ArticlePDF Available

Study of the gas-phase fragmentation behaviour of sulfonated peptides

Authors:

Abstract and Figures

A series of singly and doubly protonated peptides bearing sulfonated residue have been studied, using both experiment and molecular modelling, to elucidate fragmentation chemistry of sulfonated peptides. Collision-induced dissociation mass spectra indicate that the sulfo group loss (neutral loss of 80 Da) is the dominant dissociation channel. Modelling results suggest the proton transfer mechanism, where upon vibrational excitation, the acidic side chain proton is transferred from the sulfo group hydroxyl to the ester oxygen resulting in S-O bond cleavage and formation of the unmodified hydroxyl containing residue and SO3. Conformations associated with potential energy profile of the reaction imply the charge remote nature of the proposed mechanism. The proposed proton transfer mechanism was compared with the intramolecular nucleophilic substitution (SN2) mechanism, the main pathway suggested for neutral loss of phosphoric acid from phosphopeptides. Both pathways (proton transfer and SN2) are available for sulfonated and phosphorylated peptides; however, each posttranslational modification favours different mechanism. The change of the bond dissociation enthalpies and the ability of stabilising the transition state structures are demonstrated as main factors responsible for each posttranslational modification activating a different pathway.
Content may be subject to copyright.
Study of the gas-phase fragmentation behaviour of sulfonated peptides
Sanja Škulj, Marko Rožman
a b s t r a c t
A series of singly and doubly protonated peptides bearing sulfonated residue have been studied, using both experiment and molecular modelling, to
elucidate fragmentation chemistry of sulfonated peptides. Collision-induced dissociation mass spectra indicate that the sulfo group loss (neutral loss of
80 Da) is the dominant dissociation channel. Modelling results suggest the proton transfer mechanism, where upon vibrational excitation, the acidic side
chain proton is transferred from the sulfo group hydroxyl to the ester oxygen resulting in S O bond cleavage and formation of the unmodified hydroxyl
containing residue and SO3. Conformations associated with potential energy profile of the reaction imply the charge remote nature of the proposed
mechanism. The proposed proton transfer mechanism was compared with the intramolecular nucleophilic substitution (SN2) mechanism, the main
pathway suggested for neutral loss of phosphoric acid from phosphopeptides. Both pathways (proton transfer and SN2) are available for sulfonated and
phosphorylated peptides; however, each posttranslational modification favours different mechanism. The change of the bond dissociation enthalpies
and the ability of stabilising the transition state structures are demonstrated as main factors responsible for each posttranslational modification
activating a different pathway.
1. Introduction
One of the most common uses of mass spectrometry (MS) is
obtaining the structure or sequence of an ion being analysed. The
gas-phase approach to generate structure-specific information
involves use of tandem MS and subsequent interpretation of frag-
ment ion spectra. In the tandem MS of the protonated peptides,
the ion of interest is isolated and then (usually) dissociated via low
energy vibration excitation (either collision-induced dissociation
(CID) or infrared multi-photon dissociation (IRMPD)). Assignment
of molecular structures to the tandem MS spectra greatly relies on
the fragmentation models used. The most comprehensive set of
rules for understanding of dissociation mechanisms of protonated
peptides is known as the mobile proton model [1–5]. The model
(introduced by Vicki Wysocki and Simon Gaskell) assumes that in
the activated protonated peptide, ionising proton(s) can migrate
to various sites, thus triggering charge-directed fragmentation
mechanisms [1–5]. The mobile proton concept has been success-
fully applied in numerous studies of fragmentation mechanisms
of various tryptic and non-tryptic peptides, cyclic peptides and
peptides bearing some posttranslational modification (PTM) [1–7].
Sulfonation, common PTM in multicellular eukaryotes, repre-
sents addition of sulfonic acid group to a protein. Modification
is detected on Tyr (mainly), Ser and Thr residues and has the
same nominal mass increase as phosphorylation (+80 Da) [8–10].
Thus, it was suggested that rates of protein sulfonation could be
underestimated due to a coexistence of both phosphorylated and
sulfonated (isobaric) forms of the same peptide [8–10]. Use of
ultra-high resolution (accurate mass) measurements, ultraviolet
and infrared photodissociation spectroscopy techniques demon-
strated that PTM modified isobaric peptides could be distinguished
[11–13]. Despite the advantages of high-resolution mass measure-
ments, photodissociation techniques and electron capture/transfer
dissociation [14], CID remains the generally used approach for
sulfonation assignments [8–10,15–17]. Protonated sulfonated pep-
tides analysed by CID undergo the facile neutral loss of the sulfur
trioxide (SO3, loss of 80 Da) from their precursor ions [8–10,15–17].
The neutral loss precedes any peptide backbone fragmentation,
thus limiting precise localization of the sulfonation sites. Under-
standing of the gas-phase dissociation pathways of sulfonated
peptides is very limited. Recently, Patrick et al. preformed structural
investigation of protonated sulfoserine dissociation pathways [18].
Low-energy CID of protonated sulfoserine produced two major
product ions: 3-member aziridine ring structure attributed to the
neutral loss of 98 Da (loss of H2SO4) and a structure identical to
the protonated serine was related to the loss of 80 Da (loss of SO3).
However, insights at the peptide level are still missing.
Better understanding of the fragmentation chemistry of sul-
fonated peptides would be of value. Previously, we provided
a description of the gas-phase dissociation of phosphorylated
peptides [7]. One may consider sulfonation and phosphoryla-
tion as similar because both modifications represent highly acidic
monoesters of their respective acids and give rise to a nominal mass
increase of 80 Da However, neutral loss during vibrational exci-
tation from protonated sulfopeptides is associated to SO3[8–10]
while that from phosphopeptides (phosphorylated Ser and Thr
residues) is associated mainly to elimination of H3PO4[6,7], i.e.
suggesting activation of different dissociation pathways.
In this work, through the combination of low-energy CID exper-
iments and molecular modelling, we present the description of the
fragmentation mechanisms of sulfonated peptides. Furthermore,
we complement the present findings on sulfopeptides with our pre-
vious results on phosphopeptides and attempt to understand why
a certain fragmentation pathway is related to specific PTM.
2. Materials and methods
2.1. Materials
Analytes and reagents were obtained from Sigma-Aldrich (St
Louis, USA) and used without further purification. Peptides TSQLL,
SAALSLLR, SAALYLLR and their posttranslational modified variants
were obtained from PolyPeptide Laboratories (Strasbourg, France).
Peptide sulfonation was achieved using the procedure described
in ref. [9]. Briefly, peptides were dissolved in trifluoroacetic acid
and reacted with 5% chlorosulfonic acid (ClSO3H) at room temper-
ature for 20 min The reaction was terminated adding H2O and the
solution neutralised with NH4OH.
2.2. Mass spectrometry
MS and tandem MS experiments were carried out on the amaZon
ion trap mass spectrometer (Bruker Daltonik, Bremen, Germany).
