Download Mass spectrometry of oligosaccharides

Document related concepts

Molecular neuroscience wikipedia , lookup

Membrane potential wikipedia , lookup

Evolution of metal ions in biological systems wikipedia , lookup

Metalloprotein wikipedia , lookup

Metabolomics wikipedia , lookup

Protein mass spectrometry wikipedia , lookup

Transcript
MASS SPECTROMETRY OF OLIGOSACCHARIDES
Joseph Zaia*
Department of Biochemistry, Boston University School of Medicine, 715
Albany St., R-806, Boston, Massachusetts 02118
Received 20 March 2003; received (revised) 6 August 2003; accepted 6 August 2003
I.
II.
III.
IV.
V.
Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
Characteristics of Tandem Mass Spectra of Carbohydrates . . . . . . . . . . . . . . . . . . . . . . . . . . . .
A. Ionization of Carbohydrates . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
1. Electrospray Ionization (ESI) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
2. Matrix-Assisted Laser Desorption/Ionization (MALDI) . . . . . . . . . . . . . . . . . . . . . . . .
B. Nomenclature for the Fragmentation of Glycoconjugates . . . . . . . . . . . . . . . . . . . . . . . . . .
C. Tandem MS of Native Oligosaccharide Molecular Ions . . . . . . . . . . . . . . . . . . . . . . . . . . .
1. Protonated Ions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
2. Deprotonated Ions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
3. Alkali and Alkaline Earth Adducted Ions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
D. Tandem MS of Permethylated and Peracetylated Oligosaccharides . . . . . . . . . . . . . . . . . .
E. Tandem MS of Reductively Aminated Carbohydrates . . . . . . . . . . . . . . . . . . . . . . . . . . .
F. Discrimination of Monosaccharide Linkages . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
G. Gas-Phase Degradation of Oligosaccharides . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
H. Computer-Based Approaches for Interpretation of Oligosaccharide Product-Ion Mass Spectra
I. Internal Residue Loss Rearrangements of Oligosaccharide Ions During CID . . . . . . . . . . . .
J. Conclusions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
Analyzers for Mass Spectrometry of Carbohydrates . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
A. Analysis of Permethylated Carbohydrates Using High Temperature GC/MS . . . . . . . . . . . .
B. Analysis of Carbohydrates with MALDI-TOF MS . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
C. Analysis of Carbohydrates with MALDI Q-oTOF MS . . . . . . . . . . . . . . . . . . . . . . . . . . .
D. Analysis of Carbohydrates with ESI Q-oTOF MS . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
E. Analysis of Carbohydrates with QIT MS . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
F. Analysis of Glycoconjugates with FT MS . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
G. Conclusions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
Tandem Mass Spectrometry of Glycopeptides . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
A. Ionization of Glycopeptides . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
B. CID of Glycopeptides . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
1. Selective Identification of Glycopeptides with Tandem MS . . . . . . . . . . . . . . . . . . . . .
2. CID of O-Linked Glycopeptides . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
3. CID of N-Linked Glycopeptides . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
C. Electron Capture Dissociation of Glycopeptides . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
D. Conclusions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
Mass Spectrometry of Sialylated Glycoconjugates . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
A. Permethylation of Sialylated Oligosaccharides . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
B. MALDI-MS of Sialylated Glycoconjugates . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
1. Anionic Dopants for Analysis of Sialylated Glycoconjugates . . . . . . . . . . . . . . . . . . . .
2. Methyl Esterification to Stabilize Sialic Acid Residues . . . . . . . . . . . . . . . . . . . . . . . .
3. Perbenzolylation to Stabilize Sialic Acid Residues . . . . . . . . . . . . . . . . . . . . . . . . . . .
4. High-Pressure MALDI of Sialylated Glycoconjugates . . . . . . . . . . . . . . . . . . . . . . . . .
————
Contract grant sponsor: NIH/NCRR; Contract grant number: P41RR10888; Contract grant sponsor: Glycosciences Research Award,
Neose Technologies.
