] [PubMed] [Google Scholar]
  • 11.Butt RH, Lee MWY, Pirshahid SA, Backlund PS, Wood S, Coorssen JR. An initial proteomic analysis of human preterm labor: placental membranes. J Proteome Res. 2006;1:391–396. doi: 10.1021/pr060282n. [DOI] [PubMed] [Google Scholar]
  • 12.Butt RH, Pfeifer TA, Delaney A, Grigliatti TA, Tetzlaff WG, Coorssen JR. Enabling coupled quantitative genomics and proteomics analyses from rat spinal cord samples. Mol Cell Proteomics. 2007;6:1574–1588. doi: 10.1074/mcp.M700083-MCP200. [DOI] [PubMed] [Google Scholar]
  • 13.Rabilloud T. Two-dimensional gel electrophoresis in proteomics: old, old fashioned, but it still climbs up the mountains. Proteomics. 2002;2:3–10. [PubMed] [Google Scholar]
  • 14.Hoving S, Voshol H, J Van Oostrum. Towards high performance two-dimensional gel electrophoresis using ultrazoom gels. Electrophor. 2000;21:2617–2621. doi: 10.1002/1522-2683(20000701)21:13<2617::AID-ELPS2617>3.0.CO;2-C. [DOI] [PubMed] [Google Scholar]
  • 15.Gygi SP, Corthals GL, Zhang Y, Rochon Y, Aebersold R. Evaluation of two-dimensional gel electrophoresis-based proteome analysis technology. Proc Natl Acad Sci USA. 2000;97:9390–9395. doi: 10.1073/pnas.160270797. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 16.Butt RH, Coorssen JR. Postfractionation for enhanced proteomic analyses: routine electrophoretic methods increase the resolution of standard 2D-PAGE. J Proteome Res. 2005;4:982–991. doi: 10.1021/pr050054d. [DOI] [PubMed] [Google Scholar]
  • 17.Chevalier F, Rofidal V, Vanova P, Bergoin A, Rossignol M. Proteomic capacity of recent fluorescent dyes for protein staining. Phytochem. 2004;65:1499–1506. doi: 10.1016/j.phytochem.2004.04.019. [DOI] [PubMed] [Google Scholar]
  • 18.De St. Groth SF, Webster RG, Datyner A (1963) Two new staining procedures for quantitative estimation of proteins on electrophoretic strips. Biochem Biophys Acta 71:377–391 [DOI] [PubMed]
  • 19.Switzer RC, Merril CR, Shifrin S. A highly sensitive silver stain for detecting proteins and peptides in polyacrylamide gels. Anal Biochem. 1979;98:231–237. doi: 10.1016/0003-2697(79)90732-2. [DOI] [PubMed] [Google Scholar]
  • 20.Berggren K, Chernokalskaya E, Steinberg TH, Kemper C, Lopez MF, Diwu Z, Haugland RP, Patton WF. Background-free, high sensitivity staining of proteins in one- and two-dimensional sodium dodecyl sulfate-polyacrylamide gels using a luminescent ruthenium complex. Electrophor. 2000;21:2509–2521. doi: 10.1002/1522-2683(20000701)21:12<2509::AID-ELPS2509>3.0.CO;2-9. [DOI] [PubMed] [Google Scholar]
  • 21.Harris LR, Churchwood MA, Butt RH, Coorssen JR. Assessing detection methods for gel-based proteomic analyses. J Proteome Res. 2007;6:1418–1425. doi: 10.1021/pr0700246. [DOI] [PubMed] [Google Scholar]
  • 22.Long GL, Winefordner JD. Limit of detection. A closer look at the IUPAC definition. Anal Chem. 2008;55:712A–724A. [Google Scholar]
  • 23.Raymond S, Weintraub L. Acrylamide gel as a supporting medium for zone electrophoresis. Science. 1959;130:711. doi: 10.1126/science.130.3377.711. [DOI] [PubMed] [Google Scholar]
  • 24.Raymond S, Wang Y-J. Preparation and properties of acrylamide gel for use in electrophoresis. Anal protein cyclization promoted by a circularly permuted Synechocystis sp. PCC6803 DnaD mini-intein. J Biol Chem. 1960;277:7790–7798. doi: 10.1074/jbc.M110303200. [DOI] [PubMed] [Google Scholar]
  • 25.Raymond S. Acrylamide gel electrophoresis. Ann NY Acad Sci. 1964;121:350–365. doi: 10.1111/j.1749-6632.1964.tb14208.x. [DOI] [PubMed] [Google Scholar]
  • 26.Meyer TS, Lamberts BL. Use of coomassie brilliant blue R250 for the electrophoresis of microgram quantities of parotid saliva proteins on acrylamide-gel strips. Biochim Biophys Acta-Gen Sub. 1965;107:144–145. doi: 10.1016/0304-4165(65)90403-4. [DOI] [PubMed] [Google Scholar]
  • 27.Compton SJ, Jones CG. Mechanism of dye response and interference in the Bradford protein assay. Anal Biochem. 1985;151:369–374. doi: 10.1016/0003-2697(85)90190-3. [DOI] [PubMed] [Google Scholar]
  • 28.Congdon RW, Muth GW, Splittgerber AG. The binding interaction of coomassie blue with proteins. Anal Biochem. 1993;213:407–413. doi: 10.1006/abio.1993.1439. [DOI] [PubMed] [Google Scholar]
  • 29.Tal M, Silberstein A, Nusser E. Why does Coomassie Brilliant Blue R interact differently with different proteins? A partial answer. J Biol Chem. 1985;260:9976–9980. [PubMed] [Google Scholar]
  • 30.Vesterberg O. Staining of protein zones after isoelectric focusing in polyacrylamide gels. Biochim Biophys Acta-Protein Structure. 1971;243:345–348. doi: 10.1016/0005-2795(71)90094-8. [DOI] [PubMed] [Google Scholar]
  • 31.Diezel W, Kopperschläger G, Hofmann E. An improved procedure for protein staining in polyacrylamide gels with a new type of Coomassie Brilliant Blue. Anal Biochem. 1972;48:617–620. doi: 10.1016/0003-2697(72)90117-0. [DOI] [PubMed] [Google Scholar]
  • 32.Malik N, Berrie A. New stain fixative for proteins separated by gel isoelectric focusing based on coomassie brilliant blue. Anal Biochem. 1972;49:173–176. doi: 10.1016/0003-2697(72)90255-2. [DOI] [PubMed] [Google Scholar]
  • 33.Reisner AH, Nemes P, Bucholtz C. The use of coomassie brilliant blue G250 perchloric acid solution for staining in electrophoresis and isoelectric focusing on polyacrylamide gels. Anal Biochem. 1975;64:509–516. doi: 10.1016/0003-2697(75)90461-3. [DOI] [PubMed] [Google Scholar]
  • 34.Vesterberg O, Hansén L, Sjösten A. Staining of proteins after isoelectric focusing in gels by new procedures. Biochim Biophys Acta-Protein Structure. 1977;491:160–166. doi: 10.1016/0005-2795(77)90052-6. [DOI] [PubMed] [Google Scholar]
  • 35.Zehr BD, Savin TJ, Hall RE. A one-step, low background coomassie staining procedure for polyacrylamide gels. Anal Biochem. 1989;182:157–159. doi: 10.1016/0003-2697(89)90734-3. [DOI] [PubMed] [Google Scholar]
  • 36.Chen H, Cheng H, Bjerknes M. One-step coomassie brilliant Blue R-250 staining of proteins in polyacrylamide gels. Anal Biochem. 1993;212:295–296. doi: 10.1006/abio.1993.1330. [DOI] [PubMed] [Google Scholar]
  • 37.Mitra P, Pal AK, Basu D, Hati RN. A staining procedure using coomassie brilliant blue G-250 in phosphoric acid for detection of proteinbands with high resolution in polyacrylamide gel and nitrocellulose membrane. Anal Biochem. 1994;223:327–329. doi: 10.1006/abio.1994.1594. [DOI] [PubMed] [Google Scholar]
  • 38.Sreeramulu G, Singh NK. Destaining of coomassie brilliant blue R-250-stained polyacrylamide gels with sodium chloride solutions. Electrophor. 1995;16:362–365. doi: 10.1002/elps.1150160162. [DOI] [PubMed] [Google Scholar]
  • 39.Nivinskas H, Cole KD. Environmentally benign staining procedures for electrophoresis gels using coomassie brilliant blue. Biotech. 1996;20:380–385. doi: 10.2144/19962003380. [DOI] [PubMed] [Google Scholar]
  • 40.Wu W, Welsh MJ. Rapid coomassie blue staining and destaining of polyacrylamide gels. Biotech. 1996;20:386–388. doi: 10.2144/19962003386. [DOI] [PubMed] [Google Scholar]
  • 41.Kang D, Gho YS, Suh M, Kang C. Highly sensitive and fast protein detection with coomassie brilliant blue in sodium dodecyl sulfate-polyacrylamide gel electrophoresis. Bull Korean Chem Soc. 2002;23:1511–1512. [Google Scholar]
  • 42.Chrambach A, Reisfeld RA, Wyckoff M, Zaccari J. A procedure for rapid and sensitive staining of protein fractionated by polyacrylamide gel electrophoresis. Anal Biochem. 1967;20:150–154. doi: 10.1016/0003-2697(67)90272-2. [DOI] [PubMed] [Google Scholar]
  • 43.Kahn R, Rubin RW. Quantitation of submicrogram amounts of protein using coomassie brilliant blue R on sodium dodecyl sulfate-polyacrylamide slab-gels. Anal Biochem. 1975;67:347–352. doi: 10.1016/0003-2697(75)90305-x. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 44.Blakesley RW, Boezi JA. A new staining technique for proteins in polyacrylamide gels using coomassie brilliant blue G250. Anal Biochem. 1977;82:580–582. doi: 10.1016/0003-2697(77)90197-x. [DOI] [PubMed] [Google Scholar]
  • 45.Choi J-K, Yoon S-H, Hong H-Y, Choi D-K, Yoo G-S. A modified coomassie blue staining of proteins in polyacrylamide gels with bismark brown R. Anal Biochem. 1996;236:82–84. doi: 10.1006/abio.1996.0134. [DOI] [PubMed] [Google Scholar]
