Identification of Cross-Linked Peptides by High ... - ACS Publications

Jan 11, 2010 - systems are not amenable to high-resolution techniques such as ... (5) Petrotchenko, E. V.; Xiao, K.; Cable, J.; Chen, Y.; Dokholyan, N...
0 downloads 0 Views 2MB Size
Anal. Chem. 2010, 82, 909–916

Identification of Cross-Linked Peptides by High-Resolution Precursor Ion Scan Amadeu H. Iglesias, Luiz Fernando A. Santos, and Fa´bio C. Gozzo* Institute of Chemistry, University of Campinas, and Instituto Nacional de Ciencia e Tecnologia de Bioanalitica, CP 6154, 13083-970 Campinas, Sa˜o Paulo, Brazil Chemical cross-linking coupled to mass spectrometry analysis has become a realistic alternative to the study of proteins structure and interactions, especially when these systems are not amenable to high-resolution techniques such as protein crystallography or nuclear magnetic resonance. One of the main bottlenecks of this approach relies on the detection of cross-linked peptides, as they are usually present in substoichiometric amounts in complex samples. It was shown that one of the main fragmentation pathways of disuccinimidyl suberate (DSS) cross-linked peptides yields diagnostic ions, whose structure is composed of a rearranged lysine side chain and the spacer arm of the linker. In this report, we demonstrate the feasibility of detecting these modified peptides based on a precursor ion scan in a quadrupole time-offlight (Q-TOF) instrument. It was shown that the fragmentation of nonmodified tryptic peptides hardly generates ions with the same nominal mass of the diagnostic ions, making the precursor ion scan very specific to N-hydroxysuccinimide (NHS)-based cross-linkers. Moreover, the experimental setup is the same as in the case of a regular cross-linking experiment, not demanding any additional experimental steps that would increase sample handling. The results obtained with protein samples allowed us to propose an algorithm that could be implemented in a software to process data from cross-linking experiments in an automated and high-throughput way. Protein structure, although of vital importance to understand biochemical processes, represents a major analytical challenge as the use of high-resolution techniques is restricted to a relatively small number of proteins. Large interest exists, therefore, for the development of additional techniques that can obtain structural information of a large number of proteins and be conveniently performed. Recently, chemical cross-linking coupled to mass spectrometry (MS3D) has become an attractive alternative to interrogate protein structure and interactions when no highresolution method is applicable to the system of interest.1-3 In MS3D experiments, a protein or protein complex is subjected to the cross-linking reaction followed by enzymatic * To whom correspondence should be addressed. E-mail: fabio@ iqm.unicamp.br. (1) Back, J. W.; Jong, L.; Muijsers, A. O.; Koster, C. G. J. Mol. Biol. 2003, 331, 303–313. (2) Sinz, A. Mass Spectrom. Rev. 2005, 25, 663–682. (3) Lee, Y. J. Mol. Biosyst. 2008, 4, 816–823. 10.1021/ac902051q  2010 American Chemical Society Published on Web 01/11/2010