*Correspondence to: Joseph Zaia, Department of Biochemistry,
Boston University School of Medicine, 715 Albany St., R-806, Boston,
MA 02118. E-mail: [email protected]
Mass Spectrometry Reviews, 2004, 23, 161– 227
# 2004 by Wiley Periodicals, Inc.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
162
. 163
. 163
. 163
. 163
. . 164
. 165
. 165
. 166
. 166
. 169
. 170
. 171
. 178
. 181
. 183
. 184
. 186
. 186
. 186
. 187
. 187
. 187
. 188
. 188
. 189
. 189
. 190
. 190
. 191
. 191
. 193
. 193
. 195
. 195
. 195
. 196
. 196
. 196
. 196
&
ZAIA
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
197
198
199
199
199
199
199
202
202
202
203
204
205
206
206
206
207
208
208
209
211
211
214
214
215
Abbreviations . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
215
References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
216
VI.
VII.
VIII.
C. ESI MS of Sialylated Oligosaccharides . . . . . . . . . . . . . . . . . . . . . . . . .
D. Tandem MS of Sialylated Oligosaccharides . . . . . . . . . . . . . . . . . . . . . . .
E. Conclusions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
Mass Spectrometry of Sulfated Oligosaccharides . . . . . . . . . . . . . . . . . . . . . .
A. Derivatization . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
B. Ionization Methods . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
1. Fast Atom Bombardment . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
2. MALDI . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
a. MALDI of Sulfated Peptides . . . . . . . . . . . . . . . . . . . . . . . . . . .
b. Direct MALDI of Sulfated Oligosaccharides . . . . . . . . . . . . . . . .
c. Use of Basic Peptides for MALDI of Polysulfated Oligosaccharides
d. MALDI Analysis of Protein-Sulfated Oligosaccharide Complexes .
3. ESI of Sulfated Oligosaccharides . . . . . . . . . . . . . . . . . . . . . . . . . . .
4. On-Line Separation Systems for Sulfated Carbohydrates . . . . . . . . . . .
C. Tandem MS of Sulfated Oligosaccharides . . . . . . . . . . . . . . . . . . . . . . . .
1. Lessons from CID of Sulfated Peptides . . . . . . . . . . . . . . . . . . . . . . .
2. Tandem MS of Mono- and Di-Sulfated Oligosaccharides . . . . . . . . . . .
3. Precursor-Ion and Neutral-Loss Scans for Sulfated Glycoconjugates . . .
4. Determination of Positional Sulfation Isomers in GAG Disaccharides . .
5. Tandem Mass Spectrometric Quantification of GAG Disaccharides . . . .
6. Tandem Mass Spectrometric Analysis of GAG Oligosaccharides . . . . .
a. CS Oligosaccharides . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
b. Heparin/HS Oligosaccharides . . . . . . . . . . . . . . . . . . . . . . . . . . .
D. Conclusions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
Overall Conclusions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
Glycosylation is a common post-translational modification to cell
surface and extracellular matrix (ECM) proteins as well as to
lipids. As a result, cells carry a dense coat of carbohydrates on
their surfaces that mediates a wide variety of cell–cell and cell–
matrix interactions that are crucial to development and function.
Because of the historical difficulties with the analysis of complex
carbohydrate structures, a detailed understanding of their roles
in biology has been slow to develop. Just as mass spectrometry
has proven to be the core technology behind proteomics, it stands
to play a similar role in the study of the functional implications of
carbohydrate expression, known as glycomics. This review
summarizes the state of knowledge for the mass spectrometric
analysis of oligosaccharides with regard to neutral, sialylated,
and sulfated compound classes. Mass spectrometric techniques
for the ionization and fragmentation of oligosaccharides are
discussed so as to give the reader the background to make
informed decisions to solve structure-activity relations in
glycomics. # 2004 Wiley Periodicals, Inc., Mass Spec Rev
23:161–227, 2004
Keywords: carbohydrates; oligosaccharides; mass spectrometry; sialylated; sulfated; glycomics
I. INTRODUCTION
With progress in proteomics comes an increasing interest
in the importance of glycosylation. Most cell-surface and
162
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
secreted proteins are glycosylated—a fact that impacts on
efforts to understand the biological relevance of specific
protein expression and modification patterns. Unlike the
core proteins, glycans are expressed as a set of variations on
a core structure and are polydisperse in nature. Therefore,
glycosylation increases the complexity of protein molecules and causes them to migrate as diffuse bands or
spots on SDS–PAGE gels to complicate efforts to identify
protein expression patterns that correlate with disease
states. Glycans serve to modulate protein function and may
influence folding, biological lifetime, and recognition of
binding partners. For example, surface carbohydrates serve
as the interface between the cell and its environment, and
define self versus non-self. Many pathogens recognize
particular carbohydrates, and structural studies have led to
progress in this area.
