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.
The origin of asymmetric charge and mass partitioning observed for gas-phase dissociation of multiply charged macromolecular complexes has been hotly debated. These experiments hold the potential to provide detailed information about the interactions between the macromolecules within the complex. Here, this unusual phenomenon of asymmetric charge partitioning is investigated for several protein homodimers. Asymmetric charge partitioning in these ions depends on a number of factors, including the internal energy, charge state, and gas-phase conformation of the complex, as well as the conformational flexibility of the protein monomer in the complex. High charge states of both cytochrome c and disulfide-reduced alpha-lactalbumin homodimers dissociate by a symmetrical charge partitioning process in which both fragment monomers carry away roughly an equal number of charges. In contrast, highly asymmetric charge partitioning dominates for the lower charge states. Cytochrome c dimer ions with eleven charges formed by electrospray ionization from two solutions in which the solution-phase conformation differs dissociate with dramatically different charge partitioning. These results demonstrate that these gas-phase complexes retain a clear "memory" of the solution from which they are formed, and that information about their solution-phase conformation can be obtained from these gas-phase dissociation experiments. Cytochrome c dimer ions formed from solutions in which the conformation of the protein is native show greater asymmetric charge partitioning with increasing ion internal energy. Cytochrome c dimers that are conformationally constrained with intramolecular cross-linkers undergo predominantly symmetric charge partitioning under conditions where asymmetric charge partitioning is observed for cytochrome c dimers without cross-links. Similar results are observed for alpha-lactalbumin homodimers. These results provide convincing evidence that the origin of asymmetric charge partitioning in these homodimers is the result of one of the protein monomers unfolding in the dissociation transition state. A mechanism that accounts for these observations is proposed.  相似文献   

2.
The influence of charge on the thermal dissociation of gaseous, protonated, homodimeric, protein ecotin ions produced by nanoflow electrospray ionization (nanoES) was investigated using the blackbody infrared radiative dissociation technique. Dissociation of the protonated dimer, (E2 + nH)n+ E2n+ where n = 14–17, into pairs of monomer ions is the dominant reaction at temperatures from 126 to 175 °C. The monomer pair corresponding to the most symmetric charge distribution is preferred, although 50–60% of the monomer product ions correspond to an asymmetric partitioning of charge. The relative abundance of the different monomer ion pairs produced from E214+, E215+, and E216+ depends on reaction time, with the more symmetric charge distribution pair dominating at longer times. The relative yield of monomer ions observed late in the reaction is independent of temperature indicating that proton transfer between the monomers does not occur during dissociation and that the different monomer ion pairs are formed from dimer ions which differ in the distribution of charge between the monomers. For E217+, the yield of monomer ions is independent of reaction time but does exhibit slight temperature dependence, with higher temperatures favoring the monomers corresponding to most symmetric charge distribution. The charge distribution in the E215+ and E216+ dimer ions influences the dissociation kinetics, with the more asymmetric distribution resulting in greater reactivity. In contrast, the charge distribution has no measurable effect on the dissociation kinetics and energetics of the E217+ dimer.  相似文献   

3.
Upon collisional activation, a series of DNA duplexes exhibited a significant degree of asymmetric dissociation with respect to charge partitioning among the single strands. That is, the charge states of the single strand product ions did not equal q/2 for even precursor charge states or (q + 1)/2 and (q − 1)/2 for odd precursor charge states (where q is the charge of the precursor). The factors that affect this asymmetric charge partitioning were assessed. The smaller, lower charged duplexes resulted in more symmetric dissociation compared with larger duplexes in higher charge states, which displayed a high degree of asymmetry upon dissociation. The composition of the duplexes influenced charge partitioning, with those containing a greater number of A/T base pairs showing more symmetric dissociation relative to the more G/C rich duplexes. The use of higher collisional energies resulted in significantly more asymmetric dissociation. Comparisons were made with the dissociation behavior previously studied for protein noncovalent complexes and past studies of the gas-phase conformations and dissociation of DNA complexes.  相似文献   