Peptides were dissolved in 50/50 ethanol/water with 0.1% formic
acid to obtain 1 M concentration. Solution was introduced into the
electrospray ionisation source by direct infusion at the flow rate of
75 L/h. The capillary voltage was set at 4500 V while high voltage
end plate offset was 500 V. The temperature and the flow rate
of the drying gas were set at 205 C and 5 L/min, respectively. The
electrospray ionisation source parameters were optimised to allow
an efficient ionisation and to reduce the in-source fragmentation
of precursor ions. The isolation width of the precursor ion was set
at 2 Da The CID excitation time was 40 ms and the amplitude was
in the 0.4–1 V range, depending on a precursor. All spectra were
acquired in the positive ion mode using a scan range from m/z 100
to 1100. DataAnalysis 4.0 and BioTools 3.2 (Bruker Daltonik GmbH,
Bremen, Germany) were used for spectra analysis and extraction of
the MS and tandem MS data.
2.3. Computational methods
In order to gain initial understanding of the potential energy
surface (PES) associated with loss of the sulfate modification, the
proposed pathways were first established on the small model sys-
tem, CH3COsSerNHCH3, and then further evaluated on the test
peptides TsSQLL and SAALsSLLR. Both the small model system
and the peptides were optimised at the B3LYP/6-31G(d) level of
theory. Both the functional and the basis set represent a good
compromise for obtaining satisfactory geometries and approxi-
mate relative energies, as demonstrated in the theoretical studies
of similar systems [7,19,20]. Stationary points (i.e. the minima and
transition states on the potential energy surface) were identified
by the harmonic frequency analysis. Transition state structures
were additionally tested by the Intrinsic Reaction Coordinate (IRC)
analysis. In order to get a more accurate description of dissocia-
tion energies, calculations using the G3(MP2)//B3LYP protocol [21]
were performed on a restricted number of the small model sys-
tem conformations. The G3(MP2)//B3LYP results were correlated
with series of the single point energy calculations in order to pin-
point a model suitable for use on the test peptides. The B2PLYP,
B3LYP, M062X and MP2 methods were used in combination with
the different basis sets: 6-31G(d), 6-31+G(d), 6-31++G(d,p), 6-
311++G(d,p) and TZVP. The smallest mean absolute deviation
was found for energies calculated at the B3LYP/TZVP level of
theory.
Combination of quenched dynamics and simulated annealing
with the AMBER 99 force field was used to sample the poten-
tial energy surface of the test peptides (TSQLL and SAALSLLR) by
the protocol identical to that previously used [7,22]. Final struc-
tures were reoptimised using the B3LYP/6-31G(d) level of theory
and the lowest energy structure was considered as the represen-
tative structure. From the representative structure, the potential
energy profiles of dissociation pathways were constructed at the
B3LYP/TZVP//B3LYP/6-31G(d) level. The Rice-Ramsperger-Kassel-
Marcus (RRKM) kinetic theory was used to describe the reaction
rate as a function of internal energy of peptides.
All quantum mechanic calculations were established using the
Gaussian 09 [23], molecular dynamic simulations were carried out
using the AMBER 12 [24] and RRKM calculations were obtained
using the MassKinetics 1.15 [25].
3. Results and discussion
3.1. CID of sulfonated peptides – mobile proton environment
The low-energy CID product ion spectra were examined in
order to set initial understanding of the gas-phase fragmentation
behaviour of sulfonated peptides. The spectra obtained by dissoci-
ation of the doubly protonated SAALsSLLR and SAALsYLLR exhibit a
very intense neutral loss of 80 Da, Fig. 1. The sulfo group loss from
the precursor ion as well as from b and y product ions dominates
all tandem MS spectra analysed in this work (Fig. 1 and Fig. S1 –
supplementary data). However, in some cases, small portion of b
and y ions retain the sulfo group and enable characterisation of the
peptides (Fig. 1). Those ions are associated with 5.7% of the total
ion intensity of all identified ions (in this work) and are not readily
observed in tandem MS spectra of sulfonated peptides (e.g. Fig. S1
b and ref. [8–10,16]). Therefore, it is sometimes difficult to obtain
site-specific information on the location of the sulfo group.
Although the tandem MS spectra of peptides described here
represent only an example, together with previous results [9,16]
they indicate that the sulfo group loss is the dominant dissociation
channel in the collisionally activated peptides where ionising pro-
ton can migrate to various sites (the mobile proton environment).
It would be of interest to collect more sulfopeptide tandem MS
spectra obtained on different instruments in order to statistically
characterise fragmentation behaviour but this is out of the scope of
the present paper.
The mechanism, which could be associated with the sulfo group
loss, includes the proton transfer from the sulfo group hydroxyl
to the ester oxygen and consequent formation of the hydroxyl
group and SO3(Scheme 1 a). The mechanism is similar to the one
associated with the loss of metaphosphoric acid in phosphorylated
Fig. 1. CID tandem MS product ion spectra of the doubly protonated (a) SAALsSLLR
and (b) SAALsYLLR. *80 Da (–SO3).
peptides [7] and was hypothesised for the sulfo group loss from
sulfoserine [18].
The potential energy profile for the sulfo group loss from
SAALsSLLR peptide is shown in Fig. 2, together with the opti-
mised structures of the energy minima and transition state. As
hypothesised, the mechanism includes transfer of the acidic side
Scheme 1. . (a) Proton transfer mechanism and (b) intramolecular nucleophilic
substitution (SN2) mechanism.
Fig. 2. Schematic representation of the potential energy profile of the sulfo group
loss from the doubly protonated SAALsSLLR peptide. The relative energies are calcu-
lated at the B3LYP/TZVP//B3LYP/6-31G(d) level of theory with respect to the most
stable SAALsSLLR conformer found. Magnified structures with interatomic distances
are available as supplementary data (Fig. S7).
chain proton from the sulfo group hydroxyl to the ester oxygen.
In the transition state, the sulfo group is additionally stabilised via
interaction with protonated amino terminus. Although there is a
possibility for activating the ionising proton from the amino termi-
nus during reaction, there is no evidence that the ionising proton is
involved in the mechanism. The protonated amino terminus–sulfo
group interaction could suggest a charge-directed pathway where
the charged amino terminus could make the sulfo hydroxyl group
a more electron deficient and thus a more efficient proton donor.
The role of the ionising proton in the proton transfer mechanism
was further examined with singly protonated TsSQLL peptide (the
mobile proton environment). In TsSQLL peptide, the protonated
amino terminus and the sulfonated serine side chain are very
close; therefore, involvement of ionising proton should be evident.
However, the identical reaction mechanism in protonated TsSQLL
peptide does not require interaction of the sulfo group with the pro-
tonated amino terminus, suggesting the charge remote nature of
the sulfo group loss pathway (Fig. S2 – supplementary data). Thus,
observed interaction in SAALsSLLR peptide can be attributed to the
secondary structure stabilisation via hydrogen bonds. Upon SO3
abstraction from the sulfonated side chain, the side chain residue
takes the form structurally analogous to the unmodified hydroxyl-
containing residue, consistent with the IRMPD observations on
sulfoserine [18]. At this point, SO3molecule probably interacts with
peptide and consequently the ion–molecule complex dissociates
making the neutral loss of SO3a two-step process (abstrac-
tion and ion–molecule complex dissociation), in accordance
with observations of phosphopeptides dissociation dynamics
[26].