  • 46.Poehling H-M, Neuhoff V. Visualization of proteins with a silver stain: a critical analysis. Electrophor. 1981;2:141–147. [Google Scholar]
  • 47.Neuhoff V, Arold N, Taube D, Ehrhardt W. Improved staining of proteins in polyacrylamide gels including isoelectric focusing gels with clear background at nanogram sensitivity using coomassie brilliant blue G-250 and R-250. Electrophor. 1988;9:255–262. doi: 10.1002/elps.1150090603. [DOI] [PubMed] [Google Scholar]
  • 48.Neuhoff V, Stamm R, Eibl H. Clear background and highly sensitive protein staining with Coomassie Blue dyes in polyacrylamide gels: a systematic analysis. Electrophor. 1985;6:427–448. [Google Scholar]
  • 49.Choi J-K, Chae H-Z, Hwang S-Y, Choi H-I, Jin L-T, Yoo G-S. Fast visible dye staining of proteins in one- and two-dimensional sodium dodecyl sulfate-polyacrylamide gels compatible with matrix assisted laser desorption/ionization-mass spectrometry. Electrophor. 2004;25:1136–1141. doi: 10.1002/elps.200305776. [DOI] [PubMed] [Google Scholar]
  • 50.Lin C-Y, Wang V, Shui H-A, Juang R-H, Hour A-L, Chen P-S, Huang H-M, Wu S-Y, Lee J-C, Tsai T-L, Chen H-M. A comprehensive evaluation of imidazole-zinc reverse stain for current proteomic researches. Proteomics. 2009;9:696–709. doi: 10.1002/pmic.200700470. [DOI] [PubMed] [Google Scholar]
  • 51.Candiano G, Bruschi M, Musante L, Santucci L, Ghiggeri GM, Carnemolla B, Orecchia P, Zardi L, Righetti PG. Blue silver: a very sensitive colloidal coomassie G-250 staining for proteome analysis. Electrophor. 2004;25:1327–1333. doi: 10.1002/elps.200305844. [DOI] [PubMed] [Google Scholar]
  • 52.Heukeshoven J, Dernick R. Increased sensitivity for coomassie staining of sodium dodecyl sulfate-polyacrylamide gels using PhastSystem development unit. Electrophor. 1988;9:60–61. doi: 10.1002/elps.1150090112. [DOI] [PubMed] [Google Scholar]
  • 53.Wong C, Sridhara S, Bardwell JCA, Jakob U. Heating greatly speeds coomassie blue staining and destaining. Biotech. 2000;28:426–432. doi: 10.2144/00283bm07. [DOI] [PubMed] [Google Scholar]
  • 54.Luo S, Wehr NB, Levine RL. Quantitation of protein on gels and blots by infrared fluorescence of coomassie blue and fast green. Anal Biochem. 2006;350:233–238. doi: 10.1016/j.ab.2005.10.048. [DOI] [PubMed] [Google Scholar]
  • 55.Wang X, Li X, Li Y. A modified Coomassie Brilliant Blue staining method at nanogram sensitivity compatible with proteomic analysis. Biotech Lett. 2007;29:1599–1603. doi: 10.1007/s10529-007-9425-3. [DOI] [PubMed] [Google Scholar]
  • 56.Choi J-K, Yoo G-S. Fast protein staining in sodium dodecyl sulfate polacrylamide gel using counter ion-dyes, coomassie brilliant blue R-250 and neutral red. Arch Pharm Res. 2002;25:704–708. doi: 10.1007/BF02976948. [DOI] [PubMed] [Google Scholar]
  • 57.Fernandez-Patron C, Hardy E, Sosa A, Seoane J, Castellanos L. Double staining of coomassie blue-stained polyacrylamide gels by imidazole-sodium dodecyl sulfate-zinc reverse staining: Sensitive detection of coomassie blue-undetected proteins. Anal Biochem. 1995;224:263–269. doi: 10.1006/abio.1995.1039. [DOI] [PubMed] [Google Scholar]
  • 58.Oakley BR, Kirsch DR, Morris NR. A simplified ultrasensitive silver stain for detecting proteins in polyacrylamide gels. Anal Biochem. 1980;105:361–363. doi: 10.1016/0003-2697(80)90470-4. [DOI] [PubMed] [Google Scholar]
  • 59.Merril CR, Goldman D, Sedman SA, Ebert MH. Ultrasensitive stain for proteins in polyacrylamide gels shows regional variation in cerebrospinal fluid proteins. Science. 1981;211:1437–1438. doi: 10.1126/science.6162199. [DOI] [PubMed] [Google Scholar]
  • 60.Morrissey JH. Silver stain for proteins in polyacrylamide gels: a modified procedure with enhanced uniform sensitivity. Anal Biochem. 1981;117:307–310. doi: 10.1016/0003-2697(81)90783-1. [DOI] [PubMed] [Google Scholar]
  • 61.Wray W, Boulikas T, Wray VP, Hancock R. Silver staining of proteins in polyacrylamide gels. Anal Biochem. 1981;118:197–203. doi: 10.1016/0003-2697(81)90179-2. [DOI] [PubMed] [Google Scholar]
  • 62.Merril CR, Dunau ML, Goldman D. A rapid sensitive silver stain for polypeptides in polyacrylamide gels. Anal Biochem. 1981;110:201–207. doi: 10.1016/0003-2697(81)90136-6. [DOI] [PubMed] [Google Scholar]
  • 63.Merril CR, Goldman D, Keuren ML. Simplified silver protein detection and image enhancement methods in polyacrylamide gels. Electrophor. 1982;3:17–23. [Google Scholar]
  • 64.Merril CR, Harrington MG, Alley V. A photodevelopment silver stain for the rapid visualization of proteins separated on polyacrylamide gels. Electrophor. 1984;5:289–297. [Google Scholar]
  • 65.Biel HJ, Gronski P, Seiler FR. Fast silver staining of polyacrylamide gels at elevated temperatures. Electrophor. 1986;7:232–235. [Google Scholar]
  • 66.Blum H, Beier H, Gross HJ. Improved silver staining of plant proteins, RNA and DNA in polyacrylamide gels. Electrophor. 1987;8:93–99. [Google Scholar]
  • 67.Shevchenko A, Wilm M, Vorm O, Mann M. Mass spectrometric sequencing of proteins from silver-stained polyacrylamide gels. Anal Chem. 1996;68:850–858. doi: 10.1021/ac950914h. [DOI] [PubMed] [Google Scholar]
  • 68.Yan JX, Wait R, Berkelman T, Harry RA, Westbrook JA, Wheeler CH, Dunn MJ. A modified silver staining protocol for visualization of proteins compatible with matrix-assisted laser desorption/ionization and electrospray ionization- mass spectrometry. Electrophor. 2000;21:3666–3672. doi: 10.1002/1522-2683(200011)21:17<3666::AID-ELPS3666>3.0.CO;2-6. [DOI] [PubMed] [Google Scholar]
  • 69.Sinha P, Poland J, Schnölzer M, Rabilloud T. A new silver staining apparatus and procedure for matrix-assisted laser desorption/ionization-time of flight analysis of proteins after two-dimensional electrophoresis. Proteomics. 2001;1:835–840. doi: 10.1002/1615-9861(200107)1:7<835::AID-PROT835>3.0.CO;2-2. [DOI] [PubMed] [Google Scholar]
  • 70.Sumner LW, Wolf-Sumner B, White SP, Asirvatham VS. Silver stain removal using H2O2 for enhanced peptide mass mapping by matrix-assisted laser desorption/ionization time-of-flight mass spectrometry. Rapid Commun Mass Spectrom. 2002;16:160–168. doi: 10.1002/rcm.555. [DOI] [PubMed] [Google Scholar]
  • 71.Richert S, Luche S, Chevallet M, Dorsselaer A, Leize-Wagner E, Rabilloud T. About the mechanism of interference of silver staining with peptide mass spectrometry. Proteomics. 2004;4:909–916. doi: 10.1002/pmic.200300642. [DOI] [PubMed] [Google Scholar]
  • 72.Guevara J, Johnston DA, Ramagali LS, Martin BA, Capetillo S, Rodriguez LV. Quantitative aspects of silver deposition in proteins resolved in complex polyacrylamide gels. Electrophor. 1982;3:197–205. [Google Scholar]
  • 73.Lopez MF, Berggren K, Chernokalskaya E, Lazarev A, Robinson MH, 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. Electrophor. 2000;21:3673–3683. doi: 10.1002/1522-2683(200011)21:17<3673::AID-ELPS3673>3.0.CO;2-M. [DOI] [PubMed] [Google Scholar]
  • 74.Mortz E, Krogh TN, Vorum H, Görg A. Improved silver staining protocols for high sensitivity protein identification using matrix-assisted laser desorption/ionization-time of flight analysis. Proteomics. 2001;1:1359–1363. doi: 10.1002/1615-9861(200111)1:11<1359::AID-PROT1359>3.0.CO;2-Q. [DOI] [PubMed] [Google Scholar]
  • 75.Irie S, Sezaki M, Kato Y. A faithful double stain of proteins in the polyacrylamide gels with coomassie blue and silver. Anal Biochem. 1982;126:350–354. doi: 10.1016/0003-2697(82)90526-7. [DOI] [PubMed] [Google Scholar]
  • 76.Berson G. Silver staining of proteins in polyacrylamide gels: increased sensitivity by a blue toning. Anal Biochem. 1983;134:230–234. doi: 10.1016/0003-2697(83)90289-0. [DOI] [PubMed] [Google Scholar]
  • 77.Dzandu JK, Deh ME, Barratt DL, Wise GE. Detection of erythrocyte membrane proteins, sialoglycoproteins, and lipids in the same polyacrylamide gel using a double-staining technique. Proc Natl Acad Sci USA. 1984;81:1733–1737. doi: 10.1073/pnas.81.6.1733. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 78.Moreno MR, Smith JF, Smith RV. Silver staining of proteins in polyacrylamide gels: Increased sensitivity through a combined coomassie blue-silver stain procedure. Anal Biochem. 1985;151:466–470. doi: 10.1016/0003-2697(85)90206-4. [DOI] [PubMed] [Google Scholar]
