首页 | 本学科首页   官方微博 | 高级检索  
相似文献
 共查询到20条相似文献,搜索用时 31 毫秒
1.
Relative quantitation of aspartyl and isoaspartyl residue mixtures from asparagine deamidation is demonstrated using electron capture dissociation without prior HPLC separation. The method utilizes the linear relationship found between the relative abundance of the isoaspartyl diagnostic ion, z(n)-57, and % isoaspartyl content based on the ECD spectra of known isoaspartyl/aspartyl mixtures of synthetic peptides. The observed linearity appears to be sequence independent because the relationship exists despite sequence variations and changes in backbone fragment abundances when isoaspartyl and aspartyl residues are interchanged. Furthermore, a new method to calculate the relative abundances of isomer from protein deamidation without synthetic peptides is proposed and tested using a linear peptide released by protein digestion that contains the deamidation site. The proteolytic peptide can be rapidly aged to the expected 3:1 (isoaspartyl:aspartyl) mixture to generate a two-point calibration standard for ECD analysis. The procedure can then be used to determine the relative abundance of deamidation products from in vivo or in vitro protein aging experiments.  相似文献   

2.
Proteolytic (18)O-labeling of peptides has been studied and optimized in order to improve the labeling efficiency and to accelerate the process without increasing the degree of incomplete labeling. Using peptides generated from tryptic digested bovine serum albumin (BSA) and cytochrome c as model proteins, it was shown that complete labeling was achieved after 2 h at pH 6. To increase the sample throughput in a bottom-up proteomic setup, tryptic digestion of proteins in-solution was replaced with tryptic digestion using immobilized trypsin. As a result, an integrated approach was made possible, where both digestion (pH 8) and (18)O/(16)O-labeling of the resulting peptides (pH 6) were done using immobilized trypsin beads. This simplified the sample handling and reduced the overall reaction time significantly: the setup enabled tryptic digestion and (18)O/(16)O-labeling without sample transfer steps within 3.5 h with average (18)O/(16)O-ratios of 0.96±0.13 in aqueous buffer. The initial results were confirmed with a more complex matrix, by spiking urine with the model proteins, yielding results comparable with the ratios obtained in buffer. Satisfying ratios were also achieved regarding urinary proteins identified in a full scale bottom-up experiment. Average (18)O/(16)O-peptide ratios of 0.83±0.13 and 0.91±0.27 indicated good performance in a highly relevant matrix for biomarker discovery.  相似文献   

3.
蛋白质主要以复合物的形式参与各项生命活动.化学交联质谱(CXMS)技术作为近年来新兴的蛋白质复合物解析技术,不仅可实现蛋白质复合物规模化解析,而且普遍适用于任意相对分子质量和纯度的蛋白质复合物样品,因此已成为X-射线晶体衍射技术、冷冻电镜技术等蛋白质复合物解析经典技术的重要补充.目前,CXMS主要采用胰蛋白酶将交联后的...  相似文献   

4.
A potential pitfall in 18O-based N-linked glycosylation site mapping   总被引:1,自引:0,他引:1  
A common procedure for identifying N-linked glycosylation sites involves tryptic digestion of the glycoprotein, followed by the conversion of glycosylated asparagine residues into (18)O-labeled aspartic acids by PNGase F digestion in (18)O water. The 3 Da mass tag created by this process is readily observable by liquid chromatography/tandem mass spectrometry (LC/MS/MS) analysis, and is often used to identify the sites of N-linked glycosylation. While using this procedure, we noticed that 60% of the asparagines identified as being glycosylated were not part of the consensus sequence required for N-linked glycosylation, and thus were not biologically possible. Investigation into the source of this unacceptably high false positive rate demonstrated that even after reversed-phase cleanup and heat denaturation, the trypsin used for proteolysis was still active and led to the incorporation of (18)O into the C-termini of the peptides during the deglycosylation step. The resulting mass shift accounted for most of the false positive sites, as the database search algorithm confused it with an (18)O-labeled Asp residue near the C-terminus of a peptide. This problem can be overcome by eliminating trypsin from the solution prior to performing the deglycosylation process, by resuspending the peptides in natural abundance water following deglycosylation, or by allowing (18)O incorporation into the C-terminus as a variable modification during the database search. These methods have been demonstrated on a model protein, and are applicable to the analyses of glycoproteins that are digested with trypsin or another serine protease prior to enzymatic release of the carbohydrate side chains. This study should alert investigators in the field to this potential and unexpected pitfall and provide strategies to overcome this phenomenon.  相似文献   