Reaction barrier for sulfo group loss of the doubly protonated
SAALsSLLR peptide is 174.8 kJ mol1. Calculated unimolecular reac-
tion rates (by the RRKM theory) show that reaching ms time range
pathway requires 1180 kJ mol1while increase up to 1880 kJ mol1
is needed for s time range (Fig. S3 – supplementary data). Rate
energy dependences of SO3loss pathway match literature available
average internal energy values of similar size peptides (approxi-
mately 900 for ms time range and 1700 kJ mol1s time range)
[27].
3.2. CID of sulfonated peptides – limited proton mobility
The precursor ion and backbone fragments in the low-energy
CID spectra of peptides under the mobile proton environment
exhibit a very intense neutral loss of 80 Da Loss of the sulfo moiety
is even more prominent under the limited proton mobility, Fig. 3.
Fig. 3. CID tandem MS product ion spectra of the singly protonated (a) SAALsSLLR
and (b) SAALsYLLR. *80 Da (–SO3).
Singly protonated SAALsSLLR and SAALsYLLR peptides preferen-
tially eliminate the modification even before backbone fragmenta-
tion occurs. Deposition of the larger amounts of additional vibronic
energy sometimes may result in additional peptide backbone frag-
mentation, in accordance with already documented findings [16].
Assessment of the potential energy surface of the sulfo group
loss for the singly protonated SAALsSLLR reveals identical charge
remote mechanism as the one under the mobile proton environ-
ment (Fig. 1a). However, there is no interaction between the sulfo
moiety and the arginine guanidino group.
According to RRKM calculations, the sulfo group loss from the
precursor ion occurs at the millisecond time scale at the internal
energy of 1010 kJ mol1(Fig S4 – supplementary data). Internal
energy window for the peptide bond cleavage is from 1000 to
1700 kJ mol1[27,28], suggesting elimination of the sulfo moiety
from the precursor ion as the principal dissociation channel, which
is in agreement with our experimental observations.
3.3. Neutral loss of PTM: sulfonated vs. phosphorylated peptides
The major fragmentation pathway occurring from energised
protonated sulfopeptide originates by loss of SO3(80 Da), Fig. 1. A
phospho modification containing peptides (except pTyr residue),
in contrast, shows pronounced peaks related to elimination of
H3PO4(98 Da) (tandem MS spectra of TpSQLL and SAALpSLLR
in Fig. S5 – supplementary data). Neutral loss of the phosphoric
acid involves the charge-directed intramolecular nucleophilic sub-
stitution (SN2) mechanism (Scheme 1b) while sulfur trioxide loss
Table 1
The proton transfer and SN2 mechanism reaction barriers (in kJ mol1) calculated at
the B3LYP/TZVP//B3LYP/6-31G(d) level of theory. The reaction barriers were calcu-
lated with respect to the most stable peptide conformation. Total electronic energies
(in Eh) and atomic coordinates are available in supporting data file (Table S1).
Peptide Charge state Reaction mechanism
Proton transfer SN2
TsSQLL 1 148.9 157.1
TpSQLL 1 165.7 134.6
SAALsSLLR 1 155.1 167.9
SAALpSLLR 1 173.5 166.9
SAALsSLLR 2 174.8 181.2
SAALpSLLR 2 200.4 180.1
Model system (sS)a1 111.1 89.3
Model system (pS)a1 160.6 93
aReaction barrier of the model system was calculated using G3(MP2)//B3LYP
protocol.
is related to charge remote proton transfer mechanism (Fig. 2a). In
both modifications, the presence of other modification dominant
mechanisms can be observed to a smaller extent, i.e. SN2 mecha-
nism in sulfonated peptides (loss of sulfuric acid, 98 Da e.g. Fig.
S1 – supplementary data) and proton transfer mechanism in phos-
phorylated peptides (loss of metaphosphoric acid, 98 Da, e.g. ref.
[6]). Both modifications represent similar highly acidic monoesters
and have potential for proton transfer and intramolecular nucleo-
philic substitution; however during collisional activation different
dissociation channels were dominant.
To compare both mechanisms, potential energy profiles for
phosphorylated/sulfonated SAALSLLR and TSQLL peptides were
constructed and the corresponding reaction barriers at the
B3LYP/TZVP//B3LYP/6-31G(d) level of theory are given in Table 1.
From reactions barriers, it follows that moving from sulfonated
to phosphorylated peptide increases the proton transfer barrier
height while SN2 threshold remains roughly the same or slightly
lower in phosphorylated peptides. An explanation is offered by
inspecting the nature of functional groups.
In the intramolecular nucleophilic substitution reaction, the
alkyl group–ester oxygen bond is cleaved while in the proton trans-
fer reaction, the ester oxygen–sulfur/phosphor bond is cleaved.
SN2 reaction breaks the same bond regardless of PTM while pro-
ton transfer mechanism breaks a different bond. Regarding the
bond strength, it is useful to consider mean bond dissociation
enthalpy since it may provide rough estimate of relative bond
strength. S O bond dissociation enthalpy is 37 kJ mol1lower
than P O bond dissociation enthalpy, suggesting that S O bond
is weaker than P O[29]. SN2 barrier lowering in peptides bear-
ing phospho modification may be attributed to extra hydroxylic
group available at phospho residue which enables additional
hydrogen bond formation and stabilisation of the transition state
structure. Intramolecular nucleophilic substitution requires side
chain–polypeptide chain interaction while proton transfer mech-
anism is restricted only to the side chain; thus, SN2 mechanism
requires “more defined” transition state conformation and there-
fore it is more susceptible to stabilisation by additional hydrogen
bond. The small model system (see Section 2) can provide indica-
tive information about the barriers taking into account the bond
strength and (on the other hand) it is free of secondary structure
stabilisation effects. SN2 barrier estimated at the sulfo and phospho
model systems is roughly the same while the proton transfer bar-
rier for the phospho model system is 50 kJ mol1which is higher
than for sulfo (Table 1).
In both mechanisms, the geometry of transition state is rel-
atively close to the product geometry and consequently RRKM
curves will have similar shape (Fig. 4 and Fig. S6). Since the size
of the precursor does not change, the main difference between
Fig. 4. RRKM theory unimolecular reaction rate constants for the proton transfer
and SN2 pathway in doubly protonated SAALsSLLR and SAALpSLLR peptide.
RRKM curves will be determined by activation energy (reaction
threshold). Accordingly, change in modification from sulfonation to
phosphorylation will raise proton transfer barrier (and lower SN2
barrier) making nucleophilic substitution mechanism predominant
(Fig. 4 and Fig S6). RRKM plots show dominance of proton transfer
mechanism for sulfonated peptides and dominance of SN2 path-
way in phosphorylated peptides, which is in agreement with our
suggestions and experimental data (Fig. 4 and Fig S6).