  • 79.Ross M, Peters L. Double staining to increase the sensitivity of protein detection in polyacrylamide gels. Biotech. 1990;9:532–533. [PubMed] [Google Scholar]
  • 80.Hänsler M, Appelt G, Rogos R. Improved detection of human pancreatic proteins separated by two-dimensional isoelectric focusing/sodium dodecyl sulphate gel electrophoresis using a combined staining procedure. J Chromatogr-Biomedical Applications. 1991;563:63–71. doi: 10.1016/0378-4347(91)80277-j. [DOI] [PubMed] [Google Scholar]
  • 81.Goldberg HA, Warner KJ. The staining of acidic proteins on polyacrylamide gels: Enhanced sensitivity and stability of "Stains-All" staining in combination with silver nitrate. Anal Biochem. 1997;251:227–233. doi: 10.1006/abio.1997.2252. [DOI] [PubMed] [Google Scholar]
  • 82.Zhou Z-D, Liu W-Y, Li M-Q. Enhanced thiosulfate-silver staining of proteins in gels using coomassie blue and chromium. Biotech Lett. 2002;24:1561–1567. [Google Scholar]
  • 83.Lauber WM, Carroll JA, Dufield DR, Kiesel JR, Radabaugh MR, Malone JP. Mass spectrometry compatibility of two-dimensional gel protein stains. Electrophor. 2001;22:906–918. doi: 10.1002/1522-2683()22:5<906::AID-ELPS906>3.0.CO;2-9. [DOI] [PubMed] [Google Scholar]
  • 84.Scheler C, Lamer S, Pan Z, Li X-P, Salnikow J, Jungblut P. Peptide mass fingerprint sequence coverage from differently stained proteins on two-dimensional electrophoresis patterns by matrix assisted laser desorption/ionization-mass spectrometry (MALDI-MS) Electrophor. 1998;19:918–927. doi: 10.1002/elps.1150190607. [DOI] [PubMed] [Google Scholar]
  • 85.Dion AS, Pomenti AA. Ammoniacal silver staining of proteins: Mechanism of glutaraldehyde enhancement. Anal Biochem. 1983;129:490–496. doi: 10.1016/0003-2697(83)90582-1. [DOI] [PubMed] [Google Scholar]
  • 86.Chevallet M, Diemer H, Luche S, Van Dorsselaer A, Rabilloud T, Leize-Wagner E. Improved mass spectrometry compatibility is afforded by ammoniacal silver staining. Proteomics. 2006;6:2350–2354. doi: 10.1002/pmic.200500567. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 87.Chevallet M, Luche S, Diemer H, Strub J-M, Van Dorsselaer A, Rabilloud T. Sweet silver: a formaldehyde-free silver staining using aldoses as developing agents, with enhanced compatibility with mass spectrometry. Proteomics. 2008;8:4853–4861. doi: 10.1002/pmic.200800321. [DOI] [PubMed] [Google Scholar]
  • 88.Jin L-T, Hwang S-Y, Yoo G-S, Choi J-K. Sensitive silver staining of protein in sodium dodecyl sulfate-polyacrylamide gels using an azo dye, calconcarboxylic acid, as a silver-ion sensitizer. Electropho. 2004;25:2494–2500. doi: 10.1002/elps.200306002. [DOI] [PubMed] [Google Scholar]
  • 89.Jin L-T, Hwang S-Y, Yoo G-S, Choi J-K. A mass spectrometry compatible silver staining method for protein incorporating a new silver sensitizer in sodium dodecyl sulfate-polyacrylamide electrophoresis gels. Proteomics. 2006;6:2334–2337. doi: 10.1002/pmic.200500596. [DOI] [PubMed] [Google Scholar]
  • 90.Jin L-T, Li X-K, Cong W-T, Hwang S-Y, Choi J-K. Previsible silver staining of protein in electrophoresis gels with mass spectrometry compatibility. Anal Biochem. 2008;383:137–143. doi: 10.1016/j.ab.2008.04.048. [DOI] [PubMed] [Google Scholar]
  • 91.Tsai C-M, Frasch CE. A sensitive silver stain for detecting lipopolysaccharides in polyacrylamide gels. Anal Biochem. 1982;119:115–119. doi: 10.1016/0003-2697(82)90673-x. [DOI] [PubMed] [Google Scholar]
  • 92.Igloi GL. A silver stain for the detection of nanogram amounts of tRNA following two-dimensional electrophoresis. Anal Biochem. 1983;134:184–188. doi: 10.1016/0003-2697(83)90281-6. [DOI] [PubMed] [Google Scholar]
  • 93.Guillemette JG, Lewis PN. Detection of subnanogram quantities of DNA and RNA on native and denaturing polyacrylamide and agarose gels by silver staining. Electrophor. 1983;4:92–94. [Google Scholar]
  • 94.Paleologue A, Reboud JP, Reboud AM. Selective silver-staining methods for RNA and proteins in the same polyacrylamide gels. Anal Biochem. 1988;169:234–238. doi: 10.1016/0003-2697(88)90279-5. [DOI] [PubMed] [Google Scholar]
  • 95.Friedman RD. Comparison of four different silver-staining techniques for salivary protein detection in alkaline polyacrylamide gels. Anal Biochem. 1982;126:346–349. doi: 10.1016/0003-2697(82)90525-5. [DOI] [PubMed] [Google Scholar]
  • 96.Higgins RC, Dahmus ME. Rapid visualization of protein bands in preparative SDS-polyacrylamide gels. Anal Biochem. 1979;93:257–260. doi: 10.1016/s0003-2697(79)80148-7. [DOI] [PubMed] [Google Scholar]
  • 97.LeBlanc GA, Cochrane BJ. A rapid method for staining proteins in acrylamide gels. Anal Biochem. 1987;161:172–175. doi: 10.1016/0003-2697(87)90668-3. [DOI] [PubMed] [Google Scholar]
  • 98.Lee C, Levin A, Branton D. Copper staining: a five-minute protein stain for sodium dodecyl sulfate-polyacrylamide gels. Anal Biochem. 1987;166:308–312. doi: 10.1016/0003-2697(87)90579-3. [DOI] [PubMed] [Google Scholar]
  • 99.Dzandu JK, Johnson JF, Wise GE. Sodium dodecyl sulfate-gel electrophoresis: staining of polypeptides using heavy metal salts. Anal Biochem. 1988;174:157–167. doi: 10.1016/0003-2697(88)90531-3. [DOI] [PubMed] [Google Scholar]
  • 100.Adams LD, Weaver KM. Detection and recovery of proteins from gels following zinc chloride staining. Appl Theor Electrophor. 1990;1:279–282. [PubMed] [Google Scholar]
  • 101.Fernandez-Patron C, Castellanos L, Rodriguez P. Reverse staining of sodium dodecyl sulfate polyacrylamide gels by imidazole-zinc salts: sensitive detection of unmodified proteins. Biotech. 1992;12:564–573. [PubMed] [Google Scholar]
  • 102.Ortiz ML, Calero M, Fernandez-Patron C, Castellanos L, Mendez E. Imidazole-SDS-Zn reverse staining of proteins in gels containing or not SDS and microsequence of individual unmodified electroblotted proteins. FEBS Lett. 1992;296:300–304. doi: 10.1016/0014-5793(92)80309-5. [DOI] [PubMed] [Google Scholar]
  • 103.Ferreras M, Gavilanes JG, Garciasegura JM. A permanent Zn2+ reverse staining method for the detection and quantification of proteins in polyacrylamide gels. Anal Biochem. 1993;213:206–212. doi: 10.1006/abio.1993.1410. [DOI] [PubMed] [Google Scholar]
  • 104.Fernandez-Patron C, Calero M, Collazo PR, Garcia JR, Madrazo J, Musacchio A, Soriano F, Estrada R, Frank R, Castellanos-Serra LR, Mendez E. Protein reverse staining: high-efficiency microanalysis of unmodified proteins detected on electrophoresis gels. Anal Biochem. 1995;224:203–211. doi: 10.1006/abio.1995.1031. [DOI] [PubMed] [Google Scholar]
  • 105.Castellanos-Serra LR, Fernandez-Patron C, Hardy E, Huerta V. A procedure for protein elution from reverse-stained polyacrylamide gels applicable at the low picomole level: an alternative route to the preparation of low abundance proteins for microanalysis. Electrophor. 1996;17:1564–1572. doi: 10.1002/elps.1150171012. [DOI] [PubMed] [Google Scholar]
  • 106.Hardy E, Santana H, Sosa A, Hernández L, Fernández-Patrón C, Castellanos-Serra L. Recovery of biologically active proteins detected with imidazole-sodium dodecyl sulfate-zinc (reverse stain) on sodium dodecyl sulfate gels. Anal Biochem. 1996;240:150–152. doi: 10.1006/abio.1996.0343. [DOI] [PubMed] [Google Scholar]
  • 107.Castellanos-Serra LR, Fernandez-Patron C, Hardy E, Santana H, Huerta V. High yield elution of proteins from sodium dodecyl sulfate–polyacrylamide gels at the low-picomole level. Application to N-terminal sequencing of a scarce protein and to in-solution biological activity analysis of on-gel renatured proteins. J Protein Chem. 1997;16:415–419. doi: 10.1023/a:1026340923032. [DOI] [PubMed] [Google Scholar]
  • 108.Matsui NM, Smith DM, Clauser KR, Fichmann J, Andrews LE, Sullivan CM, Burlingame AL, Epstein LB. Immobilized pH gradient two-dimensional gel electrophoresis and mass spectrometric identification of cytokine-regulated proteins in ME-180 cervical carcinoma cells. Electrophor. 1997;18:409–417. doi: 10.1002/elps.1150180315. [DOI] [PubMed] [Google Scholar]