digestion and mass spectrometry analysis. The identification of cross-linked peptides can be used to reveal several structural features of proteins, such as solvent accessibility, protein dynamics, and distance constraints as well as interacting partners and interacting domains in protein complexes. As this technique is based on mass spectrometric analysis for the identification of cross-linked peptides, it inherits all the attractive features of mass spectrometry, such as high sensitivity, fast analysis, and data interpretation, as well as applicability to virtually any protein. The main bottleneck of this methodology, however, relies on the identification and correct assignment of cross-linked peptides, since these species are formed in substoichiometric amounts.1 Several strategies have been developed to circumvent this limitation, such as isotope-coded cross-linkers,4,5 reagents with affinity tags,6,7 digestion in a mixture of H216O and H218O,8,9 and cleavable cross-linkers.10,11 All these methods, however, require additional sample handling and in most of the cases additional steps to the cross-linking reaction itself, decreasing sample throughput sensitivity. As chemical cross-linkers behave similarly to post-translational modifications (PTM), it would be desirable to apply MS-based methods traditionally used for PTM in order to aid in the detection of those peptides. One of those methods, precursor ion scan (PIS) on a triple quadrupole instrument (and more recently on hybrid triple quadrupole linear ion trap), has been widely used in the study of phosphopeptides,12-14 site determination of aldehydes modification,15 identification and differentiation of N- and O-linked (4) Sinz, A. Angew. Chem., Int. Ed. 2007, 46, 660–662. (5) Petrotchenko, E. V.; Xiao, K.; Cable, J.; Chen, Y.; Dokholyan, N. V.; Borchers, C. H. Mol. Cell. Proteomics 2009, 8, 273–286. (6) Kang, S.; Mou, L.; Lanman, J.; Velu, S.; Brouillette, W. J.; Prevelige, P. E., Jr. Rapid Commun. Mass Spectrom. 2009, 23, 1719–1726. (7) Zhang, H.; Tang, X.; Munske, G. R.; Tolic, N.; Anderson, G. A.; Bruce, J. E. Mol. Cell. Proteomics 2009, 8, 409–420. (8) Gao, Q.; Doneanu, C. E.; Shaffer, S. A.; Adman, E. T.; Goodlett, D. R.; Nelson, S. D. J. Biol. Chem. 2006, 281, 20404–20417. (9) Gardsvoll, H.; Gilquin, B.; Du, M. H. L.; Me´ne`z, A.; Jorgensen, T. J. D.; Ploug, M. J. Biol. Chem. 2006, 281, 19260–19272. (10) Kasper, P. T.; Back, J. W.; Vitale, M.; Hartog, A. F.; Roseboom, W.; Koning, L. J.; Maarseveen, J. H.; Muijsers, A. O.; Koster, C. G.; Jong, L. ChemBioChem 2007, 8, 1281–1292. (11) Lu, Y.; Tanasova, M.; Borhan, B.; Reid, G. E. Anal. Chem. 2008, 80, 9279– 9287. (12) Huddleston, M. J.; Annan, R. S.; Bean, M. F.; Carr, S. A. J. Am. Soc. Mass Spectrom. 1993, 4, 710–717. (13) Neubauer, G.; Mann, M. Anal. Chem. 1999, 71, 235–242. (14) Williamson, B. L.; Marchese, J.; Morrice, N. A. Mol. Cell. Proteomics 2006, 5, 337–346. (15) Bolgar, M. S.; Gaskell, S. J. Anal. Chem. 1996, 68, 2325–2330.

Analytical Chemistry, Vol. 82, No. 3, February 1, 2010

909

Figure 1. Structure of the diagnostic ions of m/z 222.1494, 239.1759, and 305.2229.