Mass spectrometry is an important tool for the structural analysis of carbohydrates, and offers precise results,
analytical versatility, and very high sensitivity. Whereas
mass spectrometric analysis options for proteins and
peptides are well-defined relative to those for carbohydrates, tandem mass spectrometric product-ion patterns are
more complex, and the results depend on the types of
derivative and precursor ion used. In most cases, one
analyzes protonated forms of underivatized proteins and
peptides and collisional-induced dissociation (CID) results
MASS SPECTROMETRY OF OLIGOSACCHARIDES
in definition of a complete sequence (Biemann & Martin,
1987) or of a partial sequence that is useful in the identification of modified peptide residues (Papayannopoulos,
1995) or in database searching (Mann & Wilm, 1994).
With carbohydrates, one has the choice of using peralkyl
or reducing terminal derivatives, or of analyzing native
structures. One also has the choice of protonated, metalcationized, or deprotonated ions. As a result, the carbohydrate mass spectrometry field is reflected by a divergence of
published structural analysis methods in recent years. For
protein mass spectrometrists who wish to learn about
carbohydrates and glycobiologists who wish to learn mass
spectrometry, the plethora of publications that describe
seemingly divergent approaches may seem overwhelming.
The intent of this review is to summarize the historical and
recent developments in carbohydrate mass spectrometry to
facilitate informed decision-making on analysis routes.
Section II explains common ionization methods used
for carbohydrates as well as characteristic fragmentation
methods for different carbohydrate compound classes.
Section III describes the characteristics of different modern
mass analyzers used for carbohydrates. The remaining
sections specifically address the analysis of glycopeptides
(Section IV), sialylated glycoconjugates (Section V), and
sulfated oligosaccharides (Section VI).
II. CHARACTERISTICS OF TANDEM MASS
SPECTRA OF CARBOHYDRATES
A. Ionization of Carbohydrates
1. Electrospray Ionization (ESI)
Conventional ESI MS (Meng, Mann, & Fenn, 1988; Fenn
et al., 1990) involves the pumping of a solution (a forced
flow) into the ion source, and has been observed to produce
relatively weak ion signals for native oligosaccharides
compared to those for peptides and proteins (Burlingame,
Boyd, & Gaskell, 1994; Reinhold, Reinhold, & Costello,
1995). Nano ESI (Wilm & Mann, 1994), on the other hand,
produces ion signals that are comparable between the
peptide and carbohydrate compound classes (Bahr et al.,
1997). It, therefore, appears that the hydrophilicity of
oligosaccharides limits the surface activity in ESI droplets
and that, with small droplets, their sensitivity is significantly enhanced. The results with nano ESI are consistent
with the conclusion that the sensitivity increase observed
when oligosaccharides are derivatized, reducing their
hydrophilicity, is due to an increase in surface activity
rather than in volatility (Karas, Bahr, & Dulcks, 2000). The
fact that the ESI of carbohydrates appears to be more
effective with the nano scale, has important implications.
Interfaces for on-line ESI LC/MS typically produce droplet
sizes that are large relative to those produced by spraying
&
from a 1–2 mm orifice nanospray emitter, and thus the spray
characteristics are typical of forced flow. An exception
appears to be the use of 10 mm fused-silica spray tips with
post-column splitting of HPLC effluents (Gangl et al.,
2001). It may be possible to design a continuous-flow
interface that produces droplet sizes small enough to
produce nanospray-like ionization responses for carbohydrates that elute from a chromatography column.