  • 109.Castellanos-Serra LR, Proenza W, Huerta V, Moritz RL, Simpson RJ. Proteome analysis of polyacrylamide gel-separated proteins visualized by reversible negative staining using imidazole-zinc salts. Electrophor. 1999;20:732–737. doi: 10.1002/(SICI)1522-2683(19990101)20:4/5<732::AID-ELPS732>3.0.CO;2-Q. [DOI] [PubMed] [Google Scholar]
  • 110.Castellanos-Serra LR, Vallin A, Proenza W, Caer JP, Rossier J. An optimized procedure for detection of proteins on carrier ampholyte isoelectric focusing and immobilized pH gradient gels with imidazole and zinc salts: Its application to the identification of isoelectric focusing separated isoforms by in-gel proteolysis and mass spectrometry analysis. Electrophor. 2001;22:1677–1685. doi: 10.1002/1522-2683(200105)22:9<1677::AID-ELPS1677>3.0.CO;2-H. [DOI] [PubMed] [Google Scholar]
  • 111.Fernandez-Patron C, Castellanos-Serra L, Hardy E, Guerra M, Estevez E, Mehl E, Frank RW. Understanding the mechanism of the zinc-ion stains of biomacromolecules in electrophoresis gels: generalization of the reverse-staining technique. Electrophor. 1998;19:2398–2406. doi: 10.1002/elps.1150191407. [DOI] [PubMed] [Google Scholar]
  • 112.Hardy E, Pupo E, Casalvilla R, Sosa AE, Trujillo LE, López E, Castellanos-Serra L. Negative staining with zinc-imidazole of gel electrophoresis-separated nucleic acids. Electrophor. 1996;17:1537–1541. doi: 10.1002/elps.1150171006. [DOI] [PubMed] [Google Scholar]
  • 113.Hardy E, Pupo E, Castellanos-Serra L, Reyes J, Fernández-Patrón C. Sensitive reverse staining of bacterial lipopolysaccharides on polyacrylamide gels by using zinc and imidazole salts. Anal Biochem. 1997;244:28–32. doi: 10.1006/abio.1996.9719. [DOI] [PubMed] [Google Scholar]
  • 114.Bosshard HF, Datyner A. The use of a new reactive dye for quantitation of prestained proteins on polyacrylamide gels. Anal Biochem. 1977;82:327–333. doi: 10.1016/0003-2697(77)90168-3. [DOI] [PubMed] [Google Scholar]
  • 115.Telser A, Rovin B. A specific stain for sulfhydryl groups in proteins after separation by polyacrylamide gel electrophoresis. Biochim Biophys Acta-Protein Structure. 1980;624:363–371. doi: 10.1016/0005-2795(80)90077-x. [DOI] [PubMed] [Google Scholar]
  • 116.Choi J-K, Tak K-H, Jin L-T, Hwang S-Y, Kwon T-I, Yoo G-S. Background-free, fast protein staining in sodium dodecyl sulfate polyacrylamide gel using counterion dyes, zincon and ethyl violet. Electrophor. 2002;23:4053–4059. doi: 10.1002/elps.200290020. [DOI] [PubMed] [Google Scholar]
  • 117.Gorovsky MA, Carlson K, Rosenbaum JL. Simple method for quantitive densitometry of polyacrylamide gels using fast green. Anal Biochem. 1970;35:359–370. doi: 10.1016/0003-2697(70)90196-x. [DOI] [PubMed] [Google Scholar]
  • 118.Allen RE, Masak KC, McAllister PK. Staining protein in isoelectric focusing gels with fast green. Anal Biochem. 1980;104:494–498. doi: 10.1016/0003-2697(80)90106-2. [DOI] [PubMed] [Google Scholar]
  • 119.Selsted ME, Becker HW. Eosin Y: A reversible stain for detecting electrophoretically resolved protein. Anal Biochem. 1986;155:270–274. doi: 10.1016/0003-2697(86)90436-7. [DOI] [PubMed] [Google Scholar]
  • 120.Lin F, Fan W, Wise GE. Eosin Y staining of proteins in polyacrylamide gels. Anal Biochem. 1991;196:279–283. doi: 10.1016/0003-2697(91)90466-7. [DOI] [PubMed] [Google Scholar]
  • 121.Pal JK, Godbole D, Sharma K. Staining of proteins on SDS polyacrylamide gels and on nitrocellulose membranes by Alta, a colour used as a cosmetic. J Biochem Biophys Meth. 2004;61:339–347. doi: 10.1016/j.jbbm.2004.06.008. [DOI] [PubMed] [Google Scholar]
  • 122.Hong HY, Choi JK, Yoo GS. Detection of proteins on polyacrylamide gels using calconcarboxylic acid. Anal Biochem. 1993;214:96–99. doi: 10.1006/abio.1993.1461. [DOI] [PubMed] [Google Scholar]
  • 123.Ali R, Sayeed SA, Khan AA. A sensitive novel staining agent for the resolved proteins in PAGE. Intl J Pept Protein Res. 1995;45:97–99. doi: 10.1111/j.1399-3011.1995.tb01026.x. [DOI] [PubMed] [Google Scholar]
  • 124.Jung D-W, Yoo G-S, Choi J-K. Mixed-dye staining method for protein detection in polyacrylamide gel electrophoresis using calconcarboxylic acid and rhodamine B. Electrophor. 1998;19:2412–2415. doi: 10.1002/elps.1150191409. [DOI] [PubMed] [Google Scholar]
  • 125.Jung D-W, Tak G-H, Lim J-W, Bae C-J, Kim G-Y, Yoo G-S, Choi J-K. Detection of proteins in polyacrylamide gels using eriochrome black T and rhodamine B. Anal Biochem. 1998;263:118–120. doi: 10.1006/abio.1998.2813. [DOI] [PubMed] [Google Scholar]
  • 126.Bickar D, Reid PD. A high-affinity protein stain for western blots, tissue prints, and electrophoretic gels. Anal Biochem. 1992;203:109–115. doi: 10.1016/0003-2697(92)90049-d. [DOI] [PubMed] [Google Scholar]
  • 127.Goldys EM, editor. Fluorescence Applications in Biotechnology and the Life Sciences. Wiley Blackwell: Hoboken; 2009. [DOI] [PubMed] [Google Scholar]
  • 128.Beeley JA, Newman F, Wilson PHR, Shimmin IC. Soldium dodecyl sulphate-polyacrylamide gel electrophoresis of human parotid salivary proteins: comparison of dansylation, coomassie blue R-250 and silver detection methods. Electrophor. 1996;17:505–506. doi: 10.1002/elps.1150170314. [DOI] [PubMed] [Google Scholar]
  • 129.Parkinson D, Redshaw JD. Visible labeling of proteins for polyacrylamide gel electrophoresis with dabsyl chloride. Anal Biochem. 1984;141:121–126. doi: 10.1016/0003-2697(84)90434-2. [DOI] [PubMed] [Google Scholar]
  • 130.Griffith IP. Immediate visualization of proteins in dodecyl sulfate-polyacrylamide gels by prestaining with remazol dyes. Anal Biochem. 1972;46:402–412. doi: 10.1016/0003-2697(72)90313-2. [DOI] [PubMed] [Google Scholar]
  • 131.Holmes KL, Lantz LM. Protein labeling with fluorescent probes. In: Darzynkiewicz Z, editor. In Cytometry. Bethesda: Elsevier; 2001. pp. 194–195. [DOI] [PubMed] [Google Scholar]
  • 132.Shaw J, Rowlinson R, Nickson J, Stone T, Sweet A, Williams K, Tonge R. Evaluation of saturation labelling two-dimensional difference gel electrophoresis fluorescent dyes. Proteomics. 2003;3:1181–1195. doi: 10.1002/pmic.200300439. [DOI] [PubMed] [Google Scholar]
  • 133.Righetti PG, Castagna A, Antonucci F, Piubelli C, Cecconi D, Campostrini N, Antonioli P, Astner H, Hamdan M. Critical survey of quantitative proteomics in two-dimensional electrophoretic approaches. J Chromatogr A. 2004;1051:3–17. doi: 10.1016/j.chroma.2004.05.106. [DOI] [PubMed] [Google Scholar]
  • 134.Kelleher NL, Lin HY, Valaskovic GA, Aaserud DJ, Fridriksson EK, McLafferty FW. Top down versus bottom up protein characterization by tandem high-resolution mass spectrometry. J Am Chem Soc. 1999;121:806–812. [Google Scholar]
  • 135.England K, Driscoll CO, Cotter TG. ROS and protein oxidation in early stages of cytotoxic drug induced apoptosis. Free Radic Res. 2006;40:1124–1137. doi: 10.1080/10715760600838209. [DOI] [PubMed] [Google Scholar]
  • 136.Baar BLM, Hulst AG, Jong AL, Wils ERJ. Characterisation of botulinum toxins type A and B, by matrix-assisted laser desorption ionisation and electrospray mass spectrometry. J Chromatogr A. 2002;970:95–115. doi: 10.1016/s0021-9673(02)00508-3. [DOI] [PubMed] [Google Scholar]
  • 137.Barger BO, White FC, Pace JL, Kemper DL, Ragland WL. Estimation of molecular weight by polyacrylamide gel electrophoresis using heat stable fluorophors. Anal Biochem. 1976;70:327–335. doi: 10.1016/0003-2697(76)90453-x. [DOI] [PubMed] [Google Scholar]
  • 138.Jackson P, Urwin VE, MacKay CD. Rapid imaging, using a cooled charge-coupled-device, of fluorescent two-dimensional polyacrylamide gels produced by labelling proteins in the first-dimensional isoelectric focusing gel with the fluorophore 2-methoxy-2,4-diphenyl-3(2H)furanone. Electrophor. 1988;9:330–339. doi: 10.1002/elps.1150090709. [DOI] [PubMed] [Google Scholar]
  • 139.Weidekamm E, Wallach DFH, Flückiger R. A new sensitive, rapid fluorescence technique for the determination of proteins in gel electrophoresis and in solution. Anal Biochem. 1973;54:102–114. doi: 10.1016/0003-2697(73)90252-2. [DOI] [PubMed] [Google Scholar]
  • 140.Ragland WL, Pace JL, Kemper DL. Fluorometric scanning of fluorescamine-labeled proteins in polyacrylamide gels. Anal Biochem. 1974;59:24–33. doi: 10.1016/0003-2697(74)90005-0. [DOI] [PubMed] [Google Scholar]