oligosaccharides in glycoproteins,16 and detection of glycopeptides17 and glycoproteins.18 In 2002, Bateman et al.19 proposed the use of a quadrupole time-of-flight (Q-TOF) instrument on a pseudo-PIS experiment in which, instead of scanning the quadrupole, alternate MS spectra were acquired with and without fragmentation. In this experiment, the quadrupole operates in the wide-band mode, while fragmentation is induced by increasing the collision energy. If the fragment ion of interest is formed, then a regular product ion scan could be acquired. This would capitalize on the advantages of this equipment, i.e., high speed of spectra acquisition associated with good sensitivity, resolution, and mass accuracy in both MS and MS/MS modes. More recently, Niggeweg et al. improved this method making it compatible with chromatographic separation.20 PIS on a Q-TOF instrument has been used so far in the identification and quantitation of phosphopeptides,21,22 detection and differentiation of arginine dimethylation,23 and on the detection of glycosylated peptides.24 PIS-based analysis depends on the presence of a reporter ion in the MS/MS spectra, i.e., a fragment that is formed in the collision cell for a class of compounds. In a previous paper by Seebacher et al., it was reported there was the formation of an ion of m/z 222 in experiments of peptides containing disuccinimidyl suberate (DSS) as the cross-linker.25 Our group has explored the fragmentation of those species, demonstrating its formation pathway.26,27 This ion (Figure 1) is formed by one rearranged lysine residue connected to the DSS moiety in its acylium form. (16) Carr, S. A.; Huddlestone, M. J.; Bean, M. F. Protein Sci. 1993, 2, 183–196. (17) Huddlestone, M. J.; Bean, M. F.; Carr, S. A. Anal. Chem. 1993, 65, 877– 884. (18) Haynes, P. A.; Aebersold, R. Anal. Chem. 2000, 72, 5402–5410. (19) Bateman, R. H.; Carruthers, R.; Hoyes, J. B.; Jones, C.; Langridge, J. I.; Millar, A.; Vissers, J. P. C. J. Am. Soc. Mass Spectrom. 2002, 13, 792–803. (20) Niggeweg, R.; Ko ¨cher, T.; Gentzel, M.; Buscaino, A.; Taipale, M.; Akhtar, A.; Wilm, M. Proteomics 2006, 6, 41–53. (21) Steen, H.; Ku ¨ ster, B.; Fernandez, M.; Pandey, A.; Mann, M. Anal. Chem. 2001, 73, 1440–1448. (22) Paradela, A.; Albar, J. P. J. Proteome Res. 2008, 7, 1809–1818. (23) Rappsilber, J.; Friesen, W. J.; Paushkin, S.; Dreyfuss, G.; Mann, M. Anal. Chem. 2003, 75, 3107–3114. (24) Frolov, A.; Hoffmann, P.; Hoffmann, R. J. Mass Spectrom. 2006, 41, 1459– 1469. (25) Seebacher, J.; Mallick, P.; Zhang, N.; Eddes, J. S.; Aebersold, R.; Gelb, M. H. J. Proteome Res. 2006, 5, 2270–2282. (26) Iglesias, A. H.; Santos, L. F. A.; Gozzo, F. C. J. Am. Soc. Mass Spectrom. 2009, 20, 557–566. (27) Santos, L. F. A.; Iglesias, A. H.; Gozzo, F. C. J. Am. Soc. Mass Spectrom. 2010, submitted for publication.

910

Analytical Chemistry, Vol. 82, No. 3, February 1, 2010

Moreover, we have also shown the formation of two other ions (m/z 239 and 305) which can also be used as reporter ions for DSS-containing peptides. Previous reports demonstrate the ability of different cross-linkers to generate gas-phase fragments which indicate the presence of this reagent in the structure of the selected ion. Back et al. proposed the use of a benzyl-derivatized disuccinimidyl ester as a cross-linker.28 Upon low-energy fragmentation, this reagent yields a stable benzyl cation marker ion which could be used to detect the presence of cross-linked species. Despite its simplicity, this reagent was used only for model peptides, and its use in the case of proteins was not shown so far. Tang et al. developed a new set of cross-linkers derived from a peptide amidation reagent which presents labile bonds upon dissociation.29 Upon dissociation, this protein interaction reporter (PIR) yields not only a reporter ion, indicating the presence of the cross-linker, but as well the two intact peptides. However, practical aspects such as the commercial unavailability and difficult synthesis, large size of the reagents, and the need for MS3 for the identification of the connected peptides have diminished the interest in such cross-linkers. More recently, Soderblom and Goshe developed a dissociative cross-linking reagent (collision-induced dissociative chemical cross-linking, CID-CXL) in which the spacer arm contains the labile dipeptide Asp-Pro.30 As in the previous case, not only are those reagents not easy to obtain but they also need MS3 spectra to confirm the nature of the peptides. Moreover, if at least one of the peptides contains an Asp-Pro bond, the identification of the cross-linked peptides will be even more troublesome. Therefore, in the present work we evaluate the detection and identification of cross-linked peptides with DSS, the most used commercially available cross-linker in MS3D experiments, by means of a PIS experiment. EXPERIMENTAL SECTION Materials. DSS and sequencing-grade modified porcine pancreas trypsin were obtained from Pierce and Promega, respectively. Peptides AGAKGAERLVKAGVR (PX), Ac-ARKGCREVTKNDLR (P1), Ac-ARGKWPREVKIHR (P2), and Ac-ARYTKDLSQRAFKGMR (P3) were obtained from Proteimax (Sa˜o Paulo, Brazil). Chicken egg lysozyme was obtained from Sigma-Aldrich. All other reagents were obtained from Sigma-Aldrich and Tedia and used without further purification. Cross-Linking Reactions. Cross-linking reactions were performed as previously described.26,27 Briefly, peptides were incubated in phosphate buffer 50 mM pH 7.5 with DSS in a 1:50 ratio for 2 h, followed by quenching with Tris buffer pH 7.6 and trypsin digestion for 3 h. Some of the samples were spiked with a bovine hemoglobin trypsin digestion standard (Waters Co.) in a 1:1 molar ratio. Collision Energy Determination. In order to determine the optimal collision energy to the formation of reporter ions, MS/ MS spectra were acquired for different DSS-containing peptides varying the collision energy. Spectra were averaged for 10 s. (28) Back, J. W.; Hartog, A. F.; Dekker, H. L.; Muijsers, A. O.; Koning, L. J.; Jong, L. J. Am. Soc. Mass Spectrom. 2001, 12, 222–227. (29) Tang, X.; Munske, G. R.; Siems, W. F.; Bruce, J. E. Anal. Chem. 2005, 77, 311–318. (30) Soderblom, E. J.; Goshe, M. B. Anal. Chem. 2006, 78, 8059–8068.