2. Matrix-Assisted Laser
Desorption/Ionization (MALDI)
The MALDI-TOF ionization efficiency for neutral carbohydrates oligomers has been observed to be constant as the
size of the molecule increases, in contrast to that for ESI,
where the ionization efficiency decreases with an increasing molecular weight (Harvey, 1993). Although the
ionization response drops off with increasing molecular
weight for mixtures of large carbohydrate polymers, the
use of MALDI for neutral oligosaccharide analysis has
advantages over ESI, particularly for applications that
involve the profiling of mixtures released from glycoproteins. Quantitation of permethylated carbohydrate mixtures
with MALDI has been demonstrated to be reproducible,
and the coefficients of variation for those analyses equal
those obtained for the same oligosaccharide mixture after
derivatization with a chromophore and chromatographic
quantitation (Viseux et al., 2001). In addition, information
on fucosylation was more detailed with the mass spectrometric approach.
The advantages of MALDI in terms of ionization
response have to be balanced against the disadvantages of
the metastable fragmentation that is caused by the higher
internal energies imparted to the ions relative to those
resulting from ESI. Although that fragmentation allows the
analysis of carbohydrate ions with post-source decay
(PSD) on a MALDI-TOF instrument (Viseux, Costello, &
Domon, 1999), it complicates MS profiling. Extensive
fragmentation of sialic acid residues necessitates either
the use of linear mode MALDI-TOF to avoid separating the
metastable products (see Section V.D), or of derivatization to stabilize those groups (See Section V.B.2). As a
further complication, detection of sialylated glycoconjugates with Fourier transform (FT) MS is limited by the
metastable fragmentation in a low-pressure MALDI source
and is mitigated through the use of collisional cooling with
high pressure (O’Connor & Costello, 2001; O’Connor,
Mirgorodskaya, & Costello, 2002). Atmospheric-pressure
sources have been developed for quadrupole ion trap (QIT)
instruments (Laiko, Moyer, & Cotter, 2000; Moyer &
Cotter, 2002; Moyer et al., 2002), and it is likely that those
sources will prove to be useful for the analysis of oligosaccharides. The development of a MALDI TOF/TOF
analyzer has been described for the analysis of peptides
163
&
ZAIA
(Medzihradszky et al., 2000; Yergey et al., 2002), and
carries the potential benefits of producing high-energy
fragmentation. Metastable fragmentation has been a
problem with that instrument, and it is likely that its
effective use for the analysis of oligosaccharides will
require a high-pressure source to limit metastable fragmentation, and thereby allow true high-energy CID. The
benefits of high-energy CID of carbohydrates are welldocumented (Domon & Costello, 1988b; Gillece-Castro &
Burlingame, 1990; Peter-Katalinic, 1994); see Section II.F
for a detailed discussion.
B. Nomenclature for the Fragmentation
of Glycoconjugates
Collisional-induced dissociation (CID) of glycoconjugates
results in the observation of ions that correspond to
cleavage of the oligosaccharide portion of the molecule.
Typically, those ions are produced in greater abundances
for the oligosaccharide portion than are those that occur in
the aglycon (non-oligosaccharide) portion of glycoconjugates. The nomenclature for oligosaccharide fragmentation used throughout the mass spectrometry field is shown
in Figure 1. Fragment ions that contain a non-reducing
terminus are labeled with uppercase letters from the
beginning of the alphabet (A, B, C), and those that contain
the reducing end of the oligosaccharide or the aglycon are
labeled with letters from the end of the alphabet (X, Y, Z);
subscripts indicate the cleaved ions. The A and X ions are
produced by cleavage across the glycosidic ring, and are
labeled by assigning each ring bond a number and counting
clockwise. Figure 1 shows examples for two cross-ring
cleavage ions. Ions produced from cleavage of successive
residues are labeled: Am, Bm, and Cm, with m ¼ 1 for the
non-reducing end and Xn, Yn, and Zn, with n ¼ 1 for
the reducing-end residue. Note that Y0 and Z0 refer to the
fragmentation of the bond to the aglycon.