5.
The possible use of divalent alkaline-earth metal ions, including Mg2+, Ca2+, Sr2+, and Ba2+, as charge carrier for electron capture dissociation of peptides was investigated. Model peptides of RGGGVGGGR and NGGGWGGGN were used to simplify the interpretation of spectral information. It was demonstrated that useful electron capture dissociation (ECD) tandem mass spectra of these metalated peptides could be generated. Interestingly, peptides metalated with different alkaline-earth metal ions generated very similar ECD tandem mass spectra. Metalated c-ions and z-ions were the predominant fragment ions. Only Mg2+-metalated peptides gave somewhat different results. Some nonmetalated c-ions were observed from ECD of [RGGGVGGGR + Mg]2+ but not from [NGGGWGGGN + Mg]2+. Together with some ab initio calculations, it was established that the bound metal ions might activate the acidity of the amide hydrogen. With the presence of high proton affinity moiety, such as N-terminal amino group and/or side chain of the arginine residues, the metalated peptide ions could exist predominantly in their zwitterion forms, in which one or two backbone amide group(s) was deprotonated and the high proton affinity functional group(s) was protonated. It was believed that electron capture leads primarily to the reduction of the mobile proton rather than the metal ions. With this zwitterion model, the formation of nonmetalated c-fragments and the generation of similar ECD spectra for peptides metalated with various alkaline-earth metal ions could readily to be explained. Another interesting observation in the ECD mass spectra of metalated peptides is related to the enhanced formation of the minor ECD products, i.e., (c - 1)(+*) and (z + 1)+ ions. Together with ab initio calculations using a truncated peptide model, various possible reaction mechanisms for the formation of these minor ECD products were evaluated. It was concluded that hydrogen transfer between the initiated formed c and z(.) species plays an important role in the formation (c - 1)(+*) and (z + 1)+ ions. Although peptides metalated with these metal ions do not have better ECD efficiency compared to the multiply-protonated peptides, it provides practical accessibility of ECD methods to analyze small peptides with no basic amino acid residues.  相似文献   

6.
C-terminal rearrangement ions [b(n-1) + H2O] (where n refers to the total number of residues of peptides) are frequently observed for peptides which contain basic amino acid(s), especially arginine, at or near their N termini in low- and high-energy collision-induced dissociation or post-source decay (PSD) spectra. Here we report a novel rearrangement, associated with PSD for serine- or threonine-containing peptides that are susceptible to C-terminal rearrangement. Based on PSD analyses of serine- or threonine-containing bradykinin and its analogs, which have been ethyl-esterified or 18O labeled at their C termini, the [b(k) + H2O] (where k denotes the position adjacent to the left of the Ser/Thr residue) ion is generally thought to be formed by the transfer of the hydroxyl moiety of a serine or threonine residue to the carbonyl group of the residue to its left accompanied by the loss of the remaining C-terminal portion of the peptide. When the Ser/Thr is at or near the C terminus, the present [b(k) + H2O] ion could be formed via two pathways, i.e., the Ser/Thr-related rearrangement and the conventional C-terminal rearrangement, which has been clearly verified by 18O labeling at the C terminus. In addition, the ions which are formally designated as [y(m)b(l) + H2O], where y(m)b(l) denotes a b-type internal ion, are also briefly described.  相似文献   

7.
Hydrogen rearrangement is an important process in radical chemistry. A high degree of H. rearrangement to and from z. ionic fragments (combined occurrence frequency 47% compared with that of z.) is confirmed in analysis of 15,000 tandem mass spectra of tryptic peptides obtained with electron capture dissociation (ECD), including previously unreported double H. losses. Consistent with the radical character of H. abstraction, the residue determining the formation rate of z' = z. + H. species is found to be the N-terminal residue in z. species. The size of the complementary c(m)' fragment turned out to be another important factor, with z' species dominating over z. ions for m < or = 6. The H. atom was found to be abstracted from the side chains as well as from alpha-carbon groups of residues composing the c' species, with Gln and His in the c' fragment promoting H. donation and Asp and Ala opposing it. Ab initio calculations of formation energies of .A radicals (A is an amino acid) confirmed that the main driving force for H. abstraction by z. is the process exothermicity. No valid correlation was found between the NC(alpha) bond strength and the frequency of this bond cleavage, indicating that other factors than thermochemistry are responsible for directing the site of ECD cleavage. Understanding hydrogen attachment to and loss from ECD fragments should facilitate automatic interpretation ECD mass spectra in protein identification and characterization, including de novo sequencing.  相似文献   