4. Conclusions
Combined experimental and theoretical investigation of the gas-
phase dissociation behaviour of sulfonated peptides yielded the
following results.
Regardless of proton mobility environment, neutral loss of 80 Da
from sulfonated peptides is related to the charge remote pro-
ton transfer mechanism. Molecular modelling predicts that during
vibrational excitation, the acidic side chain proton is transferred
from sulfo group hydroxyl to the ester oxygen. Protonation of
the ester oxygen leads to S O bond cleavage and to formation of
unmodified hydroxyl containing residue and SO3.
In comparison with other dissociation channels (e.g. b/y ion for-
mation), the neutral loss product ion(s) should be dominant in the
tandem MS spectrum. Furthermore, with limiting proton mobility,
their abundance should increase even more.
Although identical dissociation pathways (the proton transfer
and the intramolecular nucleophilic substitution) are available for
sulfonated and phosphorylated peptides, each PTM activates dif-
ferent mechanism. Preference of the proton transfer mechanism
in sulfonated peptides is a consequence of the fact that the S O
bond is weaker compared to P O bond. On the other hand, due to
the stronger P O bond and the possibility of additional stabilisa-
tion of the transition state (additional OH group), phosphorylated
peptides follow the intramolecular nucleophilic substitution
mechanism.
Acknowledgements
This manuscript is dedicated to Prof. Simon J. Gaskell in happy
celebration of his 65th birthday. The Ministry of Science, Education
and Sports of Republic of Croatia supported this work.
Appendix A. Supplementary data
Supplementary data associated with this article can be found, in
the online version, at http://dx.doi.org/10.1016/j.ijms.2015.07.023
References
[1] O. Burlet, C.Y. Yang, S.J. Gaskell, Influence of cysteine to cysteic acid oxidation
on the collision-activated decomposition of protonated peptides: evidence for
intraionic interactions, J. Am. Soc. Mass Spectrom. 3 (1992) 337–344.
[2] A.L. McCormack, Á. Somogyi, A.R. Dongré, V.H. Wysocki, Fragmentation of pro-
tonated peptides: surface-induced dissociation in conjunction with a quantum
mechanical approach, Anal. Chem. 65 (1993) 2859–2872.
[3] K.A. Cox, S.J. Gaskell, M. Morris, J. Whiting, Role of the site of protonation in
the low-energy decompositions of gas-phase peptide ions, J. Am. Soc. Mass
Spectrom. 7 (1996) 522–531.
[4] A.R. Dongre´
ı, J.L. Jones, A. Somogyi, V.H. Wysocki, Influence of peptide com-
position, gas-phase basicity, and chemical modification on fragmentation
efficiency: evidence for the mobile proton model, J. Am. Chem. Soc. 118 (1996)
8365–8374.
[5] R. Boyd, A. Somogyi, The mobile proton hypothesis in fragmentation of
protonated peptides: a perspective, J. Am. Soc. Mass. Spectrom. 21 (2010)
1275–1278.
[6] A.M. Palumbo, G.E. Reid, Evaluation of gas-phase rearrangement and com-
peting fragmentation reactions on protein phosphorylation site assignment
using collision induced dissociation-MS/MS and MS3, Anal. Chem. 80 (2008)
9735–9747.
[7] M. Roˇ
zman, Modelling of the gas-phase phosphate group loss and
rearrangement in phosphorylated peptides, J. Mass. Spectrom. 46 (2011)
949–955.
[8] C. Seibert, T.P. Sakmar, Toward a framework for sulfoproteomics: synthesis
and characterization of sulfotyrosine-containing peptides, Pept. Sci. 90 (2007)
459–477.
[9] K.F. Medzihradszky, Z. Darula, E. Perlson, M. Fainzilber, R.J. Chalkley, H.
Ball, D. Greenbaum, M. Bogyo, D.R. Tyson, R.A. Bradshaw, A.L. Burlingame,
O-sulfonation of serine and threonine: mass spectrometric detection and
characterization of a new posttranslational modification in diverse proteins
throughout the eukaryotes, Mol. Cell. Proteomics 3 (2004) 429–440.
[10] K.A. Dave, F. Whelan, C. Bindloss, S.G.B. Furness, A. Chapman-Smith, M.L.
Whitelaw, J.J. Gorman, Sulfonation and phosphorylation of regions of the
dioxin receptor susceptible to methionine modifications, Mol. Cell. Proteomics
8 (2009) 706–719.
[11] R.E. Bossio, A.G. Marshall, Baseline resolution of isobaric phosphorylated and
sulfated peptides and nucleotides by electrospray ionization FTICR MS: another
step toward mass spectrometry-based proteomics, Anal. Chem. 74 (2002)
1674–1679.
[12] A.L. Patrick, C.N. Stedwell, N.C. Polfer, Differentiating sulfopeptide and phos-
phopeptide ions via resonant infrared photodissociation, Anal. Chem. 86 (2014)
5547–5552.
[13] M.R. Robinson, K.L. Moore, J.S. Brodbelt, Direct identification of tyrosine sulfa-
tion by using ultraviolet photodissociation mass spectrometry, J. Am. Soc. Mass
Spectrom. 25 (2014) 1461–1471.
[14] K.F. Medzihradszky, S. Guan, D.A. Maltby, A.L. Burlingame, Sulfopeptide frag-
mentation in electron-capture and electron-transfer dissociation, J. Am. Soc.
Mass Spectrom. 18 (2007) 1617–1624.
[15] T. Yagami, K. Kitagawa, S. Futaki, Liquid secondary-ion mass spectrometry of
peptides containing multiple tytosine-O-sulfates, Rapid Commun. Mass Spec-
trom. 9 (1995) 1335–1341.
[16] J.F. Nemeth-Cawley, S. Karnik, J.C. Rouse, Analysis of sulfated peptides using
positive electrospray ionization tandem mass spectrometry, J. Mass Spectrom.
36 (2001) 1301–1311.
[17] J.L. Wolfender, F. Chu, H. Ball, F. Wolfender, M. Fainzilber, M.A. Baldwin, A.L.
Burlingame, Identification of tyrosine sulfation in Conus pennaceus conotoxins
-PnIA and -PnIB: further investigation of labile sulfo- and phosphopep-
tides by electrospray, matrix-assisted laser desorption/ionization (MALDI) and
atmospheric pressure MALDI mass spectrometry, J. Mass Spectrom. 34 (1999)
447–454.
[18] A.L. Patrick, C.N. Stedwell, B. Schindler, I. Compagnon, G. Berden, J. Oomens,
N.C. Polfer, Insights into the fragmentation pathways of gas-phase protonated
sulfoserine, Int. J. Mass Spectrom. 379 (2015) 26–32.
[19] M. Roˇ
zman, Aspartic acid side chain effect – experimental and theoretical
insight, J. Am. Soc. Mass Spectrom. 18 (2007) 121–127.