  • 141.Jackowski G, Liew CC. Fluorescamine staining of nonhistone chromatin proteins as revealed by two-dimensional polyacrylamide gel electrophoresis. Anal Biochem. 1980;102:321–325. doi: 10.1016/0003-2697(80)90161-x. [DOI] [PubMed] [Google Scholar]
  • 142.Tonge R, Shaw J, Middleton B, Rowlinson R, Rayner S, Young J, Pognan F, Hawkins E, Currie I, Davison M. Validation and development of fluorescence two-dimensional differential gel electrophoresis proteomics technology. Proteomics. 2001;1:377–396. doi: 10.1002/1615-9861(200103)1:3<377::AID-PROT377>3.0.CO;2-6. [DOI] [PubMed] [Google Scholar]
  • 143.Berlier JE, Rothe A, Buller G, Bradford J, Gray DR, Filanoski BJ, Telford WG, Yue S, Liu J, Cheung C-Y, Chang W, Hirsch JD, Beechem JM, Haughland RP, Haugland RP. Quantitative comparison of long-wavelength alexa fluor dyes to Cy dyes: fluorescence of the dyes and their bioconjugates. J Histochem Cytochem. 2003;51:1699–1712. doi: 10.1177/002215540305101214. [DOI] [PubMed] [Google Scholar]
  • 144.Corzett TH, Fodor IK, Choi MW, Walsworth VL, Chromy BA, Turteltaub KW, McCutchen-Maloney SL. Statistical analysis of the experimental variation in the proteomic characterization of human plasma by two-dimensional difference gel electrophoresis. J Prot Res. 2006;5:2611–2619. doi: 10.1021/pr060100p. [DOI] [PubMed] [Google Scholar]
  • 145.Zhou G, Li H, DeCamp D, Chen S, Shu H, Gong Y, Flaig M, Gillespie JW, Hu N, Taylor PR, Emmert-Buck MR, Liotta LA, Petricoin EF, III, Zhao Y. 2D differential in-gel electrophoresis for the identification of esophageal scans cell cancer-specific protein markers. Mol Cell Proteomics. 2002;1:117–123. doi: 10.1074/mcp.m100015-mcp200. [DOI] [PubMed] [Google Scholar]
  • 146.Tyagarajan K, Pretzer E, Wiktorowicz JE. Thiol-reactive dyes for fluorescence labeling of proteomic samples. Electrophor. 2003;24:2348–2358. doi: 10.1002/elps.200305478. [DOI] [PubMed] [Google Scholar]
  • 147.Chan H-L, Gharbi S, Gaffney PR, Cramer R, Waterfield MD, Timms JF. Proteomic analysis of redox- and ErbB2-dependent changes in mammary luminal epithelial cells using cysteine- and lysine-labelling two-dimensional difference gel electrophoresis. Proteomics. 2005;5:2908–2926. doi: 10.1002/pmic.200401300. [DOI] [PubMed] [Google Scholar]
  • 148.Pretzer E, Wiktorowicz JE. Saturation fluorescence labeling of proteins for proteomic analyses. Anal Biochem. 2008;374:250–262. doi: 10.1016/j.ab.2007.12.014. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 149.O'Keefe DO. Quantitative electrophoretic analysis of proteins labeled with monobromobimane. Anal Biochem. 1994;222:86–94. doi: 10.1006/abio.1994.1458. [DOI] [PubMed] [Google Scholar]
  • 150.Urwin VE, Jackson P. Two-dimensional polyacrylamide gel electrophoresis of proteins labeled with the fluorophore monobromobimane prior to first-dimensional isoelectric focusing: Imaging of the fluorescent protein spot patterns using a cooled charge-coupled device. Anal Biochem. 1993;209:57–62. doi: 10.1006/abio.1993.1082. [DOI] [PubMed] [Google Scholar]
  • 151.Swatton JE, Prabakaran S, Karp NA, Lilley KS, Bahn S. Protein profiling of human postmortem brain using 2-dimensional fluorescence difference gel electrophoresis (2-D DIGE) Mol Psychiatry. 2004;9:128–143. doi: 10.1038/sj.mp.4001475. [DOI] [PubMed] [Google Scholar]
  • 152.Sitek B, Lüttges J, Marcus K, Klöppel G, Schmiegel W, Meyer HE, Hahn SA, Stühler K. Application of fluorescence difference gel electrophoresis saturation labelling for the analysis of microdissected precursor lesions of pancreatic ductal adenocarcinoma. Proteomics. 2005;5:2665–2679. doi: 10.1002/pmic.200401298. [DOI] [PubMed] [Google Scholar]
  • 153.Riederer IM, Riederer BM. Differential protein labeling with thiol-reactive infrared DY-680 and DY-780 maleimides and analysis by two-dimensional gel electrophoresis. Proteomics. 2007;7:1753–1756. doi: 10.1002/pmic.200601007. [DOI] [PubMed] [Google Scholar]
  • 154.Dietz L, Bosque A, Pankert P, Ohnesorge S, Merz P, Anel A, Schnölzer M, Thierse H-J. Quantitative DY-maleimide-based proteomic 2-DE-labeling strategies using human skin proteins. Proteomics. 2009;9:4298–4308. doi: 10.1002/pmic.200900051. [DOI] [PubMed] [Google Scholar]
  • 155.Tsolakos N, Techanukul T, Wallington A, Zhao Y, Jones C, Nagy J, Wheeler JX. Comparison of two combinations of cyanine dyes for prelabelling and gel electrophoresis. Proteomics. 2009;9:1727–1730. doi: 10.1002/pmic.200800563. [DOI] [PubMed] [Google Scholar]
  • 156.Nishihara JC, Champion KM. Quantitative evaluation of proteins in one- and two-dimensional polyacrylamide gels using a fluorescent stain. Electrophor. 2002;23:2203–2215. doi: 10.1002/1522-2683(200207)23:14<2203::AID-ELPS2203>3.0.CO;2-H. [DOI] [PubMed] [Google Scholar]
  • 157.Cong W-T, Hwang S-Y, Jin L-T, Choi J-K. Improved conditions for fluorescent staining of proteins with 4, 4′-dianilino-1, 1′-binaphthyl-5, 5′-disulfonic acid in SDS-PAGE. Electrophor. 2008;29:4487–4494. doi: 10.1002/elps.200800124. [DOI] [PubMed] [Google Scholar]
  • 158.Cong W-T, Hwang S-Y, Jin L-T, Choi J-K. Sensitive fluorescent staining for proteomic analysis of proteins in 1-D and 2-D SDS-PAGE and its comparison with SYPRO Ruby by PMF. Electrophor. 2008;29:4304–4315. doi: 10.1002/elps.200800150. [DOI] [PubMed] [Google Scholar]
  • 159.Ball MS, Karuso P. Mass spectral compatibility of four proteomics stains. J Proteome Res. 2007;6:4313–4320. doi: 10.1021/pr070398z. [DOI] [PubMed] [Google Scholar]
  • 160.Kreig RC, Paweletz CP, Liotta LA, Petricon EF., III Dilution of protein gel stain results in retention of staining capacity. Biotech. 2003;35:376–378. doi: 10.2144/03352pt01. [DOI] [PubMed] [Google Scholar]
  • 161.Ahnert N, Patton WF, Schulenberg B. Optimized conditions for diluting and reusing a fluorescent protein gel stain. Electrophor. 2004;25:2506–2510. doi: 10.1002/elps.200406004. [DOI] [PubMed] [Google Scholar]
  • 162.Rabilloud T, Strub J-M, Luche S, Girardet JL, Dorsselaer A, Lunardi J. Ruthenium II tris (bathophenanthroline disulfonate), a powerful fluorescent stain for detection of proteins in gel with minimal interference in subsequent mass spectrometry analysis. Proteome. 2000;1:1–14. [Google Scholar]
  • 163.Rabilloud T, Strub J-M, Luche S, Van Dorsselaer A, Lunardi J. A comparison between Sypro Ruby and ruthenium II tris (bathophenanthroline disulfonate) as fluorescent stains for protein detection in gels. Proteomics. 2001;1:699–704. doi: 10.1002/1615-9861(200104)1:5<699::AID-PROT699>3.0.CO;2-C. [DOI] [PubMed] [Google Scholar]
  • 164.Berggren KN, Schulenberg B, Lopez MF, Steinberg TH, Bogdanova A, Smejkal G, Wang A, Patton WF. An improved formulation of SYPRO Ruby protein gel stain: comparison with the original formulation and with a ruthenium II tris (bathophenanthroline disulfonate) formulation. Proteomics. 2002;2:486–498. doi: 10.1002/1615-9861(200205)2:5<486::AID-PROT486>3.0.CO;2-X. [DOI] [PubMed] [Google Scholar]
  • 165.Francis F, Gerkens P, Harmel N, Mazzucchelli G, Pauw E, Haubruge E. Proteomics in Myzus persicae: effect of aphid host plant switch. Insect Biochem Mol Biol. 2006;36:219–227. doi: 10.1016/j.ibmb.2006.01.018. [DOI] [PubMed] [Google Scholar]
  • 166.Gerber IB, Laukens K, Witters E, Dubery IA. Lipopolysaccharide-responsive phosphoproteins in Nicotiana tabacum cells. Plant Physiol Biochem. 2006;44:369–379. doi: 10.1016/j.plaphy.2006.06.015. [DOI] [PubMed] [Google Scholar]
  • 167.Rinnerthaler M, Jarolim S, Heeren G, Palle E, Perju S, Klinger H, Bogengruber E, Madeo F, Braun RJ, Breitenbach-Koller L, Breitenbach M, Laun P. MMI1 (YKL056c, TMA19), the yeast orthologue of the translationally controlled tumor protein (TCTP) has apoptotic functions and interacts with both microtubules and mitochondria. Biochim Biophys Acta-Bioenergetics. 2006;1757:631–638. doi: 10.1016/j.bbabio.2006.05.022. [DOI] [PubMed] [Google Scholar]
  • 168.Berger K, Wissmann D, Ihling C, Kalkhof S, Beck-Sickinger A, Sinz A, Paschke R, Führer D. Quantitative proteome analysis in benign thyroid nodular disease using the fluorescent ruthenium II tris(bathophenanthroline disulfonate) stain. Mol Cell Endocrinol. 2004;227:21–30. doi: 10.1016/j.mce.2004.08.001. [DOI] [PubMed] [Google Scholar]