8.
Designing small peptides that are capable of binding Cu2+ ions mainly through the side‐chain functionalities is a hard task because the amide nitrogen atoms strongly compete for Cu2+ ion coordination. However, the design of such peptides is important for obtaining biomimetic small systems of metalloenyzmes as well as for the development of artificial systems. With this in mind, a cyclic decapeptide, C‐Asp, which contained three His residues and one Asp residue, and its linear derivative, O‐Asp, were synthesized. The C‐Asp peptide has two Pro? Gly β‐turn‐inducer units and, as a result of cyclization, and as shown by CD spectroscopy, its backbone is constrained into a more defined conformation than O‐Asp, which is linear and contains a single Pro? Gly unit. A detailed potentiometric, mass spectrometric, and spectroscopic study (UV/Vis, CD, and EPR spectroscopy) showed that at a 1:1 Cu2+/peptide ratio, both peptides formed a major [CuHL]2+ species in the pH range 5.0–7.5 (C‐Asp) and 5.5–7.0 (O‐Asp). The corrected stability constants of the protonated species (log K*CuH(O?Asp)=9.28 and log K*CuH(C?Asp)=10.79) indicate that the cyclic peptide binds Cu2+ ions with higher affinity. In addition, the calculated value of Keff shows that this higher affinity for Cu2+ ions prevails at all pH values, not only for a 1:1 ratio but even for a 2:1 ratio. The spectroscopic data of both [CuHL]2+ species are consistent with the exclusive coordination of Cu2+ ions by the side‐chain functionalities of the three His residues and the Asp residue in a square‐planar or square‐pyramidal geometry. Nonetheless, although these data show that, upon metal coordination, both peptides adopt a similar fold, the larger conformational constraints that are present in the cyclic scaffold results in different behaviour for both [CuHL]2+ species. CD and NMR analysis revealed the formation of a more rigid structure and a slower Cu2+‐exchange rate for [CuH(C‐Asp)]2+ compared to [CuH(O‐Asp]2+. This detailed comparative study shows that cyclization has a remarkable effect on the Cu2+‐coordination properties of the C‐Asp peptide, which binds Cu2+ ions with higher affinity at all pH values, stabilizes the [CuHL]2+ species in a wider pH range, and has a slower Cu2+‐exchange rate compared to O‐Asp.  相似文献   

9.
Proteolytic (18)O labeling is a very powerful tool for differential analysis applied to proteome studies. However, it is a relatively new technique and the optimization of the labeling process still needs some attention. We found that the two-step post-proteolytic labeling should be favored over the conventional digestion of proteins in H(2) (18)O, since the former allows for higher sample concentrations and thus more favorable kinetics. It was demonstrated that the inhibitory effect of urea on (18)O incorporation could be compensated by the use of higher sample concentrations. Furthermore, it was shown that heat-deactivation of trypsin prevents (18)O/(16)O back-exchange. In addition, no non-specific hydrolysis of the peptides could be observed as a result of the heating. Heat inactivation of trypsin opens the way for the use of capillary electrophoresis as a separation technique in proteolytic labeling studies, as it abolishes the need for use of detrimental additives. Analysis of a labeled protein digest by capillary isoelectric focusing/mass spectrometry showed the applicability of the method. No back-exchange was observed across the entire electropherogram.  相似文献   