When considering tandem MS of oligosaccharides,
there are several options regarding the state of the precursor
ion, the choice of which will dramatically influence the
product-ion pattern, and thus the structural information,
produced. For native oligosaccharides, the options in-
clude protonated molecule ions, [M þ nH]nþ, deprotonated
molecule ions, [M–nH]n, and natriated molecule ions,
[M þ Na]þ. Permethylated and peracetylated derivatives
are analyzed as either [M þ H]þ or [M þ Na]þ with
MALDI, or as multiply charged ions with ESI. Table 1
shows the combination of ionization states and derivatives
that have been studied in the literature. Many of the principles of carbohydrate fragmentation with tandem MS were
developed with fast atom bombardment (FAB) ionization
(also known as liquid secondary ion mass spectrometry),
and this review takes those principle into consideration
along with later developments.
Note that metal-adducted ions, in which charge is
carried by protons, will be denoted [M(X) þ H]þ. This
nomenclature assumes the adduction of the metal cation
with the displacement of a proton, and the positive ion is
formed by the addition of a proton. It is understood that
the sodium cation is paired with an acidic group. This
nomenclature is used here and elsewhere (Costello, Juhasz,
& Perreault, 1994; O’Connor, Mirgorodskaya, & Costello,
2002; Zaia & Costello, 2003) because it simplifies the
notation of precursor and product ions, and facilitates the
labeling in tandem mass spectra. Note that cationized ions,
in which charge is carried by the metal, are given as
[M þ mX]mnþ, where m ¼ the number of metal ions and
n ¼ the cationic charge.
Tandem MS of carbohydrates was first studied with
FAB ionization, a technique that imposed severe constraints on the analysis. Neutral (containing no amino
groups) and basic (containing hexosamine residues)
carbohydrates ionize relatively poorly with FAB, and the
tandem MS analysis was practically limited to molecules
of <1000 Da. Acidic oligosaccharides (those containing
Neu5Ac or sulfate) were observed to produce relatively
strong ions with negative FAB (Egge & Peter-Katalinc,
1987). The ion signals were observed to be much stronger
when the oligosaccharides or glycoconjugates were permethylated or peracetylated. The FAB ionization process
was also energetic enough to produce glycosidic-bond and
cross-ring cleavages—a phenomenon that was used to
obtain structural information on molecules that had been
purified to homogeneity (Dell et al., 1983a,b; Egge, Dell, &
Von Nicolai, 1983; Carr & Reinhold, 1984). It was also
observed that neutral gas collisions enhanced the yield of
fragment ions generated from oligosaccharides (Carr et al.,
1985). The development of tandem mass analyzers for
TABLE 1. Ionization options for tandem MS of oligosaccharides
FIGURE 1. Nomenclature for glycoconjugate product ions generated by
tandem MS (Modified from Domon & Costello, 1988b).
164
MASS SPECTROMETRY OF OLIGOSACCHARIDES
biological macromolecules allowed the selection of individual precursor ions and subsequent fragmentation, thus
eliminating the need to purify oligosaccharides to homogeneity. Using FAB tandem MS, it became generally
accepted to analyze sodium adducts of permethylated
oligosaccharides as [M þ Na]þ ions—as will be discussed
in detail below.
The development of ESI greatly expanded the capabilities of the tandem MS of oligosaccharides by improving,
relative to FAB, the strength of the ion signals produced
from a given quantity of oligosaccharide. Although the
permethylated oligosaccharides were observed to produce
the strongest signals, the direct analysis of native oligosaccharides was possible for larger precursor ions than was
possible with FAB. That increase in sensitivity was such
that researchers had the choice to ionize native or permethylated oligosaccharides or glycoconjugates as protonated, deprotonated, or alkali-adducted ions. Because the
information content of product-ion mass spectra depends
to a large extent on the state of the precursor ion, it is
important to examine the available information in this
area. The following sections summarize the fragmentation behavior of oligosaccharides and oligosaccharide
portions of glycoconjugates with regard to the state of
the precursor ion so as to allow the reader to choose the
appropriate analysis options. Recent work in this area is
summarized, with regard to overall trends for carbohydrate
fragmentation.