  • 169.Krause K, Karger S, Schierhorn A, Poncin S, Many M-C, Fuhrer D. Proteomic profiling of cold thyroid nodules. Endocrinol. 2007;148:1754–1763. doi: 10.1210/en.2006-0752. [DOI] [PubMed] [Google Scholar]
  • 170.Witzel K, Surabhi G-K, Jyothsnakumari G, Sudhakar C, Matros A, Mock H-P. Quantitative proteome analysis of barley seeds using ruthenium(II)-tris-(bathophenanthroline-disulphonate) staining. J Proteome Res. 2007;6:1325–1333. doi: 10.1021/pr060528o. [DOI] [PubMed] [Google Scholar]
  • 171.Tokarski C, Cren-Olivé C, Fillet M, Rolando C. High-sensitivity staining of proteins for one- and two-dimensional gel electrophoresis using post migration covalent staining with a ruthenium fluorophore. Electrophor. 2006;27:1407–1416. doi: 10.1002/elps.200500426. [DOI] [PubMed] [Google Scholar]
  • 172.Bell PJL, Karuso P. Epicocconone, a novel fluorescent compound from the fungus Epicoccum nigrum. J Am Chem Soc. 2003;125:9304–9305. doi: 10.1021/ja035496+. [DOI] [PubMed] [Google Scholar]
  • 173.Mackintosh JA, Choi H-Y, Bae S-H, Veal DA, Bell PJ, Ferrari BC, Dyk DD, Verrills NM, Paik Y-K, Karuso P. A fluorescent natural product for ultra sensitive detection of proteins in one-dimensional and two-dimensional gel electrophoresis. Proteomics. 2003;3:2273–2288. doi: 10.1002/pmic.200300578. [DOI] [PubMed] [Google Scholar]
  • 174.Svensson E, Hedberg JJ, Malmport E, Bjellqvist B. Fluorescent in-gel protein detection by regulating the pH during staining. Anal Biochem. 2006;355:304–306. doi: 10.1016/j.ab.2006.03.046. [DOI] [PubMed] [Google Scholar]
  • 175.Tannu NS, Sanchez-Brambila G, Kirby P, Andacht TM. Effect of staining reagent on peptide mass fingerprinting from in-gel trypsin digestions: a comparison of SyproRuby and DeepPurple. Electrophor. 2006;27:3136–3143. doi: 10.1002/elps.200500740. [DOI] [PubMed] [Google Scholar]
  • 176.Volkova K, Kovalska V, Yarmoluk S. Modern techniques for protein detection on polyacrylamide gels: problems arising from the use of dyes of undisclosed structures for scientific purposes. Biotech Histochem. 2007;82:201–208. doi: 10.1080/10520290701707660. [DOI] [PubMed] [Google Scholar]
  • 177.Nock CM, Ball MS, White IR, Skehel JM, Bill L, Karuso P. Mass spectrometric compatibility of Deep Purple and SYPRO Ruby total protein stains for high-throughput proteomics using large-format two-dimensional gel electrophoresis. Rapid Commun Mass Spectrom. 2008;22:881–886. doi: 10.1002/rcm.3483. [DOI] [PubMed] [Google Scholar]
  • 178.Smejkal GB, Robinson MH, Lazarev A. Comparison of fluorescent stains: relative photostability and differential staining of proteins in two-dimensional gels. Electrophor. 2004;25:2511–2519. doi: 10.1002/elps.200406005. [DOI] [PubMed] [Google Scholar]
  • 179.Coghlan DR, Mackintosh JA, Karuso P. Mechanism of reversible fluorescent staining of protein with epicocconone. Org Lett. 2005;7:2401–2404. doi: 10.1021/ol050665b. [DOI] [PubMed] [Google Scholar]
  • 180.Kang C, Kim HJ, Kang D, Jung DY, Suh M. Highly sensitive and simple fluorescence staining of proteins in sodium dodecyl sulfate-polyacrylamide-based gels by using hydrophobic tail-mediated enhancement of fluorescein luminescence. Electrophor. 2003;24:3297–3304. doi: 10.1002/elps.200305605. [DOI] [PubMed] [Google Scholar]
  • 181.Hartman BK, Udenfriend S. A method for immediate visualization of proteins in acrylamide gels and its use for preparation of antibodies to enzymes. Anal Biochem. 1969;30:391–394. doi: 10.1016/0003-2697(69)90132-8. [DOI] [PubMed] [Google Scholar]
  • 182.Daban J-R, Aragay AM. Rapid fluorescent staining of histones in sodium dodecyl sulfate-polyacrylamide gels. Anal Biochem. 1984;138:223–228. doi: 10.1016/0003-2697(84)90792-9. [DOI] [PubMed] [Google Scholar]
  • 183.Vincent A, Scherrer K. A rapid and sensitive method for detection of proteins in polyacrylamide SDS gels: Staining with ethidium bromide. Mol Biol Rep. 1979;5:209–214. doi: 10.1007/BF00782890. [DOI] [PubMed] [Google Scholar]
  • 184.Horowitz PM, Bowman S. Ion-enhanced fluorescence staining of sodium dodecyl sulfate-polyacrylamide gels using bis(8-p-toluidino-1-naphthalenesulfonate) Anal Biochem. 1987;165:430–434. doi: 10.1016/0003-2697(87)90292-2. [DOI] [PubMed] [Google Scholar]
  • 185.Kazmin D, Edwards RA, Turner RJ, Larson E, Starkey J. Visualization of proteins in acrylamide gels using ultraviolet illumination. Anal Biochem. 2002;301:91–96. doi: 10.1006/abio.2001.5488. [DOI] [PubMed] [Google Scholar]
  • 186.Daban J-R, Bartolomé S, Samsó M. Use of the hydrophobic probe Nile red for the fluorescent staining of protein bands in sodium dodecyl sulfate-polyacrylamide gels. Anal Biochem. 1991;199:169–174. doi: 10.1016/0003-2697(91)90085-8. [DOI] [PubMed] [Google Scholar]
  • 187.Suzuki Y, Yokoyama K, Namatame I. Rapid and easy protein staining for SDS-PAGE using an intramolecular charge transfer-based fluorescent reagent. Electrophor. 2006;27:3332–3337. doi: 10.1002/elps.200600101. [DOI] [PubMed] [Google Scholar]
  • 188.Dvortsov IA, Lunina NA, Chekanovskaya LA, Shedova EN, Gening LV, Velikodvorskaya GA. Ethidium bromide is good not only for staining of nucleic acids but also for staining of proteins after polyacrylamide gel soaking in trichloroacetic acid solution. Anal Biochem. 2006;353:293–295. doi: 10.1016/j.ab.2006.03.001. [DOI] [PubMed] [Google Scholar]
  • 189.Pluder F, Beck-Sickinger AG. One-step procedure for staining of proteins with ruthenium II tris (bathophenanthroline disulfonate) Anal Biochem. 2007;361:299–301. doi: 10.1016/j.ab.2006.11.003. [DOI] [PubMed] [Google Scholar]
  • 190.Cong W-T, Jin L-T, Hwang S-Y, Choi J-K. Fast fluorescent staining of protein in sodium dodecyl sulfate polyacrylamide gels by palmatine. Electrophor. 2008;29:417–423. doi: 10.1002/elps.200700382. [DOI] [PubMed] [Google Scholar]
  • 191.Leibowitz MJ, Wang RW. Visualization and elution of unstained proteins from polyacrylamide gels. Anal Biochem. 1984;137:161–163. doi: 10.1016/0003-2697(84)90364-6. [DOI] [PubMed] [Google Scholar]
  • 192.Yamamoto H, Nakatani M, Shinya K, Kim B-H, Kakuno T. Ultraviolet imaging densitometry of unstained gels for two-dimensional electrophoresis. Anal Biochem. 1990;191:58–64. doi: 10.1016/0003-2697(90)90387-o. [DOI] [PubMed] [Google Scholar]
  • 193.Koutny LB, Yeung ES. On-line detection of proteins in gel electrophoresis by ultraviolet absorption and by native fluorescence utilizing a charge-coupled device imaging system. Anal Chem. 1993;65:183–187. doi: 10.1021/ac00007a019. [DOI] [PubMed] [Google Scholar]
  • 194.Roegener J, Lutter P, Reinhardt R, Bluggel M, Meyer HE, Anselmetti D. Ultrasensitive detection of unstained proteins in acrylamide gels by native UV fluorescence. Anal Chem. 2002;75:157–159. doi: 10.1021/ac020517o. [DOI] [PubMed] [Google Scholar]
  • 195.Ladner CL, Yang J, Turner RJ, Edwards RA. Visible fluorescent detection of proteins in polyacrylamide gels without staining. Anal Biochem. 2004;326:13–20. doi: 10.1016/j.ab.2003.10.047. [DOI] [PubMed] [Google Scholar]
  • 196.Sluszny C, Yeung ES. One- and two-dimensional miniaturized electrophoresis of proteins with native fluorescence detection. Anal Chem. 2004;76:1359–1365. doi: 10.1021/ac035336g. [DOI] [PubMed] [Google Scholar]
  • 197.Zhang H, Yeung ES. Ultrasensitive native fluorescence detection of proteins with miniaturized polyacrylamide gel electrophoresis by laser side-entry excitation. Electrophor. 2006;27:3609–3618. doi: 10.1002/elps.200600020. [DOI] [PubMed] [Google Scholar]
  • 198.Dahlberg AE, Dingman CW, Peacock AC. Electrophoretic characterization of bacterial polyribosomes in agarose-acrylamide composite gels. J Mol Biol. 1969;41:139–146. doi: 10.1016/0022-2836(69)90131-4. [DOI] [PubMed] [Google Scholar]
  • 199.Green MR, Pastewka JV, Peacock AC. Differential staining of phosphoproteins on polyacrylamide gels with a cationic carbocyanine dye. Anal Biochem. 1973;56:43–51. doi: 10.1016/0003-2697(73)90167-x. [DOI] [PubMed] [Google Scholar]
  • 200.Green MR, Pastewka JV. Identification of sialic acid-rich glycoproteins on polyacrylamide gels. Anal Biochem. 1975;65:66–72. doi: 10.1016/0003-2697(75)90491-1. [DOI] [PubMed] [Google Scholar]