4.
The ‘softness’ of the electrospray ionization (ESI) method provides a direct link between solution chemistry and the inherent gas-phase environment of mass Spectrometry. Available results related to the preservation of non-covalent associations into the gas phase after ESI are reviewed. These associations include the possible retention of elements of higher order protein structure, non-covalent polypeptide–heme associations and enzyme complexes. Experimental results are presented showing that non-covalently bound polypeptide and protein dimer ions are relatively common as low level contributions to ESI mass spectra. It is argued that these dimers are reflective of multimeric species in solution since Coulombic barriers preclude dimerization after ESI although uncertainty remains regarding whether they exist prior to the formation of highly charged droplets. The dissociation of dimers is facile and for proteins can yield monomers having a broad distribution of charge states. The detection of non-covalently associated dimers requires gentle ESI mass spectrometer interface conditions, yielding relatively low levels of internal excitation. Under such conditions incomplete molecular ion desolvation can result in experimental artifacts for tandem mass spectrometric experiments. ESI mass Spectrometry may have broad potential for the study of noncovalent liquid phase associations.  相似文献   

5.
The assembly of the B subunits of Shiga toxins (Stx) 1 and 2 and the influence of solution conditions (protein concentration, temperature, pH, and ionic strength) on it are investigated using temperature-controlled nanoflow electrospray (nano-ES) ionization and Fourier-transform ion cyclotron resonance mass spectrometry. Despite the similar higher order structure predicted by X-ray crystallography analysis, the B(5) homopentamers of Stx1 and Stx2 exhibit differences in stability under the solution conditions investigated. At solution temperatures ranging from 0 to 60 degrees C and subunit concentrations ranging from 5 to 85 microM, the Stx1 B subunit exists almost entirely as the homopentamer in aqueous solutions, independent of the ionic strength. In contrast, the degree of assembly of Stx2 B subunit is strongly dependent on temperature, subunit concentration, and ionic strength. At subunit concentrations of more than 50 microM, the Stx2 B subunit exists predominantly as a pentamer, although smaller multimers (dimer, trimer, and tetramer) are also evident. At lower concentrations, the Stx2 B subunit exists predominantly as monomer and dimer. The relative abundance of multimeric species of the Stx2 B subunit was insensitive to the ion source conditions, suggesting that gas-phase dissociation of the pentamer ions in the source does not influence the mass spectrum. Blackbody infrared radiative dissociation of the protonated B(5) ions of Stx2 at the +12 and +13 charge states proceeds, at reaction temperatures of 120 to 180 degrees C, predominantly by the ejection of a single subunit from the complex. Dissociation into dimer and trimer ions constitutes a minor pathway. It follows that the dimer and trimer ions and, likely, the monomer ions observed in the nano-ES mass spectra of Stx2 B subunit originated in solution and not from gas-phase reactions. It is concluded that, under the solution conditions investigated, the homopentamer of Stx2 B subunit is thermodynamically less stable than that of Stx1 B subunit. Arrhenius activation parameters determined for the protonated Stx2 B(5) ions at the +12 and +13 charge states were compared with values reported for the corresponding B(5) ions of Stx1 B subunit. In contrast to the differential stability of the Stx1 and Stx2 B pentamers in solution, the dissociation activation energies (E(a)) determined for the gaseous complexes are indistinguishable at a given charge state. The similarity in the E(a) values suggests that the protonated pentamer ions of both toxins are stabilized by similar intersubunit interactions in the gas phase, a result that is in agreement with the X-ray crystal structures of the holotoxins.  相似文献   