Product-ion mass spectra of glycoconjugates are
considerably more complicated than are those of peptides
because of their branching structure. For such branching
structures, fragmentation occurs from the non-reducing
end of each antenna and from the reducing end to give
rise to a higher complexity. The input of energy into the
molecule by collision is most likely to break single bond
glycosidic linkages. Using low energy CID, fragmentation
of glycosidic linkages is most likely; fragmentation across
the sugar rings is less so because two covalent bonds must
be cleaved. It does occur, however, and provides important
information on the location of substituents on branching
monosaccharide residues. Research has shown that CID
product-ion mass spectra provide information on the
stereochemistry of individual sugar residues (Mueller
et al., 1988), the linkage position (Laine et al., 1988), and
branching structure (Carr et al., 1985; Domon & Costello,
1988a). Oligosaccharides that contain the same monosaccharides linked with a different branching structure
often show distinct product-ion patterns because the steric
environments differ between such isomers and result in
different bond energies and ion abundances in product-ion
mass spectra (Laine et al., 1988). Today, it is not possible to
determine all of the linkages and branching patterns for a
complex branched oligosaccharide. However, much work
has been conducted in the past 15 years to improve the
&
degree to which the product-ion patterns generated from
those molecules may be directly interpreted, as opposed to
their being used as mass fingerprints, whose pattern reflects
but does not predict the structure.
C. Tandem MS of Native Oligosaccharide
Molecular Ions
1. Protonated Ions
The CID fragmentation patterns differ according to the
oligosaccharide sequence, size, and class of subunits, as
discussed below, and are used to differentiate subtle
structural differences. CID fragmentation of [M þ H]þ
ions generated from asialo glycoconjugates results in the
observation of abundant Bm and Yn ions; those patterns are
useful in assigning the oligosaccharides sequence (Domon
& Costello, 1988b). Cleavage to the reducing side of
HexNAc residues is favored, resulting in abundant fragment ions (Egge, Dell, & Von Nicolai, 1983). Therefore,
such fragmentation is useful to define the presence of
lactosamine (Gal-GlcNAc) disaccharides in the antennae
of N-linked oligosaccharides. CID fragmentation of neutral
glycoconjugates (gangliosides) results in the observation
of abundant Bm and Yn ions, indicating that charge is not
localized in those ions (Domon & Costello, 1988b). The
abundances of ions produced from cross-ring cleavages is
comparatively low for [M þ H]þ ions generated from that
compound class. Those results are similar to the metastable
fragmentation observed for those molecules in FAB MS
(Egge & Peter-Katalinc, 1987; Gunnarson, 1987).
When the protonated ions generated from small
oligosaccharides are subjected to CID, a greater diversity
of information is produced than that obtained from
large oligosaccharides (Dell, 1990; Gillece-Castro &
Burlingame, 1990). Specifically, glycosidic bond cleavages are more likely to predominate in product-ion mass
spectra of large ions, with the result that less detailed information regarding the branching structure is obtained.
Comparatively more subtle information, with regard to
cross-ring cleavage ions, may thus be obtained for small
ions. Those observations have led to efforts to increase the
information content of product-ion mass spectra.
Conditions during ESI MS of native oligosaccharides
can be manipulated to favor the production of protonated,
natriated, or ammoniated precursor ions (Duffin et al.,
1992). Natriated oligosaccharides are produced in the presence of a low concentration of sodium acetate. Ammoniated adducts formed in the presence of ammonium acetate
readily decay to protonated ions due to the energy required
to desolvate ions during the electrospray process. Sialylated oligosaccharides were observed to produce abundant
[M–H] ions in the negative-ion mode with ESI MS. The
yield of glycosidic bond product ions in high-energy CID
165
&
ZAIA
was observed to be higher for protonated relative to
natriated native oligosaccharide ions (Orlando, Bush, &
Fenselau, 1990; Duffin et al., 1992). Cleavages to the
reducing side of HexNAc residues are observed in
abundance, particularly to those residues located on
complex N-linked antennae. High-mannose oligosaccharides produce a series of product-ions from successive
losses of hexose residues, resulting in a generally more
uniform product ion pattern than observed for complex Nlinked sugars that contain HexNAc residues near the nonreducing termini. Thus, the pattern of CID product ions is
useful to define the class of N-linked oligosaccharide (highmannose, hybrid, or complex).