Liquid Chromatography/Mass Spectrometry Analysis. LC/MS/MS analyses were performed in a Q-TOF Ultima (Waters, Milford) coupled online to a nanoAcquity ultraperformance liquid chromatography (UPLC) system. An amount of 1 pmol of each peptide was loaded and desalted on a 180 µm × 20 mm Waters Symmetry C18. After the desalting step, sample was directed to a 100 µm × 100 mm Waters BEH130 C18 column at a flow rate of 1.1 µL/min. Mobile phases A and B consisted of 0.1% formic acid/ water and 0.1% formic acid/acetonitrile, respectively. The gradient conditions used are as follows: 0 min with 10% of B, then it linearly increased to 40% B in 25 min, then it increased up to 70% B in 28 min where it remained until 40 min and in the next minute it was decreased to 10% of B. Typical operating conditions of the mass spectrometer in PIS experiments are as follows: 3.5 kV (capillary voltage), 100 V (cone voltage), 100 °C (source temperature), and 10 eV/25-50 eV for the low and high collision energies. Anytime there was a fragment of m/z 222.1, 239.1, or 305.2 ± 0.2 Da, a product ion scan was acquired of the five most intense ions in the MS spectrum. RESULTS AND DISCUSSION Collision Energy Optimization. In previous works,26,27 the study of fragmentation patterns of both intra- and intermolecular cross-linked peptides revealed a set of diagnostic ions present whenever the peptide is bound to cross-linkers (Figure 1). For DSS, for example, fragment ions of m/z 222.1494 and 239.1759 are observed whenever the DSS molecule is attached to a peptide, being it an intra- or intermolecular species as well as a dead-end. The ion of m/z 222.1494 corresponds to the loss of ammonia from m/z 239.1759. The fragment of m/z 305.2229, on the other hand, is specific to species where two lysine side chains are connected by DSS, that is, intra- or intermolecular cross-linked peptides. This set of diagnostic ions can, therefore, be used in PIS experiments to identify cross-linked peptides in the presence of nonmodified, regular species present in protein digests. Another remarkable feature of this set of ions is that no isobaric a, b, or y fragments exist for the 20 common amino acids. This selectivity, coupled to the high mass accuracy of TOF analyzers in both MS and MS/ MS modes, should confer high specificity for the identification of cross-linked peptide. In the previous reports on the fragmentation of cross-linked peptides, the relative intensity of these diagnostic ions depended strongly on the collision energy, so this parameter was optimized using model cross-linked peptides (Scheme 1).26,27 The intensity plots of the diagnostic ions as a function of collision energy for intra- and intermolecular species with 2 and 3 charges were determinate (plots for m/z 222, 239, and 305 can be found in Supplementary Figures 1-3, respectively, in the Supporting Information). Formation of m/z 222 is strongly favorite at lower collision energy in the case of the N-terminus cross-link (55 eV × 30 eV, Supporting Information Supplementary Figure 1A). It is not possible, however, to attribute this difference on the type of cross-linking, since these peptides have very different masses. When the precursor masses are similar, the collision energy apexes are very close, indicating that a common collision energy can be applied as a function of m/z, just like regular peptides. This trend is also true for 2+ and 3+ intermolecular species (Supporting Information Supplementary Figure 1, parts C and D, respectively) as well as for the two other marker ions of m/z 239