6.
The effects of binding two small-molecule inhibitors to Agrobacterium sp. strain ATCC 21400 (Abg) β-glucosidase on the conformations and stability of gas-phase ions of Abg have been investigated. Biotin-iminosugar conjugate (BIC) binds noncovalently to Abg while 2,4-dinitro-2-deoxy-2-fluoro-β-d-glucopyranoside (2FG-DNP) binds covalently with loss of DNP. In solution, Abg is a dimer. Mass spectra show predominantly dimer ions, provided care is taken to avoid dissociation of dimers in solution and dimer ions in the ion sampling interface. When excess inhibitor, either covalent or noncovalent, is added to solutions of Abg, mass spectra show peaks almost entirely from 2:2 inhibitor-enzyme dimer complexes. Tandem mass spectrometry experiments show similar dissociation channels for the apo-enzyme and 2FG-enzyme dimers. The +21 dimer produces +10 and +11 monomers. The internal energy required to dissociate the +21 2FG-enzyme to its monomers (767?±?30 eV) is about 36 eV higher than that for the apo-enzyme dimer (731?±?6 eV), reflecting the stabilization of the free enzyme dimer by the 2FG inhibitor. The primary dissociation channels for the noncovalent BIC-enzyme dimer are loss of neutral and charged BIC. The internal energy required to induce loss of BIC is 482?±?8 eV, considerably less than that required to dissociate the dimers. For a given charge state, ions of the covalent and noncovalent complexes have about 15 % and 25 % lower cross sections, respectively, compared with the apo-enzyme. Thus, binding the inhibitors causes the gas-phase protein to adopt more compact conformations. Noncovalent binding surprisingly produces the greatest change in protein ion conformation, despite the weaker inhibitor binding.
Figure
?  相似文献   

7.
The conformations of gas-phase ions of hemoglobin, and its dimer and monomer subunits have been studied with H/D exchange and cross section measurements. During the H/D exchange measurements, tetramers undergo slow dissociation to dimers, and dimers to monomers, but this did not prevent drawing conclusions about the relative exchange levels of monomers, dimers, and tetramers. Assembly of the monomers into tetramers, hexamers, and octamers causes the monomers to exchange a greater fraction of their hydrogens. Dimer ions, however, exchange a lower fraction of their hydrogens than monomers or tetramers. Solvation of tetramers affects the exchange kinetics. Solvation molecules do not appear to exchange, and solvation lowers the overall exchange level of the tetramers. Cross section measurements show that monomer ions in low charge states, and tetramer ions have compact structures, comparable in size to the native conformations in solution. Dimers have remarkably compact structures, considerably smaller than the native conformation in solution and smaller than might be expected from the monomer or tetramer cross sections. This is consistent with the relatively low level of exchange of the dimers.  相似文献   

8.
The formation and dissociation of dimer complexes consisting of a transition metal ion and two polyether ligands is examined in a quadrupole ion trap mass spectrometer. Reactions of three transition metals (Ni, Cu, Co) with three crown ethers and four acyclic ethers (glymes) are studied. Singly charged species are created from ion-molecule reactions between laser-desorbed monopositive metal ions and the neutral polyethers. Doubly charged complexes are generated from electrospray ionization of solutions containing metal salts and polyethers. For the singly charged complexes, the capability for dimer formation by the ethers is dependent on the number of available coordination sites on the ligand and its ability to fully coordinate the metal ion. For example, 18-crown-6 never forms dimer complexes, but 12-crown-4 readily forms dimers. For the more flexible acyclic ethers, the ligands that have four or more oxygen atoms do not form dimer complexes because the acyclic ligands have sufficient flexibility to wrap around the metal ion and prevent attachment of a second ligand. For the doubly charged complexes, dimers are observed for all of the crown ethers and glymes, thus showing no dependence on the flexibility or number of coordination sites of the polyether. The nonselectivity of dimer formation is attributed to the higher charge density of the doubly charged metal center, resulting in stronger coordination abilities. Collisionally activated dissociation is used to evaluate the structures of the metal-polyether dimer complexes. Radical fragmentation processes are observed for some of the singly charged dimer complexes because these pathways allow the monopositive metal ion to attain a more favorable 2 + oxidation state. These radical losses are observed for the dimer complexes but not for the monomer complexes because the dimer structures have two independent ligands, a feature that enhances the coordination geometry of the complex and allows more flexibility for the rearrangements necessary for loss of radical species. Dissociation of the doubly charged complexes generated by electrospray ionization does not result in losses of radical neutrals because the metal ions already exist in favorable 2+ oxidation states.  相似文献   