10.
The dissociation of intermolecularly crosslinked peptides was evaluated for a series of peptides with proline or aspartic acid residues positioned adjacent to the crosslinking sites (lysine residues). The peptides were crosslinked with either disuccinimidyl suberate (DSS) or disuccinimidyl L-tartrate (DST), and the influence of proline and aspartic acid residues on the fragmentation patterns were investigated for precursor ions with and without a mobile proton. Collisionally activated dissociation (CAD) spectra of aspartic acid-containing crosslinked peptide ions, doubly-charged with both protons sequestered, were dominated by cleavage C-terminal to the Asp residue, similar to that of unmodified peptides. The proline-containing crosslinked peptides exhibited a high degree of internal ion formation, with the resulting product ions having an N-terminal proline residue. Upon dissociation of the doubly-charged crosslinked peptides, twenty to fifty percent of the fragment ion abundance was accounted for by multiple cleavage products. Crosslinked peptides possessing a mobile proton yielded almost a full series of b- and y-type fragment ions, with only proline-directed fragments still observed at high abundances. Interestingly, the crosslinked peptides exhibited a tendency to dissociate at the amide bond C-terminal to the crosslinked lysine residue, relative to the N-terminal side. One could envision updating computer algorithms to include these crosslinker specific product ions--particularly for precursor ions with localized protons--that provide complementary and confirmatory information, to offer more confident identification of both the crosslinked peptides and the location of the crosslink, as well as affording predictive guidelines for interpretation of the product-ion spectra of crosslinked peptides.  相似文献   

11.
In the present work we report on a novel and fast protocol for accurate bottom-up protein quantification that overcomes the drawbacks of in-gel digestion and MALDI analysis, while maintaining their benefits. It relies on the following steps: (i) gel electrophoresis separation of proteins, (ii) fast in-gel protein digestion with trypsin, (iii) (18)O-labeling through the decoupled method, (iv) quantification through selected peptides previously chosen using the (18)O inverse labeling approach and that, finally, (v) it takes advantage of software specifically developed to select the peptides that will drive the quantification of the protein in an automated mode. We have accurately quantified the following six proteins: glycogen phosphorylase, BSA, ovalbumin, carbonic anhydrase, trypsin inhibitor, and α-lactalbumin. As a case study we have quantified the protein vitellogenin in plasma of Cyprinus carpio exposed to high levels of estrogens. The proposed new protocol was validated against the traditional ELISA method; both were found to provide comparable results (non-parametric Mann-Whitney U-test).  相似文献   

12.
Oxidation of cysteine residues to cysteic acids in C-terminal arginine-eontaining peptides (such as those derived by tryptic digestion of proteins) strongly promotes the formation of multiple members of the Y? series of fragment ions following low energy collision-activated decomposition (CAD) of the protonated peptides, Removal of the arginine residue abolishes the effect, which is also attenuated by conversion of the arginine to dimethylpyrim-idylornithine. The data indicate the importance of an intraionic interaction between the cysteic acid and arginine side-chains. Low energy CAD of peptides which include cysteic acid and histidine residues, also provides evidence for intraionic interactions. It is proposed that these findings are consistent with the general hypothesis that an increased heterogeneity (with respect to location of charge) of the protonated peptide precursor ion population is beneficial to the generation of a high yield of product ions via several charge-directed, low energy fragmentation pathways. Furthermore, these data emphasize the significance of gas-phase conformations of protonated peptides in determining fragmentation pathways.  相似文献   

13.
MS/MS data derived from the [M-H](-) ions of desulfated caerulein peptides provide (i) sequencing information from a combination of alpha, beta and gamma backbone cleavages, and (ii) identification of specific amino acid side chains by side-chain cleavages [e.g. Ser (-CH(2)O), Thr (-CH(3)CHO) and Asp (-H(2)O)] (fragmentations having no counterparts in positive ion spectra). In addition, delta and/or gamma backbone cleavage ions from Asp residues identify the position of these residues in the peptide. In contrast, neither delta nor gamma cleavage ions are observed from either the Gln2 residue nor from Phe residues. Full structural information can be obtained from a consideration of the positive and negative ion MS/MS data in concert.  相似文献   

14.
Electron capture dissociation (ECD) of model peptides adducted with first row divalent transition metal ions, including Mn2+, Fe2+, Co2+, Ni2+, Cu2+, and Zn2+, were investigated. Model peptides with general sequence of ZGGGXGGGZ were used as probes to unveil the ECD mechanism of metalated peptides, where X is either V or W; and Z is either R or N. Peptides metalated with different divalent transition metal ions were found to generate different ECD tandem mass spectra. ECD spectra of peptides metalated by Mn2+ and Zn2+ were similar to those generated by ECD of peptides adducted with alkaline earth metal ions. Series of c-/z-type fragment ions with and without metal ions were observed. ECD of Fe2+, Co2+, and Ni2+ adducted peptides yielded abundant metalated a-/y-type fragment ions; whereas ECD of Cu2+ adducted peptides generated predominantly metalated b-/y-type fragment ions. From the present experimental results, it was postulated that electronic configuration of metal ions is an important factor in determining the ECD behavior of the metalated peptides. Due presumably to the stability of the electronic configuration, metal ions with fully-filled (i.e., Zn2+) and half filled (i.e., Mn2+) d-orbitals might not capture the incoming electron. Dissociation of the metal ions adducted peptides would proceed through the usual ECD channel(s) via “hot-hydrogen” or “superbase” intermediates, to form series of c-/z - fragments. For other transition metal ions studied, reduction of the metal ions might occur preferentially. The energy liberated by the metal ion reduction would provide enough internal energy to generate the “slow-heating” type of fragment ions, i.e., metalated a-/y- fragments and metalated b-/y- fragments.  相似文献   