[20] K.A. Herrmann, V.H. Wysocki, E.R. Vorpagel, Computational investigation
and hydrogen/deuterium exchange of the fixed charge derivative tris(2,4,6-
trimethoxyphenyl) phosphonium: implications for the aspartic acid cleavage
mechanism, J. Am. Soc. Mass Spectrom. 16 (2005) 1067–1080.
[21] A.G. Baboul, L.A. Curtiss, P.C. Redfern, K. Raghavachari, Gaussian-3 theory using
density functional geometries and zero-point energies, J. Chem. Phys. 110
(1999) 7650–7657.
[22] M. Roˇ
zman, S.J. Gaskell, Non-covalent interactions of alkali metal cations with
singly charged tryptic peptides, J. Mass. Spectrom. 45 (2010) 1409–1415.
[23] M.J. Frisch, G.W. Trucks, H.B. Schlegel, G.E. Scuseria, M.A. Robb, J.R. Cheeseman,
J.A. Montgomery Jr., T. Vreven, K.N. Kudin, J.C. Burant, J.M. Millam, S.S. Iyengar,
J. Tomasi, V. Barone, B. Mennucci, M. Cossi, G. Scalmani, N. Rega, G.A. Petersson,
H. Nakatsuji, M. Hada, M. Ehara, K. Toyota, R. Fukuda, J. Hasegawa, M. Ishida, T.
Nakajima, Y. Honda, O. Kitao, H. Nakai, M. Klene, X. Li, J.E. Knox, H.P. Hratchian,
J.B. Cross, C. Adamo, J. Jaramillo, R. Gomperts, R.E. Stratmann, O. Yazyev, A.J.
Austin, R. Cammi, C. Pomelli, J.W. Ochterski, P.Y. Ayala, K. Morokuma, G.A.
Voth, P. Salvador, J.J. Dannenberg, V.G. Zakrzewski, S. Dapprich, A.D. Daniels,
M.C. Strain, O. Farkas, D.K. Malick, A.D. Rabuck, K. Raghavachari, J.B. Foresman,
J.V. Ortiz, Q. Cui, A.G. Baboul, S. Clifford, J. Cioslowski, B.B. Stefanov, G. Liu, A.
Liashenko, P. Piskorz, I. Komaromi, R.L. Martin, D.J. Fox, T. Keith, M.A. Al-Laham,
C.Y. Peng, A. Nanayakkara, M. Challacombe, P.M.W. Gill, B. Johnson, W. Chen,
M.W. Wong, C. Gonzalez, J.A. Pople, Gaussian 09, Gaussian, Inc., Pittsburgh, PA,
2009.
[24] D.A. Case, T.A. Darden, T.E. Cheatham III, C.L. Simmerling, J. Wang, R.E. Duke,
R. Luo, R.C. Walker, W. Zhang, K.M. Merz, B. Roberts, S. Hayik, A. Roitberg, G.
Seabra, J. Swails, A.W. Goetz, I. Kolossvai, K.F. Wong, F. Paesani, J. Vanicek, R.M.
Wolf, J. Liu, X. Wu, S.R. Brozell, T. Steinbrecher, H. Gohlke, Q. Cai, X. Ye, J. Wang,
M.-J. Hsieh, G. Cui, D.R. Roe, D.H. Mathews, M.G. Seetin, R. Salomon-Ferrer, C.
Sagui, V. Babin, T. Luchko, S. Gusarov, A. Kovalenko, P.A. Kollman, AMBER 12,
University of California, San Francisco, 2012.
[25] L. Drahos, K. Vékey, MassKinetics: a theoretical model of mass spectra incor-
porating physical processes, reaction kinetics and mathematical descriptions,
J. Mass Spectrom. 36 (2001) 237–263.
[26] J. Laskin, R.P.W. Kong, T. Song, I.K. Chu, Effect of the basic residue on the ener-
getics and dynamics of dissociation of phosphopeptides, Int. J. Mass Spectrom.
295 (2012) 330–332.
[27] K. Vékey, A. Somogyi, V.H. Wysocki, Average activation energies of low-
energy fragmentation processes of protonated peptides determined by a new
approach, Rapid Commun. Mass Spectrom. 10 (1996) 911–918.
[28] J. Laskin, T.H. Bailey, J.H. Futrell, Fragmentation energetics for angiotensin II
and its analogs from time- and energy-resolved surface-induced dissociation
studies, Int. J. Mass Spectrom. 234 (2004) 89–99.
[29] W. Reusch, Virtual Textbook of Organic Chemistry, Department of Chemistry,
Michigan State University, http:// www2.chemistry.msu.edu/faculty/reusch/
VirtTxtJml/react2.htm#rx6
... Fifty-seven peptides were identified as phosphorylated, and 4 were identified as tyrosine-sulfated, respectively belonging to 46 and 3 protein groups (Table S2). MaxQuant searches tyrosine sulfopeptides by the 79.9568 mass shift in the precursor and the typical SO 3 neutral loss, as described in the literature for CID fragmentation of sulfopeptides in the positive-ionization mode, 8 which is dominated by the neutral loss of SO 3 (79.956817) due to the proton-transfer mechanism being preferred over the S N 2 dissociation mechanism, which is, on the contrary, the main dissociation path for phosphopeptides and produces the H 3 PO 4 (97.976898) ...
... As such, if peptides with a given sequence modified with either sulfation or phosphorylation are nearly isobaric in the MS scan (difference between the −H 2 PO 3 and −HSO 3 moieties is only 9.6 mDa 1 ), the fragmentation mechanism in the MS/MS mode is different and can distinguish phosphorylation from sulfation, although site location cannot be afforded. 8 For comparison, Proteome Discover was also employed to identify peptides in the enriched serum. Results provided a total 2257 peptides identifications, with 35 and 6 phospho-and sulfopeptides, respectively. ...
Article
Protein tyrosine O-sulfation is an important post-translational modification, as it has been correlated to inflammation, virus infection and signal pathways. Nevertheless, methods for the characterization of protein sulfation by sulfopeptide enrichment are currently limited. In this paper, two standard compounds, representative of mono- and di-sulfated peptides, were used to compare the enrichment capability of five sorbent materials: two commercial weak anion exchange mixed mode sorbents (Strata X-AW and Oasis WAX) and three phosphopeptide enrichment materials based on affinity chromatography to either immobilized metals (IMAC) or to metal oxides, i.e., Fe3+, TiO2 or Ti4+. The sulfopeptides were analyzed by UHPLC-multiple reaction monitoring analysis and were stable under all the tested experimental conditions. Recoveries of the enrichment step from spiked bovine serum albumin digests were above 80% for the commercial Fe-IMAC kit and the Strata X-AW sorbent. Shotgun proteomics was used on the same samples to evaluate the selectivity, calculated as the number of co-enriched peptides, and it was found better for the Fe-IMAC kit. Therefore, the Fe-IMAC protocol was embedded in a shotgun proteomics workflow and applied to serum spiked with the sulfopeptides before protein dephosphorylation and digestion. The recovery of the entire analytical workflow was 20%, which was compatible with previous data on TiO2 phosphopeptide enrichment. Despite the potential, no sulfopeptide was confidently identified in serum digests by conventional shotgun proteomics, probably due to very low abundance of native sulfoproteins, poor ionization efficiency of sulfopeptides in the positive mode, lack of unambiguous sulfopeptide identification by bioinformatics software. In this context, the use of negative ionization mode with high resolution mass spectrometry appeared promising to improve sensibility and allow sulfopeptide identification in complex samples.