9.
The dissection of specific and nonspecific protein complexes in the gas phase is studied by collisionally activated decomposition. In particular, the gas phase dissection of multiple protonated homodimeric Human Galectin I, E. Coli Glyoxalase I, horse heart cytochrome c, and Hen egg Lysozyme have been investigated. Both the Human Galectin I and E. Coli Glyoxalase I enzymes are biologically active as a dimer, exhibiting molecular weights of approximately 30 kDa. Cytochrome c and Lysozyme are monomers, but may aggregate to some extent at high protein concentrations. The gas phase dissociation of these multiple protonated dimer assemblies does lead to the formation of monomers. The charge distribution over the two concomitant monomers following the dissociation of these multiple protonated dimers is found to be highly dissimilar. There is no evident correlation between the solution phase stability of the dimeric proteins and their gas-phase dissociation pattern. Additionally, in the collisionally activated decomposition spectra diffuse ion signals are observed, which are attributed to monomer ions formed via slow decay of the collisionally activated dimer ions inside the reflectron time-of-flight. Although, the formation of these diffuse metastable ions may complicate the interpretation of collisionally activated decomposition mass spectra, especially when studying noncovalent protein complexes, a simple mathematical equation may be used to reveal their origin and pathway of formation.  相似文献   

10.
In the gas-phase, ions of protein complexes typically follow an asymmetric dissociation pathway upon collisional activation, whereby an expelled small monomer takes a disproportionately large amount of the charges from the precursor ion. This phenomenon has been rationalized by assuming that upon activation, a single monomer becomes unfolded, thereby attracting charges to its newly exposed basic residues. Here, we report on the atypical gas-phase dissociation of the therapeutically important, heterodimeric calcium/calmodulin-dependent serine/threonine phosphatase calcineurin, using a combination of tandem mass spectrometry, ion mobility mass spectrometry, and computational modeling. Therefore, a hetero-dimeric calcineurin construct (62?kDa), composed of CNa (44?kDa, a truncation mutant missing the calmodulin binding and auto-inhibitory domains), and CNb (18?kDa), was used. Upon collisional activation, this hetero-dimer follows the commonly observed dissociation behavior, whereby the smaller CNb becomes highly charged and is expelled. Surprisingly, in addition, a second atypical dissociation pathway, whereby the charge partitioning over the two entities is more symmetric is observed. The presence of two gas-phase conformational isomers of calcineurin as revealed by ion mobility mass spectrometry (IM-MS) may explain the co-occurrence of these two dissociation pathways. We reveal the direct relationship between the conformation of the calcineurin precursor ion and its concomitant dissociation pathway and provide insights into the mechanisms underlying this co-occurrence of the typical and atypical fragmentation mechanisms.  相似文献   

11.
A series of calculations, varying from simple electrostatic to more detailed semi-empirical based molecular dynamics ones, were carried out on charged gas phase ions of the cytochrome c(') dimer. The energetics of differing charge states, charge partitionings, and charge configurations were examined in both the low and high charge regimes. As well, preliminary free energy calculations of dissociation barriers are presented. It is shown that one must always consider distributions of charge configurations, once protein relaxation effects are taken into account, and that no single configuration dominates. All these results also indicate that in the high charge limit, the dissociation of protein complex ions is governed by electrostatic repulsion from the net charges, the consequences of which are enumerated and discussed. There are two main trends deriving from this, namely that charges will move so as to approximately maintain constant surface charge density, and that the lowest barrier to dissociation is the one that produces fragment ions with equal charges. In particular, it is shown that the charge-to-mass ratio of a fragment ion is not the key physical parameter in predicting dissociation products. In fact, from the perspective of the division of total charge, many dissociation pathways reported to be "asymmetric" in the literature should be more properly labelled as "symmetric" or "near-symmetric". The Coulomb repulsion model assumes that the timescale for charge transfer is faster than that for protein structural changes, which in turn is faster than that for complex dissociation.  相似文献   