15.
Deamidation of proteins occurs spontaneously under physiological conditions. Asparaginyl (Asn) residues may deamidate into aspartyl (Asp) residues, causing a change in both the charge and the conformation of peptides. It has been previously proposed by Capasso et al. that deamidation of relatively unrestrained Asn residues proceeds through a succinimide intermediate. This mechanism has been modeled by Konuklar et al. and the rate determining step for the deamidation process in neutral media has been shown to be the cyclization step leading to the succinimide intermediate. In the present study, possible water-assisted mechanisms, for both concerted and stepwise succinimide formation, were computationally explored using the B3LYP method with 6-31+G* basis set. Single point solvent calculations were carried out in water, by means of integral equation formalism-polarizable continuum model (IEF-PCM) at the B3LYP/6-31++G* level of theory. A novel route leading to the succinimide intermediate via tautomerization of the Asn side chain amide functionality has been proposed. The energetics of these pathways have been subject to a comparative study to identify the most probable mechanism for the deamidation of peptides in solution.  相似文献   

16.
The fragmentation behavior of nitrated and S-nitrosylated peptides were studied using collision induced dissociation (CID) and metastable atom-activated dissociation mass spectrometry (MAD-MS). Various charge states, such as 1+, 2+, 3+, 2–, of modified and unmodified peptides were exposed to a beam of high kinetic energy helium (He) metastable atoms resulting in extensive backbone fragmentation with significant retention of the post-translation modifications (PTMs). Whereas the high electron affinity of the nitrotyrosine moiety quenches radical chemistry and fragmentation in electron capture dissociation (ECD) and electron transfer dissociation (ETD), MAD does produce numerous backbone cleavages in the vicinity of the modification. Fragment ions of nitrosylated cysteine modifications typically exhibit more abundant neutral losses than nitrated tyrosine modifications because of the extremely labile nature of the nitrosylated cysteine residues. However, compared with CID, MAD produced between 66% and 86% more fragment ions, which preserved the labile –NO modification. MAD was also able to differentiate I/L residues in the modified peptides. MAD is able to induce radical ion chemistry even in the presence of strong radical traps and therefore offers unique advantages to ECD, ETD, and CID for determination of PTMs such as nitrated and S-nitrosylated peptides.  相似文献   

17.
Microwave assisted acid cleavage was applied directly to intact adenovirus type 5 to achieve denaturation and proteolysis in a single reaction. The speed of the digestion, coupled with the simplicity of MALDI analysis, allowed peptide products to be profiled in less than 5 min. Identification of peptides from a range of proteins by MALDI-TOF confirms that both denaturation and proteolysis were achieved using low concentrations of acetic acid (12.5%) and short incubations (1.5 to 2 min) at high temperatures (140° C). These conditions favor production of peptides that carry Asp on their C-termini. When this cleavage reaction is carried out in highly enriched H(2) (18)O, a single atom of (18)O is introduced site-specifically into the carboxyl terminal. The labeling reaction is applied to label reporter peptides from human adenovirus type 5 harvested from HeLa cells. Small peptide products of endogenous processing were also observed.  相似文献   