  • 201.King LE, Morrison M. The visualization of human erythrocyte membrane proteins and glycoproteins in SDS polyacrylamide gels employing a single staining procedure. Anal Biochem. 1976;71:223–230. doi: 10.1016/0003-2697(76)90031-2. [DOI] [PubMed] [Google Scholar]
  • 202.Campbell K, MacLennan D, Jorgensen A. Staining of the Ca2+-binding proteins, calsequestrin, calmodulin, troponin C, and S-100, with the cationic carbocyanine dye "Stains-all". J Biol Chem. 1983;258:11267–11273. [PubMed] [Google Scholar]
  • 203.Keyser JW. Staining of serum glycoproteins after electrophoretic separation in acrylamide gels. Anal Biochem. 1964;9:249–252. doi: 10.1016/0003-2697(64)90111-3. [DOI] [PubMed] [Google Scholar]
  • 204.Zacharius RM, Zell TE, Morrison JH, Woodlock JJ. Glycoprotein staining following electrophoresis on acrylamide gels. Anal Biochem. 1969;30:148–152. doi: 10.1016/0003-2697(69)90383-2. [DOI] [PubMed] [Google Scholar]
  • 205.Matthieu J-M, Quarles RH. Quantitative scanning of glycoproteins on polyacrylamide gels stained with periodic acid-Schiff reagent (PAS) Anal Biochem. 1973;55:313–316. doi: 10.1016/0003-2697(73)90321-7. [DOI] [PubMed] [Google Scholar]
  • 206.Doerner KC, White BA. Detection of glycoproteins separated by nondenaturing polyacrylamide gel electrophoresis using the periodic acid-Schiff stain. Anal Biochem. 1990;187:147–150. doi: 10.1016/0003-2697(90)90433-a. [DOI] [PubMed] [Google Scholar]
  • 207.Wardi AH, Michos GA. Alcian blue staining of glycoproteins in acrylamide disc electrophoresis. Anal Biochem. 1972;49:607–609. doi: 10.1016/0003-2697(72)90472-1. [DOI] [PubMed] [Google Scholar]
  • 208.Eckhardt AE, Hayes CE, Goldstein IJ. A sensitive fluorescent method for the detection of glycoproteins in polyacrylamide gels. Anal Biochem. 1976;73:192–197. doi: 10.1016/0003-2697(76)90154-8. [DOI] [PubMed] [Google Scholar]
  • 209.Furlan M, Perret BA, Beck EA. Staining of glycoproteins in polyacrylamide and agarose gels with fluorescent lectins. Anal Biochem. 1979;96:208–214. doi: 10.1016/0003-2697(79)90574-8. [DOI] [PubMed] [Google Scholar]
  • 210.Muñoz G, Marshall S, Cabrera M, Horvat A. Enhanced detection of glycoproteins in polyacrylamide gels. Anal Biochem. 1988;170:491–494. doi: 10.1016/0003-2697(88)90663-x. [DOI] [PubMed] [Google Scholar]
  • 211.Racusen D. Glycoprotein detection in polyacrylamide gel with thymol and sulfuric acid. Anal Biochem. 1979;99:474–476. doi: 10.1016/s0003-2697(79)80035-4. [DOI] [PubMed] [Google Scholar]
  • 212.Steinberg TH, Top KPO, Berggren KN, Kemper C, Jones L, Diwu Z, Haugland RP, Patton WF. Rapid and simple single nanogram detection of glycoproteins in polyacrylamide gels and on electroblots. Proteomics. 2001;1:841–855. doi: 10.1002/1615-9861(200107)1:7<841::AID-PROT841>3.0.CO;2-E. [DOI] [PubMed] [Google Scholar]
  • 213.Hart C, Schulenberg B, Steinberg TH, Leung W-Y, Patton WF. Detection of glycoproteins in polyacylamide gels and on electroblots using Pro-Q Emerald 488 dye, a fluorescent periodate schiff-base stain. Electrophor. 2003;24:588–598. doi: 10.1002/elps.200390069. [DOI] [PubMed] [Google Scholar]
  • 214.Misenheimer TM, Hahr AJ, Harms AC, Annis DS, Mosher DF. Disulfide connectivity of recombinant C-terminal region of human thrombospondin 2. J Biol Chem. 2001;276:45882–45887. doi: 10.1074/jbc.M104218200. [DOI] [PubMed] [Google Scholar]
  • 215.Pío R, Martínez A, Unsworth EJ, Kowalak JA, Bengoechea JA, Zipfel PF, Elsasser TH, Cuttitta F. Complement factor H is a serum-binding protein for adrenomedullin, and the resulting complex modulates the bioactivities of both partners. J Biol Chem. 2001;276:12292–12300. doi: 10.1074/jbc.M007822200. [DOI] [PubMed] [Google Scholar]
  • 216.Khidekel N, Arndt S, Lamarre-Vincent N, Lippert A, Poulin-Kerstien KG, Ramakrishnan B, Qasba PK, Hsieh-Wilson LC. A chemoenzymatic approach toward the rapid and sensitive detection of O-GlcNAc posttranslational modifications. J Am Chem Soc. 2003;125:16162–16163. doi: 10.1021/ja038545r. [DOI] [PubMed] [Google Scholar]
  • 217.Clark PM, Dweck JF, Mason DE, Hart CR, Buck SB, Peters EC, Agnew BJ, Hsieh-Wilson LC. Direct in-gel fluorescence detection and cellular imaging of O-GlcNAc-modified proteins. J Am Chem Soc. 2008;130:11576–11577. doi: 10.1021/ja8030467. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 218.Patton WF. Detection technologies in proteome analysis. J Chromatogr B Anal Technol Biomed Life Sci. 2002;771:3–31. doi: 10.1016/s1570-0232(02)00043-0. [DOI] [PubMed] [Google Scholar]
  • 219.Cutting JA, Roth TF. Staining of Phospho-proteins on acrylamide gel electropherograms. Anal Biochem. 1973;54:386–394. doi: 10.1016/0003-2697(73)90367-9. [DOI] [PubMed] [Google Scholar]
  • 220.Hegenauer J, Ripley L, Nace G. Staining acidic phosphoproteins (phosvitin) in electrophoretic gels. Anal Biochem. 1977;78:308–311. doi: 10.1016/0003-2697(77)90038-0. [DOI] [PubMed] [Google Scholar]
  • 221.Steinberg TH, Agnew BJ, Gee KR, Leung W-Y, Goodman T, Schulenberg B, Hendrickson J, Beechem JM, Haugland RP, Patton WF. Global quantitative phosphoprotein analysis using multiplexed proteomics technology. Proteomics. 2003;3:1128–1144. doi: 10.1002/pmic.200300434. [DOI] [PubMed] [Google Scholar]
  • 222.Schulenberg B, Goodman TN, Aggeler R, Capaldi RA, Patton WF. Characterization of dynamic and steady-state protein phosphorylation using a fluorescent phosphoprotein gel stain and mass spectrometry. Electrophor. 2004;25:2526–2532. doi: 10.1002/elps.200406007. [DOI] [PubMed] [Google Scholar]
  • 223.Agrawal GK, Thelen JJ. Development of a simplified, economical polyacrylamide gel staining protocol for phosphoproteins. Proteomics. 2005;5:4684–4688. doi: 10.1002/pmic.200500021. [DOI] [PubMed] [Google Scholar]
  • 224.Kinoshita E, Kinoshita-Kikuta E, Takiyama K, Koike T. Phosphate-binding tag, a new tool to visualize phosphorylated proteins. Mol Cell Proteomics. 2006;5:749–757. doi: 10.1074/mcp.T500024-MCP200. [DOI] [PubMed] [Google Scholar]
  • 225.Yamada S, Nakamura H, Kinoshita E, Kinoshita-Kikuta E, Koike T, Shiro Y. Separation of a phosphorylated histidine protein using phosphate affinity polyacrylamide gel electrophoresis. Anal Biochem. 2007;360:160–162. doi: 10.1016/j.ab.2006.10.005. [DOI] [PubMed] [Google Scholar]
  • 226.Barbieri CM, Stock AM. Universally applicable methods for monitoring response regulator aspartate phosphorylation both in vitro and in vivo using Phos-tag-based reagents. Anal Biochem. 2008;376:73–82. doi: 10.1016/j.ab.2008.02.004. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 227.Chung MC-M. A specific iron stain for iron-binding proteins in polyacrylamide gels: application to transferrin and lactoferrin. Anal Biochem. 1985;148:498–502. doi: 10.1016/0003-2697(85)90258-1. [DOI] [PubMed] [Google Scholar]
  • 228.Leong LM, Tan BH, Ho KK. A specific stain for the detection of nonheme iron proteins in polyacrylamide gels. Anal Biochem. 1992;207:317–320. doi: 10.1016/0003-2697(92)90018-3. [DOI] [PubMed] [Google Scholar]
  • 229.Kuo C-F, Fridovich I. A stain for iron-containing proteins sensitive to nanogram levels of iron. Anal Biochem. 1988;170:183–185. doi: 10.1016/0003-2697(88)90106-6. [DOI] [PubMed] [Google Scholar]
  • 230.Frings CS, Foster LB, Cohen PS. Electrophoretic separation of serum lipoproteins in polyacrylamide gel. Clin Chem. 1971;17:111–114. [PubMed] [Google Scholar]
  • 231.Muniz N. Measurement of plasma lipoproteins by electrophoresis on polyacrylamide gel. Clin Chem. 1977;23:1826–1833. [PubMed] [Google Scholar]
  • 232.McNamara JR, Campos H, Adolphson JL, Ordovas JM, Wilson PW, Albers JJ, Usher DC, Schaefer EJ. Screening for lipoprotein[a] elevations in plasma and assessment of size heterogeneity using gradient gel electrophoresis. J Lipid Res. 1989;30:747–755. [PubMed] [Google Scholar]
  • 233.Rainwater DL, Moore PH, Jr, Shelledy WR, Dyer TD, Slifer SH. Characterization of a composite gradient gel for the electrophoretic separation of lipoproteins. J Lipid Res. 1997;38:1261–1266. [PubMed] [Google Scholar]