12.
The unfolding enthalpy of the native state of ubiquitin in solution is 5 to 8 times that of its gaseous ions, as determined by electron capture dissociation (ECD) mass spectrometry. Although two-state folding occurs in solution, the three-state gaseous process proposed for this by Clemmer and co-workers based on ion mobility data is supported in general by ECD mass spectra, including relative product yields, distinct Delta H(unfolding) values between states, site-specific melting temperatures, and folding kinetics indicating a cooperative process. ECD also confirms that the 13+ ions represent separate conformers, possibly with side-chain solvated alpha-helical structures. However, the ECD data on the noncovalent bonding in the 5+ to 13+ ions, determined overall in 69 of the 75 interresidue sites, shows that thermal unfolding proceeds via a diversity of intermediates whose conformational characteristics also depend on charge site locations. As occurs with increased acidity in solution, adding 6 protons to the 5+ ions completely destroys their tertiary noncovalent bonding. However, solvation of the newly protonated sites to the backbone instead increases the stability of the secondary structure (possibly an alpha-helix) of these gaseous ions, while in solution these new sites aid denaturation by solvation in the aqueous medium. Extensive ion equilibration can lead to even more compact and diverse conformers. The three-state unfolding of gaseous ubiquitin appears to involve ensembles of individual chain conformations in a "folding funnel" of parallel reaction paths. This also provides a further caution for characterizing solution conformers from their gas-phase behavior.  相似文献   

13.
Thermal dissociation experiments, implemented with blackbody infrared radiative dissociation and Fourier-transform ion cyclotron resonance mass spectrometry, are performed on gaseous protonated and deprotonated ions of the homopentameric B subunits of Shiga toxin 1 (Stx1 B5) and Shiga toxin 2 (Stx2 B5) and the homotetramer streptavidin (S4). Dissociation of the gaseous, multisubunit complexes proceeds predominantly by the loss of a single subunit. Notably, the fractional partitioning of charge between the product ions, i.e., the leaving subunit and the resulting multimer, for a given complex is, within error, constant over the range of charge states investigated. The Arrhenius activation parameters (E(a), A) measured for the loss of subunit decrease with increasing charge state of the complex. However, the parameters for the protonated and deprotonated ions, with the same number of charges, are indistinguishable. The influence of the complex charge state on the dissociation pathways and the magnitude of the dissociation E(a) are modeled theoretically with the discrete charge droplet model (DCDM) and the protein structure model (PSM), wherein the structure of the subunits is considered. Importantly, the major subunit charge states observed experimentally for the Stx1 B5(n+/-) ions correspond to the minimum energy charge distribution predicted by DCDM and PSM assuming a late dissociative transition-state (TS); while for structurally-related Stx2 B5(n+) ions, the experimental charge distribution corresponds to an early TS. It is proposed that the lateness of the TS is related, in part, to the degree of unfolding of the leaving subunit, with Stx1 B being more unfolded than Stx2 B. PSM, incorporating significant subunit unfolding is necessary to account for the product ions observed for the S4(n+) ions. The contribution of Coulombic repulsion to the dissociation E(a) is quantified and the intrinsic activation energy is estimated for the first time.  相似文献   

14.
Previous gas-phase dissociation experiments of protein-protein complexes have resulted in product ion distributions that are asymmetric by charge and mass, providing limited insight into the chemical nature of subunit organization and interaction. In these experiments, a symmetric charge distribution results from an "energy sudden" collision of protein-protein complexes with a surface, indicating that it may be possible to probe the suboligomeric structure of noncovalent complexes in the gas phase. It is proposed that energy sudden surface activation of cytochrome C homodimers results in dissociation without significant unfolding of one of the monomeric subunits. Previously proposed mechanisms for the dissociation of protein-protein complexes are discussed in the context of these results. These experiments demonstrate the potential to preserve the structural details of subunit interaction within a protein-protein complex and help elucidate the asymmetric nature of macromolecular dissociation in the gas phase.  相似文献   