18.
Hydrogen (1H/2H) exchange combined with mass spectrometry (HX-MS) has become a recognized method for the analysis of protein structural dynamics. Presently, the incorporated deuterons are typically localized by enzymatic cleavage of the labeled proteins and single residue resolution is normally only obtained for a few residues. Determination of site-specific deuterium levels by gas-phase fragmentation in tandem mass spectrometers would greatly increase the applicability of the HX-MS method. The biggest obstacle in achieving this goal is the intramolecular hydrogen migration (i.e., hydrogen scrambling) that occurs during vibrational excitation of gas-phase ions. Unlike traditional collisional ion activation, electron capture dissociation (ECD) is not associated with substantial vibrational excitation. We investigated the extent of intramolecular backbone amide hydrogen (1H/2H) migration upon ECD using peptides with a unique selective deuterium incorporation. Our results show that only limited amide hydrogen migration occurs upon ECD, provided that vibrational excitation prior to the electron capture event is minimized. Peptide ions that are excessively vibrationally excited in the electrospray ion source by, e.g., high declustering potentials or during precursor ion selection (via sideband excitation) in the external linear quadrupole ion trap undergo nearly complete hydrogen (1H/2H) scrambling. Similarly, collision-induced dissociation (CID) in the external linear quadrupole ion trap results in complete or extensive hydrogen (1H/2H) scrambling. This precludes the use of CID as a method to obtain site-specific information from proteins that are labeled in solution-phase 1H/2H exchange experiments. In contrast, the deuteration levels of the c- and z-fragment ions generated from ECD closely mimic the known solution deuteration pattern of the selectively labeled peptides. This excellent correlation between the results obtained from gas phase and solution suggests that ECD holds great promise as a general method to obtain single residue resolution in proteins from solution 1H/2H exchange experiments.  相似文献   

19.
We examined the fragmentation of the electrospray-produced [M-H]- and [M-2H]2- ions of a number of peptides containing two acidic amino acid residues, one being aspartic acid (Asp) or glutamic acid (Glu), and the other being cysteine sulfinic acid [C(SO2H)] or cysteine sulfonic acid [C(SO3H)], on an ion-trap mass spectrometer. We observed facile neutral losses of H2S and H2SO2 from the side chains of cysteine and C(SO2H), respectively, whereas the corresponding elimination of H2SO3 from the side chain of C(SO3H) was undetectable for most peptides that we investigated. In addition, the collisional activation of the [M-H]- ions of the C(SO2H)-containing peptides resulted in the cleavage of the amide bond on the C-terminal side of the C(SO2H) residue. Moreover, collisional activation of the [M-2H]2- ions of the above Asp-containing peptides led to the cleavage of the backbone N-Calpha bond of the Asp residue to give cn and/or its complementary [zn-H2O] ions. Similar cleavage also occurred for the singly deprotonated ions of the otherwise identical peptides with a C-terminal amide functionality, but not for the [M-H]- ions of same peptides with a free C-terminal carboxylic acid. Furthermore, ab initio calculation results for model cleavage reactions are consistent with the selective cleavage of the backbone N-Calpha bond in the Asp residue.  相似文献   

20.
18O labeling: a tool for proteomics.   总被引:3,自引:0,他引:3  
An evaluation of the proteolytic labeling and quantification of proteins for diagnostic purposes using trypsin and 18O-enriched H2O is presented. We demonstrate that comparative or relative quantitation can be performed effectively with this approach. We have developed a protocol that allows the conservation of the labeled peptides in natural abundance water without fear of back-exchange providing that pH is sufficiently low to quench the catalytic activity of trypsin, but not so low as to promote chemical back-exchange. Because the labeling efficiency depends on the nature of the peptide, a simple linear relationship between the relative 16O/18O digest buffer mixture content (x) and labeling efficiency (y) does not exist; rather it follows a probability based y = x(2) relationship. As such, the extent of peptide labeling using 16O/18O digest buffer mixture ratios may deviate significantly from that expected based on a linear relationship. The evaluation of the relative Ziptip efficiency indicated a loss in sample recovery as the peptide concentration was reduced using normal conditions, suggesting that there is a limit below which there are diminishing returns. In addition, the adsorptive losses due to Speedvac dry down and recovery indicated modest (20%) losses that may vary widely (0-50%) from peptide to peptide. The in-solution digestion efficiency of standard protein mixtures as a function of concentration revealed a linear decrease with decreasing concentration. This is consistent with enzyme kinetic effects and emphasizes a potential quantitation error that could arise when evaluating differential expression based on peptide detection. The results from our studies demonstrate the power of 18O labeling as an optimization tool for proteomics process development.  相似文献   

设为首页 | 免责声明 | 关于勤云 | 加入收藏

Copyright©北京勤云科技发展有限公司  京ICP备09084417号