  • 234.Jencks WP, Durrum EL, Jetton MR. Paper electrophoresis as a quantitative method: the staining of serum lipoproteins. J Clin Investig. 1955;34:1437–1448. doi: 10.1172/JCI103193. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 235.Noble RP. Electrophoretic separation of plasma lipoproteins in agarose gel. J Lipid Res. 1968;9:693–700. [PubMed] [Google Scholar]
  • 236.Smejkal GB, Hoff HF. Filipin staining of lipoproteins in polyacrylamide gels: sensitivity and photobleaching of the fluorophore and its use in a double staining method. Electrophor. 1994;15:922–925. doi: 10.1002/elps.11501501133. [DOI] [PubMed] [Google Scholar]
  • 237.Williams NK, Prosselkov P, Liepinsh E, Line I, Sharipo A, Littler DR, Curmi PMG, Otting G, Dixon NE (2002) In vivo protein cyclization promoted by a circularly permuted Synechocystis sp. PCC6803 Dna D mini-intein. J Biol Chem 277:7790–7798 [DOI] [PubMed]
  • 238.Griffin BA, Adams SR, Tsien RY. Specific covalent labeling of recombinant protein molecules inside live cells. Science. 1998;281:269–272. doi: 10.1126/science.281.5374.269. [DOI] [PubMed] [Google Scholar]
  • 239.Adams SR, Campbell RE, Gross LA, Martin BR, Walkup GK, Yao Y, Llopis J, Tsien RY. New biarsenical ligands and tetracysteine motifs for protein labeling in vitro and in vivo: synthesis and biological applications. J Am Chem Soc. 2002;124:6063–6076. doi: 10.1021/ja017687n. [DOI] [PubMed] [Google Scholar]
  • 240.Keller PJ, Cohen E, Hollinshead Beeley JA. Canine pancreatic ribosomes: II. The protein moiety. J Biol Chem. 1968;243:1271–1276. [PubMed] [Google Scholar]
  • 241.Lawrence JM, Hedwig EH, Grant DR (1970) Analysis of wheat flour proteins by polyacrylamide gel electrophoresis. Cereal Chem 47:98–109
  • 242.Bertolini MJ, Tankersley DL, Schroeder DD. Staining and destaining polyacrylamide gels: a comparison of coomassie blue and fast green protein dyes. Anal Biochem. 1976;71:6–13. doi: 10.1016/0003-2697(76)90003-8. [DOI] [PubMed] [Google Scholar]
  • 243.Castellanos-Serra L, Hardy E. Detection of biomolecules in electrophoresis gels with salts of imidazole and zinc II: a decade of research. Electrophor. 2001;22:864–873. doi: 10.1002/1522-2683()22:5<864::AID-ELPS864>3.0.CO;2-Y. [DOI] [PubMed] [Google Scholar]
  • 244.Talbot DN, Yphantis DA. Fluorescent monitoring of SDS gel electrophoresis. Anal Biochem. 1971;44:246–253. doi: 10.1016/0003-2697(71)90367-8. [DOI] [PubMed] [Google Scholar]
  • 245.Eng PR, Parkes CO. SDS electrophoresis of fluorescamine-labeled proteins. Anal Biochem. 1974;59:323–325. doi: 10.1016/0003-2697(74)90041-4. [DOI] [PubMed] [Google Scholar]
  • 246.Aragay AM, Diaz P, Daban J-R. A fluorescent method for the rapid staining and quantitation of proteins in sodium dodecyl sulfate-polyacrylamide gels. Electrophor. 1985;6:527–531. [Google Scholar]
  • 247.Alba FJ, Bermudez A, Bartolomé S, Daban J-R. Detection of five nano-grams of protein by two-minute nile red staining of unfixed SDS gels. Biotech. 1996;21:625–626. doi: 10.2144/96214bm12. [DOI] [PubMed] [Google Scholar]
  • 248.Ünlü M, Morgan ME, Minden JS. Difference gel electrophoresis. A single gel method for detecting changes in protein extracts. Electrophor. 1997;18:2071–2077. doi: 10.1002/elps.1150181133. [DOI] [PubMed] [Google Scholar]
  • 249.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;276:129–143. doi: 10.1006/abio.1999.4364. [DOI] [PubMed] [Google Scholar]
  • 250.Luo L-D, Wirth PJ. Consecutive silver staining and autoradiography of 35S and 32P-labeled cellular proteins: application for the analysis of signal transducing pathways. Electrophor. 1993;14:127–136. doi: 10.1002/elps.1150140121. [DOI] [PubMed] [Google Scholar]
  • 251.Kemper C, Steinberg TH, Jones L, Patton WF. Simultaneous, two-color fluorescence detection of total protein profiles and beta-glucuronidase activity in polyacrylamide gel. Electrophor. 2001;22:970–976. doi: 10.1002/1522-2683()22:5<970::AID-ELPS970>3.0.CO;2-5. [DOI] [PubMed] [Google Scholar]
  • 252.Berggren KN, Chernokalskaya E, Lopez MF, Beechem JM, Patton WF. Comparison of three different fluorescent visualization strategies for detecting Escherichia coli ATP synthase subunits after sodium dodecyl sulfate-polyacrylamide gel electrophoresis. Proteomics. 2001;1:54–65. doi: 10.1002/1615-9861(200101)1:1<54::AID-PROT54>3.0.CO;2-5. [DOI] [PubMed] [Google Scholar]
  • 253.Evans CA, Tonge R, Blinco D, Pierce A, Shaw J, Lu Y, Hamzah HG, Gray A, Downes CP, Gaskell SJ, Spooncer E, Whetton AD. Comparative proteomics of primitive hematopoietic cell populations reveals differences in expression of proteins regulating motility. Blood. 2004;103:3751–3759. doi: 10.1182/blood-2003-09-3294. [DOI] [PubMed] [Google Scholar]
  • 254.Wheelock ÅM, Morin D, Bartosiewicz M, Buckpitt AR. Use of a fluorescent internal protein standard to achieve quantitative two-dimensional gel electrophoresis. Proteomics. 2006;6:1385–1398. doi: 10.1002/pmic.200402083. [DOI] [PubMed] [Google Scholar]
  • 255.Forsen E, Abadal G, Ghatnekar-Nilsson S, Teva J, Verd J, Sandberg R, Svendsen W, Perez-Murano F, Esteve J, Figueras E, Campabadal F, Montelius L, Barniol N, Boisen A. Ultrasensitive mass sensor fully integrated with complementary metal-oxide-semiconductor circuitry. App Phys Lett. 2005;87:043507. [Google Scholar]
  • 256.Berth M, Moser FM, Kolbe M, Bernhardt J. The state of the art in the analysis of two-dimensional gel electrophoresis images. App Microbiol Biotechnol. 2007;76:1223–1243. doi: 10.1007/s00253-007-1128-0. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 257.Millioni R, Miuzzo M, Sbrignadello S, Murphy E, Puricelli L, Tura A, Bertacco E, Rattazzi M, Iori E, Tessari P. Delta2D and proteomweaver: performance evaluation of two different approaches for 2-DE analysis. Electrophor. 2010;8:1311–1317. doi: 10.1002/elps.200900766. [DOI] [PubMed] [Google Scholar]
  • 258.Yergey AL, Coorssen JR, Backlund PS, Blank PS, Humphrey GA, Zimmerberg J, Campbell JM, Vestal ML. De novo sequencing of peptides using MALDI/TOF-TOF. J Am Soc Mass Spectrom. 2002;13:784–791. doi: 10.1016/S1044-0305(02)00393-8. [DOI] [PubMed] [Google Scholar]
  • 259.Aebersold R, Mann M. Mass spectrometry-based proteomics. Nature. 2003;422:198–207. doi: 10.1038/nature01511. [DOI] [PubMed] [Google Scholar]
  • 260.Wysocki VH, Resing KA, Zhang Q, Cheng G. Mass spectrometry of peptides and proteins. Methods. 2005;35:211–222. doi: 10.1016/j.ymeth.2004.08.013. [DOI] [PubMed] [Google Scholar]
  • 261.Domon B, Aebersold R. Mass spectrometry and protein analysis. Science. 2006;312:212–217. doi: 10.1126/science.1124619. [DOI] [PubMed] [Google Scholar]
  • 262.Han X, Aslanian A, Yates Iii JR. Mass spectrometry for proteomics. Curr Opin Chem Biol. 2008;12:483–490. doi: 10.1016/j.cbpa.2008.07.024. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 263.Aebersold R, Goodlett DR. Mass spectrometry in proteomics. Chem Rev. 2001;101:269–296. doi: 10.1021/cr990076h. [DOI] [PubMed] [Google Scholar]
  • 264.Bantscheff M, Boesche M, Eberhard D, Matthieson T, Sweetman G, Kuster B. Robust and sensitive itraq quantification on an ltq orbitrap mass spectrometer. Mol Cell Proteomics. 2008;7:1702–1713. doi: 10.1074/mcp.M800029-MCP200. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 265.Godoy L, Olsen J, Souza G, Li G, Mortensen P, Mann M. Status of complete proteome analysis by mass spectrometry: SILAC labeled yeast as a model system. Genome Biol. 2006;7:R50. doi: 10.1186/gb-2006-7-6-r50. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 266.Kierszenbaum F, Levison SA, Dandliker WB. Fractionation of fluorescent-labeled proteins according to the degree of labeling. Anal Biochem. 1969;28:563–572. doi: 10.1016/0003-2697(69)90210-3. [DOI] [PubMed] [Google Scholar]
  • 267.Karp NA, Feret R, Rubtsov DV, Lilley KS. Comparison of DIGE and post-stained gel electrophoresis with both traditional and SameSpots analysis for quantitative proteomics. Proteomics. 2008;8:948–960. doi: 10.1002/pmic.200700812. [DOI] [PubMed] [Google Scholar]
  • 268.Martinez JLC. International Congress on Analytical Proteomics. 2011. http://sing.ei.uvigo.es/ICAP/

Articles from Journal of Chemical Biology are provided here courtesy of Springer-Verlag

RESOURCES