... 12 Covalent modification strategies were used to distinguish the two modifications. 13 As for direct analysis, several fragmentation strategies were investigated over the years, including collision-induced dissociation (CID), 14,15 electron-capture dissociation (ECD), electron-transfer dissociation (ETD), 16 UV photodissociation, 17 negative-ion ECD, 18 ion/ion charge inversion/attachment with dipolar direct current collisional activation, 19 and hydrogen attachment/ abstraction dissociation. 20 Recently, ultrahigh-resolution MS by new generation instrumentation was suggested as a possible solution to this issue in the full scan acquisition mode, while ETD and electron-transfer/collision-induced dissociation (ETciD) could provide information on site localization and the sulfopeptide sequence. ...
Article
Full-text available
Site localization of protein sulfation by high-throughput proteomics remains challenging despite the technological improvements. In this study, sequence analysis and site localization of sulfation in tryptic peptides were determined under a conventional nano-liquid chromatography-mass spectrometry configuration. Tryptic sulfopeptide standards were used to study different fragmentation strategies, including collision-induced dissociation (CID), higher-energy collisional dissociation (HCD), electron-transfer dissociation (ETD), electron-transfer/higher-energy collision dissociation (EThcD), and electron-transfer/collision-induced dissociation (ETciD), in the positive ionization mode. Sulfopeptides displayed only neutral loss of SO3 under CID, while the sequence could be determined for all other tested fragmentation techniques. Results were compared to the same sequences with phosphotyrosine, indicating important differences, as the sequence and modification localization could be studied by all fragmentation strategies. However, the use of metal adducts, especially potassium, provided valuable information for sulfopeptide localization in ETD and ETD-hybrid strategies by stabilizing the modification and increasing the charge state of sulfopeptides. In these conditions, both the sequence and localization could be obtained. In-source neutral loss of SO3 under EThcD provided diagnostic peaks suitable to distinguish the sulfopeptides from the nearly isobaric phosphopeptides. Further confirmation on the modification type was found in the negative ionization mode, where phosphopeptides always had the typical phosphate product ion corresponding to PO3-.
... Besides, as tyrosine sulfation may occur not only on soluble proteins but also on membrane proteins, detection of such PTM turns out even more challenging. [11] Even if the sulfoester function may be in some case intramolecularly strengthened in presence of proximal basic residues, it remains necessary to stabilize it for MS detection of intact sulfoproteins. As non-covalent tag, quaternary ammonium derivatives, initially used as matrix enrichment, proved to be effective in negative ion mode while not being able to completely avoid desulfation. ...
Article
MALDI-MS is of crucial importance for detection and identification of relevant molecules involved in biological pathways allowing a better understanding of physiological processes advantageously related to diagnosis strategies. However, many relevant biomarkers remain undetectable due to several parameters including low molecular weight, weak abundance, MS ionization potential or localization in a complex environment. To overcome these limitations, labeling strategies were developed to target reactive organic moieties of biomolecules of interest. More recently, activity-based probes or antibodies were implemented to match the current need for in vivo biomolecule specific labeling. Furthermore, with the development of MALDI-MS imaging, matrix-free derivatization techniques are currently in the spotlight. MALDI-MS labeling requires the development of a panel of tags designed to target different functional groups promoting signal enhancement for many biomolecules. This review recapitulates recent derivatization methods and available labeling reagents while putting into perspective the importance of MALDI-MS labels in the biochemical research context.
Thesis
Mass spectrometry-based proteomic protocols can identify thousands of expressed proteins with widely varying concentrations. However, acidic post-translational modifications (PTMs), e.g., phosphorylation and sulfation, are difficult to examine, due to the reduced ionization efficiency of highly acidic peptides with positive ion mode nanoelectrospray ionization (nESI). This thesis presents methods for improved detection of acidic, modified peptides and natural product biosynthetic active site peptides in both positive and negative ion mode. Trace addition of trifluoroethanol (TFE) to aqueous samples suppresses corona discharge typically observed in negative ion mode nESI experiments. TFE greatly (~8 fold) improves nESI spray stability without altering observed protein, peptide, and small molecule charge states. This phenomenon is likely due to the highly electronegative fluorine atom’s ability to scavenge electrons, thus stemming plasma formation. In negative ion mode nanoflow liquid chromatography-mass spectrometry (nLC-MS) experiments, TFE addition increases the number of identified peptides by 18%. The relatively simple addition of TFE to sample solutions for improved negative ion mode nESI can be readily employed for improved analysis of widely varied compounds in direct infusion and nLC-MS experiments. The demonstrated compatibility of TFE with nLC-MS allowed for systematic examination of mobile phase and detection polarity effects on peptide identifications. Regardless of mobile phase pH and detection polarity, overall sequence coverage for a six protein digest was similar. However, multiply phosphorylated peptides were only detected at pH 11 in negative ion mode and sulfopeptides were detected most effectively (~55 fold improvement) in negative ion mode and with maximum ion abundance at pH 11. This work demonstrates that alkaline pH separations coupled with negative ion mode nESI provides an efficient method for the detection of highly acidic multiply phosphorylated peptides and sulfopeptides in a background of tryptic peptides typically examined in most proteomic studies. Under typical positive ion mode analysis, sulfopeptides undergo proton mediated loss of the PTM, hampering identification and detection. Alkylamines were found to selectively adduct to sulfopeptides in positive ion mode nESI, allowing for discernment of isobaric phosphorylation and sulfation PTMs without the need for high resolution instrumentation. Alkylamine ion-pairing occurs at 98-99% efficiency regardless of solution pH or base concentration. Characteristic [SO3+alkylamine] neutral losses are observed upon slight collisional activation. This unique transition enables sulfopeptide identification and discovery with positive ion mode data-independent nLC tandem MS. Experiments to discover four sulfopeptide standards in a background of tryptic peptides resulted in 17 sulfopeptide identifications. To our knowledge, the work presented is the first protocol developed for positive ion mode sulfopeptide discovery without the need for tedious chemical modification of a sample proteome prior to analysis. Optimization of tryptic digestion and separation conditions were imperative for MS detection of covalently tethered intermediates in the polyketide synthase (PKS) bryostatin A, module 3 (BryAM3), which introduces a unique beta-branch and O-methylation in the biosynthesis of bryostatin-1, a potent protein kinase C inhibitor. BryAM3 was successfully (98%) phosphopantetheinylated to generate holo acyl carrier protein (ACP). Malonyl extender unit loading on holo ACP was also successfully achieved (68%) utilizing a non-native kirromycin C trans acyl transferase (KirCAT). Unexpectedly, KirCAT also catalyzed malonation of BryAM3 non-active site cysteine residues and direct ACP loading of thiophenol-activated substrates was observed. These experiments indicate that great care must be taken when performing in vitro studies with this and potentially other trans PKSs.