15.
The folding pathways of gas-phase cytochrome c ions produced by electrospray ionization have been studied by an ion trapping/ion mobility technique that allows conformations to be examined over extended timescales (10 ms to 10 s). The results show that the +9 charge state emerges from solution as a compact structure and then rapidly unfolds into several substantially more open structures, a transition that requires 30-60 ms; over substantially longer timescales (250 ms to 10 s) elongated states appear to refold into an array of folded structures. The new folded states are less compact than those that are apparent during the initial unfolding. Apparently, unfolding to highly open conformations is a key step that must occur before +9 ions can sample more compact states that are stable at longer times.  相似文献   

16.
Hemoglobin (Hb) (α2β2) is a tetrameric protein–protein complex. Collision cross sections, hydrogen exchange levels, and tandem mass spectrometry have been used to investigate the properties of gas-phase monomer, dimer, and tetramer ions of adult human hemoglobin (Hb A, α2β2), and two variant hemoglobins: fetal hemoglobin (Hb F, α2γ2) and sickle hemoglobin (Hb S, α2β2, E6V[β]). All three proteins give similar mass spectra. Monomers of Hb S and Hb F have similar cross sections, ca. 10% greater than those of Hb A. Cross sections of dimer ions of Hb S are 11% greater than those of Hb A and 6% greater than those of Hb F. Tetramers of Hb S are 13% larger than tetramers of Hb A or Hb F. Monomers and dimers of all three Hb have similar hydrogen-deuterium exchange (HDX) levels. Tetramers of Hb S exchange 16% more hydrogens than Hb A and Hb F. In tandem mass spectrometry, monomers of Hb S and Hb F require ca. 10% greater internal energy for heme loss than Hb A. Dimers (+11) of Hb A and Hb S dissociate to monomers with asymmetrical charge division; dimers of Hb F (+11) dissociate with nearly equal charge division. Tetramer ions dissociate to monomers and trimers, unlike solution Hb, which dissociates to dimers. The most stable dimers are from Hb S; the most stable tetramers from Hb F. The results with Hb S show that a single mutation in the β chain can change the physical properties of this gas-phase protein–protein complex.  相似文献   

17.
The hydration of doubly protonated gas-phase ions of gramicidin S formed by electrospray ionization was investigated. Under “gentle” electrospray conditions, a near Gaussian distribution of (M + 2H + nH2O)2+ ions with n up to 50 can be readily formed. These extensively hydrated gas-phase ions should have structures similar to those in solution. For intermediate extents of hydration, the “naked” or unsolvated ion is present in unusually high abundance. This is attributed to a competition between solvation of the charges by water vs intramolecular self-solvation via hydrogen bonding. In addition, “magic” numbers of attached water molecules are observed for n = 8, 11, and 14. These magic numbers are attributed to favorable arrangements of water molecules surrounding the charge and surface of the peptide in the gas phase. These results are indicative of a gentle stepwise transformation from the solution-phase structure of the ion to the preferred gas-phase structure as solvent evaporates from the hydrated ions.  相似文献   

18.
Ion mobility (IM) and tandem mass spectrometry (MS/MS) coupled with native MS are useful for studying noncovalent protein complexes. Collision induced dissociation (CID) is the most common MS/MS dissociation method. However, some protein complexes, including glycogen phosphorylase B kinase (PHB) and L-glutamate dehydrogenase (GDH) examined in this study, are resistant to dissociation by CID at the maximum collision energy available in the instrument. Surface induced dissociation (SID) was applied to dissociate the two refractory protein complexes. Different charge state precursor ions of the two complexes were examined by CID and SID. The PHB dimer was successfully dissociated to monomers and the GDH hexamer formed trimeric subcomplexes that are informative of its quaternary structure. The unfolding of the precursor and the percentages of the distinct products suggest that the dissociation pathways vary for different charge states. The precursors at lower charge states (+21 for PHB dimer and +27 for GDH hexamer) produce a higher percentage of folded fragments and dissociate more symmetrically than the precusors at higher charge states (+29 for PHB dimer and +39 for GDH hexamer). The precursors at lower charge state may be more native-like than the higher charge state because a higher percentage of folded fragments and a lower percentage of highly charged unfolded fragments are detected. The combination of SID and charge reduction is shown to be a powerful tool for quaternary structure analysis of refractory noncovalent protein complexes, as illustrated by the data for PHB dimer and GDH hexamer.
Figure
?  相似文献   