2. Deprotonated Ions
Neutral oligosaccharides have been observed to ionize
relatively poorly relative to their basic (amino sugarcontaining) and acidic counterparts. Sialylated oligosaccharides are amenable to analysis as [M–H] ions, and are
observed to produce product ions as abundant as those
from metastable fragmentation with FAB ionization
(Derappe et al., 1986; Egge & Peter-Katalinc, 1987). A
series of molecules with Neu5Ac at the non-reducing
terminus fragment to produce ions that almost exclusively
contain the acidic group. This is contrasted with the
fragmentation of an asialo ganglioside in which both
reducing and non-reducing fragmentation results (Domon
& Costello, 1988b). Fragmentation to the reducing side of
HexNAc is facile in the CID of [M–H] ions.
The importance of mobile protons in analyzing the
fragmentation pattern becomes evident when comparing
studies of deprotonated polysaccharides with their protonated forms. Interestingly, the high-energy CID production mass spectrum of the [M–H] ion generated from
maltoheptaose (1,4-linked glucose heptamer) displays
only 2,4Am ions (Gillece-Castro & Burlingame, 1990). By
contrast, the [M–H] ion produced from the chitobiose
core oligosaccharide, a branched (Man)2-Man-GlcNAcGlcNAc structure, displays abundant cross-ring cleavages
with limited glycosidic bond fragmentation. The linear
maltoheptaose spectrum displays a smooth decrease in ion
abundances as the m/z values decrease, whereas the pattern
generated from the chitobiose core is defined by a very
abundant 2,4A3 ion. The discontinuous product-ion pattern
of the chitobiose indicates branching. Protonated ions
generated from the same molecules produce very abundant
glycosidic bond cleavages with the very low abundance of
cross-ring cleavages.
The profusion of glycosidic bond cleavages of native
oligosaccharides in CID product-ion mass spectra seems
to be related to the presence of free hydroxyl groups.
Hydroxylic hydrogen migration has been implicated in
glycosidic-bond cleavages (Hofmeister, Zhou, & Leary,
166
1991), and the abundance of such ions appear to limit the
abundance of cross-ring cleavage ions. In accordance,
certain cross-ring cleavage ions, observed in the production mass spectra of permethylated oligosaccharides, have
been observed to be absent from spectra acquired on native
oligosaccharides (Harvey, Bateman, & Green, 1997).
In protonated or alkali-cationized oligosaccharides,
the presence of HexNAc residues places a charge on the
amide nitrogen. The proximity of that charge to the
reducing-side glycosidic oxygen evidently predisposes that
bond to scission. Although a similar pattern is observed for
deprotonated oligosaccharides, it is not observed for
chondroitin sulfate glycosaminoglycans which consist of
[UroA(1–3)GalNAcSulfate(1–4)]n, in the negative ESI
mode (Zaia, McClellan, & Costello, 2001; McClellan et al.,
2002). The CID production mass spectrum of a triply
charged CS hexamer (n ¼ 3) results in abundant glycosidic
bond cleavages to the reducing side of uronic acid residues
but not to the reducing side of GalNAc residues. This
difference is evidence that the glycosidic bond to the
reducing side of GalNAc is significantly less labile than is
the bond to the reducing side of UroA residues. This result
is in marked contrast to the observations made for
deprotonated oligosaccharides with FAB MS (Domon &
Costello, 1988b). For further discussion on this issue, see
Section VI.C.6.a.
3. Alkali and Alkaline Earth Adducted Ions
The fragmentation of natriated native oligosaccharides by
high-energy CID results in different product-ion patterns
than are obtained from protonated ions (Orlando, Bush, &
Fenselau, 1990). Specifically, the production of cross-ring
cleavages is enhanced relative to the levels observed in
the product-ion mass spectra of protonated molecular
ions. Furthermore, the abundance of ions produced from
cross-ring cleavages were reduced as the collision energy
was lowered from 8 to 2 keV; that reduction indicates that
the use of high-energy CID increases the information
content of product-ion spectra.