Scheme 1. (A) Schematic Representation of the Designed Peptides, Followed by Reaction with DSS and Trypsin Digestion and (B) Intra- and Intermolecular Species Formed from Synthetic Peptides Used in This Studya

a “x” presents a variable number and type of amino acid except lysine and arginine.

and 305 (Supporting Information Supplementary Figures 2 and 3). It is also worth mentioning that the 2+ intramolecular species of m/z 449.2 do not yield fragment ion m/z 305 at any collision energy as expected, since in this case the lysine residue is connected to the N-terminus of the peptide and not to another lysine. On the basis of these results, all other MS/MS spectra were acquired using the collision energy proportional to peptide m/z, which varied from 25 to 50 eV. PIS on Cross-Linked Samples. The first sample analyzed by PIS on the Q-TOF was PX (Figure 2). This peptide contains three possible sites for cross-linking: two lysines and the free N-terminus. As can be seen in the low-energy chromatogram of the DSS reaction product, the major species present are the two K-K (inter) and the K-N-terminus (intra- and intermolecular) cross-linked peptides. Non-cross-linked digested peptides and dead-end (non) digested species account for the other peaks observed. The extracted ion current (XIC) for the diagnostic ions in the high-energy chromatograms are shown in Figure 3. The first observation is the higher number of species in the XIC for the ions of m/z 222 and 239 when compared to 305. This is in agreement to the fact that m/z 305 is only generated from intraor intermolecular cross-linked peptides, whereas m/z 222 and 239 are generated from all species containing the cross-linker moiety. It can be also noticed in the chromatogram for m/z 222 ion that the only intense species corresponds to the intermolecular crosslinked peptide; the other low-intensity signals were due to the presence of M + 1 from m/z 304 ions. The XIC for m/z 222 and 239, on the other hand, presents several species in addition to the peptides mentioned above. These ions were mainly generated from dead-end species. Owing to the intrinsic complex nature of the cross-linking reaction, there are several possibilities of deadend peptides in this type of experiment: (i) single dead-end tryptic peptide, (ii) single dead-end with a missed cleavage site, (iii) multiple dead-end (two or three sites) tryptic peptides, and (iv) multiple dead-ends with a missed cleavage site. This large number of possibilities explains the large number of signals detected in this experiment. Analytical Chemistry, Vol. 82, No. 3, February 1, 2010

911

Figure 2. Low-energy chromatogram for peptide PX after reaction with DSS and trypsin digestion.

Figure 3. Extracted ion current for the three diagnostic ions for sample PX after reaction with DSS and trypsin digestion: m/z 222 (A), 239 (B), and 305(C).

The superimposed chromatograms of m/z 222 and 239 (Supporting Information Supplementary Figure 4) compare the intensity of both ions. In general, m/z 222 presents higher intensity then m/z 239, indicating that the loss of ammonia from m/z 239 is favorable in the energetic conditions used. For some low-intensity precursors, however, m/z 239 presents a higher intensity, and in some cases 912

Analytical Chemistry, Vol. 82, No. 3, February 1, 2010

only one of the ions is generated, as in the case of precursors with retention times of 14.5 and 19.4 min. Therefore, simultaneous monitoring of both ions is required for more comprehensive analysis. This approach is more advantageous in Q-TOF type instruments (compared to tandem quadrupoles) because the addition of an additional monitoring channel does not lead to sensitivity loss.

Figure 4. Extracted ion currents of the diagnostic ions (curves A, B, and C correspond, respectively, to m/z 222, 239, and 305) for samples containing hemoglobin trypsin digest without (lower lines) and with (upper lines) the addition of PX reacted with DSS and digested with trypsin.