19.
Multiply charged poly(ethylene glycol) ions of the form (M+nNa) n+ derived from electrospray ionization have been subjected to reactions with negative ions in the quadrupole ion trap. Mixtures of multiply charged positive ions ranging in average mass from about 2000 to about 14,000 Da were observed to react with perfluorocarbon anions by either proton transfer or fluoride transfer. Iodide anions reacted with the same positive ions by attachment. In no case was fragmentation of the polymer ion observed. In all cases, the multiply charged positive ion charge states could be readily reduced to +1, thereby eliminating the charge state overlap observed in the normal electrospray mass spectrum. With all three reaction mechanisms, however, the +1 product ions were comprised of mixtures of products with varying numbers of sodium ions, and in the case of iodide attachment and fluoride transfer, varying numbers of halogen anions. These reactions shift the mass distributions to higher masses and broaden the distributions. The extents to which these effects occur are functions of the magnitudes of the initial charges and the width of the initial charge state distributions. Care must be taken in deriving information about the polymer molecular weight distribution from the singly charged product ions arising from these ion/ion reactions. The cluster ions containing iodide were shown to be intermediates in sodium ion transfer. Dissociation of the adduct ions can therefore lead to a +1 product ion population that is comprised predominantly of M+Na+ ions. However, a strategy based on the dissociation of the iodide cluster ions is limited by difficulties in dissociating high mass-to-charge ions in the quadrupole ion trap.  相似文献   

20.
Noncovalent peptide-peptide and peptide-water interactions in small model systems were examined using an electrospray mass spectrometer equipped with a high-pressure drift cell. The results of these aggregation and hydration experiments were interpreted with the aid of molecular mechanics (MM) and density functional theory (DFT) calculations. The systems investigated include bare deprotonated monomers and dimers [P(1,2)-H](-) and hydrated deprotonated monomers and dimers [P(1,2)-H](-).(H(2)O)(n)() for the peptides dialanine (P = AA) and diglycine (P = GG). Mass spectra indicated that both peptides AA and GG form exclusively dimer ions in the electrospray process. Monomeric ions were generated by high-energy injection of the dimers into the drift cell. Temperature-dependent hydration equilibrium experiments carried out in the drift cell yielded water binding energies ranging from 11.7 (first water molecule) to 7.1 kcal/mol (fourth water) for [AA-H](-) and 11.0 to 7.4 kcal/mol for [GG-H](-). The first water molecule adding to the dimer ions [AA-H](-).(AA) and [GG-H](-).(GG) is bound by 8.4 and 7.5 kcal/mol, respectively. The hydration mass spectra for the monomers and dimers provide a means to compare the ability of water and a neutral peptide to solvate a deprotonated peptide [P-H](-). The data indicate that a similar degree of solvation is achieved by four water molecules, [P-H](-).(H(2)O)(4), or one neutral peptide, [P-H](-).(P). Temperature-dependent kinetics experiments yielded activation energies for dissociation of the dimers [AA-H](-).(AA) and [GG-H](-).(GG) of 34.9 and 32.2 kcal/mol, respectively. MM and DFT calculations carried out for the dialanine system indicated that the dimer binding energy is 24.3 kcal/mol, when the [AA-H](-) and AA products are relaxed to their global minimum structures. However, a value of 38.9 kcal/mol is obtained if [AA-H](-) and AA dissociate but retain the structures of the moieties in the dimer, suggesting the transition state occurs early in the dissociation process. Similar results were found for the diglycine dimer.  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