Article
Bacitracin is a mixture of antimicrobial peptides. Although the structures of bacitracins have been fully studied, the structure elucidation of the oxidation products of bacitracin has rarely been reported. In the present work, the bacitracin was treated by hydrogen peroxide solution to form bacitracin sulfonic acids (SO3H) directly, and the structures of bacitracin-SO3Hs were studied by high performance liquid chromatography-quadrupole time of flight-tandem mass spectrometry (HPLC-QTOF-MS/MS). It was demonstrated that the components of bacitracin, such as bacitracin A, B, and C could be site-specifically transformed to the corresponding oxidized product containing the SO3H group. The structure characterization of the oxidized products was elucidated based on the accurate molecular mass and the product ions obtained by MS/MS. Some new kinds of bacitracin J-SO3H, vinyl-bacitracin A-SO3H, bacitracin Y–SO3H and deamino-bacitracin Y–SO3H were identified. Furthermore, we also investigated the novel fragmentation behavior of the oxidized products. This is the first report to propose a neutral loss mechanism of SO3 that explains the characteristic ions of the vinyl-bacitracin A-SO3H and bacitracin Y–SO3H. Based on the structure analysis of the oxidation products of bacitracin, three transformation routes of bacitracin-SO3H generated from bacitracin were hypothesized and systematically summarized for the first time.
Article
Tandem mass spectrometry (MS/MS) is the primary method for discovering, identifying, and localizing post-translational modifications (PTMs) in proteins. However, conventional positive ion mode collision induced dissociation (CID)-based MS/MS often fails to yield site-specific information for labile and acidic modifications due to low ionization efficiency in positive ion mode and/or preferential PTM loss. While a number of alternative methods have been developed to address this issue, most require specialized instrumentation or indirect detection. In this work, we present an amine-reactive TEMPO-based free radical initiated peptide sequencing (FRIPS) approach for negative ion mode analysis of phosphorylated and sulfated peptides. FRIPS-based fragmentation generates sequence informative ions for both phosphorylated and sulfated peptides with no significant PTM loss. Furthermore, FRIPS is compared to positive ion mode CID, electron transfer dissociation (ETD) as well as negative ion mode electron capture dissociation (niECD) and CID, both in terms of sequence coverage and fragmentation efficiency for phospho- and sulfo-peptides. Because FRIPS-based fragmentation has no particular instrumentation requirements and shows limited PTM loss, we propose this approach as a promising alternative to current techniques for analysis of labile and acidic PTMs.
Article
The fragmentation chemistry of protonated sulfoserine was probed using a combination of collision-induced dissociation (CID) mass spectrometry, infrared multiple photon dissociation (IRMPD) spectroscopy, and density functional theory (DFT) calculations. The IRMPD spectra of the dominant fragment ions at m/z 106 and 88 (i.e., loss of SO3 and H2SO4) were obtained and used to determine the corresponding structures. By comparison to a synthetic standard and calculations, it was determined that the m/z 106 ion is structurally identical to protonated serine. The m/z 88 fragment ion was assigned an aziridine structure based on a comparison to theory, analogous to the structure previously proposed by others for phosphoric acid loss from phosphoserine. This work provides the first spectroscopic insights into the dissociation pathways of a sulfated amino acid, laying the groundwork for future studies on related amino acids and peptides with this important, labile post-translational modification.
Article
Sulfation is a common post-translational modification of tyrosine residues in eukaryotes; however, detection using traditional liquid chromatography-mass spectrometry (LC-MS) methods is challenging based on poor ionization efficiency in the positive ion mode and facile neutral loss upon collisional activation. In the present study, 193 nm ultraviolet photodissociation (UVPD) is applied to sulfopeptide anions to generate diagnostic sequence ions, which do not undergo appreciable neutral loss of sulfate even using higher energy photoirradiation parameters. At the same time, neutral loss of SO3 is observed from the precursor and charge-reduced precursor ions, a spectral feature that is useful for differentiating tyrosine sulfation from the nominally isobaric tyrosine phosphorylation. LC-MS detection limits for UVPD analysis in the negative mode were determined to be around 100 fmol for three sulfated peptides, caerulein, cionin, and leu-enkephalin. The LC-UVPD-MS method was applied for analysis of bovine fibrinogen, and its key sulfated peptide was confidently identified. Figure ᅟ
Article
The post-translational modifications sulfation and phosphorylation pose special challenges to mass spectral analysis due to their isobaric nature and their lability in the gas-phase, as both types of peptides dissociate through similar channels upon collisional activation. Here, we present resonant infrared photodissociation, based on diagnostic sulfate and phosphate OH stretches, as a means to differentiate sulfated from phosphorylated peptides within the framework of a mass spectrometry platform. The approach is demonstrated for a number of tyrosine-containing peptides, ranging from dipeptides (YG, pYG, and sYG) over tripeptides (GYR, GpYR, and GsYR), to more biologically relevant enkephalin peptides (YGGFL, pYGGFL, and sYGGFL). In all cases, the diagnostic ranges for sulfate OH stretches are established as 3580-3600 cm-1, and can thus be distinguished from other characteristic hydrogen stretches, such as carboxylic acid OH, alcohol OH, and phosphate OH stretches.
Article
The dissociation of singly or multiply protonated peptide ions by using low-energy collisional activation (CA) is highly dependent on the sites of protonation. The presence of strongly basic amino acid residues in the peptide primary structure dictates the sites of protonation, which generates a precursor ion population that is largely homogeneous with respect to charge sites. Attempts to dissociate this type of precursor ion population by low-energy CA result in poor fragmentation via few pathways. The work described here represents a systematic investigation of the effects of charge heterogeneity in the precursor ion population of a series of model peptides in low-energy CA experiments. Incorporation of acidic residues in the peptide RLC*IFSC*FR (where C* indicates a cysteic acid residue), for example, balances the charge on the basic arginine residues, which enables the ionizing protons to reside on a number of less basic sites along the peptide backbone. This results in a precursor ion population that is heterogeneous with respect to charge site. Low-energy CA of these ions results in diverse and efficient fragmentation. Molecular modeling has been utilized to demonstrate that energetically preferred conformations incorporate an intraionic interaction between arginine and cysteic acid residues.
Article
Time- and collision-energy-resolved surface-induced dissociation (SID) of protonated peptides containing phosphoserine (s) was studied using a Fourier transform ion cyclotron resonance (FT-ICR) mass spectrometer configured for SID experiments. We examined fragmentation of three singly protonated peptides: GGGsGGG, KGGsGGG, and RGGsGGG. Fragmentation of GGGsGGG occurs under the mobile proton condition, while the ionizing proton is sequestered by the basic residue, resulting in the nonmobile proton condition in dissociation of the two basic peptides: KGGsGGG and RGGsGGG. RRKM modeling of the experimental data demonstrates that the energetics and dynamics of H3PO4 loss are different under mobile and nonmobile proton conditions. Specifically, fragmentation of GGGsGGG is characterized by a higher dissociation barrier, 1.68 eV, and higher activation entropy, 11 e.u. (e.u. = entropy unit), than fragmentation of the basic peptides. Similar threshold energies of 1.36 eV and 1.40 eV and activation entropies of -4.9 e.u. and 0.3 e.u. were obtained for KGGsGGG and RGGsGGG, respectively. We propose that the loss of H3PO4 from phosphoserine is a two-step process, in which the phosphate abstraction from the phosphorylated side chain is followed by dissociation of the ion-molecule complex.