The next step consisted of simulating a sample obtained in experiments with protein complexes. In this case, we spiked the PX sample with an equimolar amount of trypsin-digested bovine hemoglobin. This protein is composed by four polypeptide chains with a total molecular weight of approximately 65 kDa and would therefore constitute a reasonable “chemical background” for our experiments. Comparison of the low-energy chromatograms of hemoglobin and hemoglobin + PX (Supporting Information Supplementary Figure 5) shows that PX peptides elute earlier and the marker ions should be present mostly in the 10-15 min elution time range, whereas regular hemoglobin peptides are eluted later and no marker ions are expected during this elution time. Figure 4 presents the normalized XIC for the diagnostic ions in the digest hemoglobin with and without PX peptides. It is clearly evident that none of the three marker ions are present in the hemoglobin sample, showing an excellent selectivity for crosslinked peptides. For the hemoglobin + PX peptides sample, on the other hand, all the three marker ions are present in the elution time range expected for the PX peptides (10-15 min). Besides, it can be noted that the XIC of the marker ions is exactly the same as that of the pure PX peptides (Figure 3), showing that matrix effects can be almost neglected, especially when coupled to a high-resolution chromatographic system (UPLC in this case), causing no considerable intensity loss when compared to the experiment of pure PX peptides.

A striking feature of these marker ions is the high selectivity. The XIC of m/z 222, 239, and 305 for digested hemoglobin with PX peptides extracted with a width of 0.02 and 1 Da (Supporting Information Supplementary Figure 6) show no major differences in the chromatogram but only a slightly higher background noise for m/z 305. The same experiment performed with lysozyme displayed the same behavior (data not shown). This makes the PIS method also applicable in low mass accuracy instruments, like quadrupoles and ion traps. The PIS approach was applied to an experiment where lysozyme was cross-linked with DSS. As in the previous case, the low-energy chromatograms for the samples with and without cross-linker show a clear difference in complexity (data not shown). XIC for ions of m/z 222, 239, and 305 once again demonstrate the high selectivity of this approach toward crosslinked species, since no detectable chromatographic peaks are observed in the absence of DSS (Figure 5). This is even clearer when both diagnostic ions for the presence of DSS (m/z 222 and 239) are analyzed simultaneously, which can be done in a TOF instrument without sensitivity loss. Regular precursor ion spectra were acquired in order to confirm the sequence of the modified peptides. For example, a small chromatographic peak for both ions with retention time of 16.5 min (Figure 6A) can be seen. Figure 6B shows the low-energy Analytical Chemistry, Vol. 82, No. 3, February 1, 2010

913

Figure 5. Extracted ion currents of the diagnostic ions (curves A, B, and C correspond, respectively, to m/z 222, 239, and 305) for lysozyme digested with trypsin, reacted (upper lines) or not (lower lines) with DSS.

Figure 6. (A) Extracted ion current for m/z 222.1 and 239.1 for a lysozyme sample modified by DSS. The gray rectangle corresponds to the chromatographic peak summed up to generate the MS spectrum shown in panel B. In panel C is shown the product ion scan for m/z 381.7, in which the inset (D) shows the presence of m/z 222.1 and 239.1 but not 305.2. 914

Analytical Chemistry, Vol. 82, No. 3, February 1, 2010

Figure 7. General flowchart showing the algorithm for DSS homologous cross-linkers. SC represents the mass of the spacer arm chain of the reagent.

MS spectrum for this retention time, and Figure 6C shows the product ion scan for m/z 381.7, a dead-end type peptide. The fragmentation pattern of this peptide allows the identification of the sequence KVFGR, where the first residue is the site of modification. This peptide sequence corresponds to the Nterminus of the protein, and therefore the modified residue is the first amino acid in the protein structure. The crystallographic structure of lysozyme (PDB ID 1W6Z) clearly shows that this residue is very solvent-exposed and therefore prone to reaction with DSS. As described above, the proposed approach for the detection of cross-linked peptides is very simple, once it does not require any additional experimental handling or reaction of the sample, making it automatable for high-throughput cross-linking experiments. The next step to make this even more user friendly would be to create an automatic pipeline for data interpretation to be implemented in software. The idea presented above could also be applied to other N-hydroxysuccinimide (NHS)-based cross-linkers. Indeed, we showed that cross-linkers homologous to DSS (DSG and DSSeb, with spacer chains containing 5 and 10 C atoms) fragment in a similar way to DSS, yielding diagnostic ions analogous to the ones presented above.27 In all cases, the mass of these ions is determined only by the rearranged side chain of the lysine residue (83.0735) and the mass of the spacer arm of the cross-linker (Figure 1). The same experiment performed with lysozyme and DSS was done also with DSSeb and DSG (Supporting Information Supplementary Figures 7 and 8), and results were very similar: owing to the high resolution of the instrument, it was possible to generate XIC for the analogous diagnostic ions with low noise levels, accounting for the high intensity of the signals in the

chromatogram. Supporting Information Supplementary Figures 7A and 8A illustrate the XIC for ions of m/z 250 and 180, which are homologous to m/z 222 (obtained from the experiments with DSS). The same was observed for the other two marker ions for DSSeb (m/z 267 and 333, in Supporting Information Supplementary Figure 7, parts B and C) and DSG (m/z 197 and 263, in Supporting Information Supplementary Figure 8, parts B and C), analogous to m/z 239 and 305 from DSS experiments. As a result, a general software algorithm is proposed, in which diagnostic ion masses are calculated based on the cross-linker spacing chain (Figure 7). Basically, this algorithm would evaluate every product ion spectrum looking for the presence of the diagnostic ions m/z SC + 83 or + 100 (within a certain mass accuracy), where SC is the mass of the cross-linker spacing chain; in the case of absence of these ions, this spectrum is discarded once it corresponds to a peptide without the cross-linker moiety. The next step is to evaluate the presence of m/z 305; its absence can indicate (i) dead-end type peptide, (ii) cross-link between K and N-terminus, or (iii) cross-link between K and another residue. In case m/z 305 is present, this ion corresponds to K-K crosslinked peptide and, therefore, corresponds to an intra- or intermolecular specie. In order to differentiate between intra- and intercross-linked species, the presence of y1 ions is verified: if it has both m/z 147.1 (y1 from Lys residue) and 175.1 (y1 from Arg residue), it means that this ion has two C-termini, i.e., an intermolecular cross-link. If it presents only one of this ions it can be (i) intramolecular peptide, (ii) intermolecular peptide in which both chains have the same C-terminus, or (iii) intermolecular peptide in which one of the cross-linked peptides is the C-terminus of the protein and therefore its last residue can be different from K and R. After this filtering step, the software would Analytical Chemistry, Vol. 82, No. 3, February 1, 2010

915

interpret product ion spectra in order to try to get the sequence of the linked peptide(s), based on the fragmentation patterns proposed previously.26,27 CONCLUSION In this report an approach based on PIS on a Q-TOF instrument to detect cross-linked peptides was demonstrated. As previously shown, DSS-linked species upon CID yield diagnostic ions which can be used in PIS experiments. In the case of DSS, we have shown that these ions are especially suitable for this approach, once there are not any y, a, and b fragment ions with same mass, resulting in experiments with high specificity. We also have proposed an algorithm for the automated analysis of product ion spectra of candidate cross-linked ions. In this approach, precursor ions are filtered by the presence of the diagnostic ions in its fragmentation spectra, which is thereafter interpreted according to the fragmentation patterns proposed in previous works. The method is demonstrated to work with other NHS-based crosslinkers as well as other instrument types, like quadrupoles and ion traps. The PIS approach is simple, does not require any

916

Analytical Chemistry, Vol. 82, No. 3, February 1, 2010

additional sample handling step, and is very selective, having the potential to overcome the problem of cross-linked peptide identification, one of the greatest bottlenecks in MS3D. ACKNOWLEDGMENT The first two authors have contributed equally to the work. This work was supported by the Sa˜o Paulo Proteome Network (FAPESP 2004/14846-0 and FINEP 01.07.0290.00) and Instituto Nacional de Ciencia e Tecnologia de Bioanalitica (CNPq 573672/ 2008-3 and FAPESP 08/57805-2). The authors also acknowledge other support from FAPESP (2007/55930-1), CAPES, and CNPq. SUPPORTING INFORMATION AVAILABLE Additional information as noted in text. This material is available free of charge via the Internet at http://pubs.acs.org.

Received for review September 11, 2009. Accepted December 8, 2009. AC902051Q