Article
Liquid chromatography/electrospray ionization mass spectrometry was used to investigate the peptide composition of the venom of Conus pennaceus, a molluscivorous cone shell from the Red Sea. Based on observed M(r)s, this venom contained all known conotoxins previously isolated and identified from this species. Interestingly, the doubly protonated species of only two of these conotoxins, alpha-PnIA and alpha-PnIB, showed additional related ions at + 40 m/z (+ 80 Da), indicating the presence of either sulfation or phosphorylation in both components. Highperformance liquid chromatographic (HPLC) fractions containing these two conotoxins were examined by matrix-assisted laser desorption/ionization (MALDI) mass spectrometry in both positive and negative ion modes, as well as by MALDI high-energy collision-induced dissociation. These experiments established the presence of a single sulfated tyrosine residue within both alpha-PnIA and alpha-PnIB, Hence their post-translationally modified sequences are GCCSLPPCAANNPDY(S)C-NH2 (alpha-PnIA) and GCCSLPPCALSNPDY(S)C-NH, (alpha-PnIB), This assignment was supported by comparison of their mass spectral behavior with that of known sulfated and phosphorylated peptides, This data clarified further the distinguishing features of the ionization and fragmentation of such modified peptides, Selective disulfide folding of synthetic alpha-PnIB demonstrated that both sulfated and non-sulfated toxins co-elute on reversed-phase HPLC and that alpha-PnIB possesses the same disulfide connectivity as other 'classical' alpha-conotoxins reported previously. Copyright (C) 1999 John Wiley & Sons, Ltd.
Article
A variation of Gaussian-3 (G3) theory is presented in which the geometries and zero-point energies are obtained from B3LYP density functional theory [B3LYP/6-31G(d)] instead of geometries from second-order perturbation theory [MP2(FU)/6-31G(d)] and zero-point energies from Hartree–Fock theory [HF/6-31G(d)]. This variation, referred to as G3//B3LYP, is assessed on 299 energies (enthalpies of formation, ionization potentials, electron affinities, proton affinities) from the G2/97 test set [J. Chem. Phys. 109, 42 (1998)]. The G3//B3LYP average absolute deviation from experiment for the 299 energies is 0.99 kcal/mol compared to 1.01 kcal/mol for G3 theory. Generally, the results from the two methods are similar, with some exceptions. G3//B3LYP theory gives significantly improved results for several cases for which MP2 theory is deficient for optimized geometries, such as CN and O2+. However, G3//B3LYP does poorly for ionization potentials that involve a Jahn–Teller distortion in the cation (CH4+, BF3+, BCl3+) because of the B3LYP/6-31G(d) geometries. The G3(MP2) method is also modified to use B3LYP/6-31G(d) geometries and zero-point energies. This variation, referred to as G3(MP2)//B3LYP, has an average absolute deviation of 1.25 kcal/mol compared to 1.30 kcal/mol for G3(MP2) theory. Thus, use of density functional geometries and zero-point energies in G3 and G3(MP2) theories is a useful alternative to MP2 geometries and HF zero-point energies. © 1999 American Institute of Physics.
Article
Relative energetics of fragmentation of protonated peptides are investigated by using electrospray ionization/surface-induced dissociation (ESI/SID) tandem mass spectrometry. ESI/SID fragmentation efficiency curves (percent fragmentation versus laboratory collision energy) are presented for 20 oligopeptides and are a measure of how easily a peptide fragments. The relative positions of the ESI/SID fragmentation efficiency curves depend on several parameters which include peptide composition (e.g., presence/absence of a basic amino acid residue) and peptide size. The ESI/SID fragmentation efficiency curves, in combination with quantum chemical calculations, provide a unique approach to substantiate and refine the mobile proton model for peptide fragmentation. Selected peptides are also investigated to further test and confirm the mobile proton model; these include doubly-protonated peptides and chemically-modified peptides (i.e., acetylated and fixed-charge derivatized peptides). Doubly-protonated peptides fragment more easily than the singly-protonated forms of the same peptides, with a sequence dependence for the difference in energy required for the fragmentation of singly- vs doubly-protonated peptides. Acetylation at the amino terminus and arginine side chain leads to a decrease in basicity and a corresponding lower energy onset for fragmentation than for the unmodified form of the peptide. Fixing the site of charge by addition of trimethylammonium acetyl to the amino terminus, i.e., eliminating the mobile proton, results in a higher energy onset than that for the protonated form of the same peptide. Curves for doubly protonated peptides with two adjacent basic residues (Arg, Arg) suggest the localization of the two protons at the two basic side chains rather than at opposite termini of the peptide.
Article
A theoretical framework and an accompanying computer program (MassKinetics, www.chemres.hu/ms/ masskinetics) is developed for describing reaction kinetics under statistical, but non-equilibrium, conditions, i.e. those applying to mass spectrometry. In this model all the important physical processes influencing product distributions are considered: reactions, including the effects of acceleration, collisions and photon exchange. These processes occur simultaneously and are taken into account by the master equation approach. The system is described by (independent) product, kinetic energy and internal energy distributions, and the time development of these distributions is studied using transition probability functions. The product distribution at the end of the experiment corresponds to the mass spectrum. Individual elements in this scheme are mostly well known: internal energy-dependent reaction rates are calculated by transition state theory (RRK or RRKM formalisms). In the course of collisions, energy transfer and other processes may occur (the latter usually resulting in the ‘loss’ of ion signal). Collisions are characterized by their probability and by energy transfer in a single collision. To describe single collisions, three collision models are used: long-lived collision complexes, partially inelastic collisions and partially inelastic collisions with cooling. The latter type has been developed here, and is capable of accounting for cooling effects occurring in collision cascades. Descriptions of photon absorption and emission are well known in principle, and these are also taken into account, in addition to changes in kinetic energy due to external (electric) fields. These changes in the system occur simultaneously, and are described by master equations (a set of differential equations). The usual form of the master equation (taking into account reactions and collisional excitation) was extended to consider also radiative energy transfer, kinetic energy changes, energy partitioning and ion loss collisions. Initial results show that close to experimental accuracy can be obtained with MassKinetics, using few or no adjustable parameters. The model/program can be used to model almost all types of mass spectrometric experiments (e.g. MIKE, CID, SORI and resonant excitation). Note that it was designed for mass spectrometric applications, but can also be used to study reaction kinetics in other non-equilibrium systems. Copyright © 2001 John Wiley & Sons, Ltd.