The Low-energy CID spectra of monolithiated disaccharides show abundant reducing end cleavages, but
little differentiation between different linkages. Dilithiated
disaccharides produce product-ion patterns that readily
differentiate linkage isomers (Zhou, Ogden, & Leary,
1990). The dilithiated ions have lithium substitution for a
hydroxyl hydrogen, and that substitution is believed to
initiate the ring-cleavage rearrangements. Using [M(Li) þ
Li]þ ions, it was possible to assign the linkages of Glc(1–
6)Glc(1–6)Glc(1–6)Glc versus the (1–4) linked isomer.
PSD allows the analysis of product-ions that result
from low-energy processes. PSD of natriated ions produced from native oligosaccharides resulted in an abundant series of glycosidic-bond and cross-ring cleavages.
MASS SPECTROMETRY OF OLIGOSACCHARIDES
Cleavages adjacent to HexNAc residues were facile, as
observed for other dissociation techniques (Spengler et al.,
1994). PSD analysis of natriated and protonated forms
of Gal(b1-3)-GlcNAc(b1,3)-Gal(b1-3)-Glc-Bz (benzyl
amino derivative) is instructive (Lemoine, Chirat, &
Domon, 1996). The natriated form produces abundant Bm
and Yn ions, indicating that charge is located without
preference to either end of the molecule. The protonated
form produces primarily Yn ions, indicating that the charge
resides on the basic benzylamino group on the reducing
terminus. The B2 ion is also observed from facile cleavage
of the GlcNAc residue. PSD of a series of protonated N- and
O-linked benzylamino structures was also shown to result
in predominantly Yn ions.
MALDI FTMS studies on alkali cationized native
oligosaccharides have shown that the yield of fragmentation correlates with the degree of branching (Cancilla et al.,
1996). Fragmentation yields were highest for oligomers
with the least branching, and were inversely related to
cation size, following the order Hþ>Liþ>Naþ>Kþ>
Rbþ>Csþ. Mechanism of fragmentation of protonated ions
is likely to be charge-induced, whereas that for cesiated
ions is likely to be charge-remote. Charge-remote fragmentation requires more energy than charge-induced
fragmentation, and the degree to which this occurs
increases with increasing cation size.
The effects of protons versus alkali metal cations may
be rationalized based on differences in the coordination of
the glycosidic oxygen. As shown in Figure 2, protonation is
likely to be localized to the glycosidic oxygen—the most
basic in the structure (Cancilla et al., 1996). Metal ions, on
the other hand, can undergo coordination with several
atoms simultaneously, resulting in less destabilization of
the glycosidic bond. This result is consistent with a higher
barrier to glycosidic-bond fragmentation for the alkali
cationized ions.
In addition to the glycosidic oxygen atom, acidic
residues coordinate alkali metal cations, and this factor has
&
been investigated as a means for stabilization and linkage
determination (Penn, Cancilla, & Lebrilla, 2000). CID
product-ion mass spectra of sialylated milk oligosaccharides were acquired on [M(X) þ X]þ ions, where X ¼ Li,
Na, K, or Cs, and showed that sialic acid residues are able to
bind two cesium cations but only one lithium or sodium
cation. It appears that small cations, with a high charge
density, repel each other, making binding of two cations
energetically unfavorable.
Alkali-metal cationization and subsequent positiveion CID MS results in three possible fragmentation pathways, as shown in Figure 3 (Cancilla et al., 1999). The
complex may dissociate to lose the metal cation (path A),
leaving the oligosaccharides as a neutral, and producing no
detailed information. The complex may undergo glycosidic bond cleavage (path B) with cation retention by one of
the fragments. The complex may also undergo cross-ring
cleavage (path C). The likelihood of dissociation of the
cation increases with size, and is greatest for Csþ and least
for Liþ and Naþ. Activation energies for glycosidic-bond
cleavage have been determined to be lowest for Liþcoordinated oligosaccharides. As a result, use of this cation
maximizes glycosidic bond fragmentation. The low energy
barrier results in the observation of abundant fragment ions
by MALDI MS, particularly for FT analyzers for which the
ion lifetime is long. Glycosidic-bond cleavages are believed to be charge-induced, and Liþ appears to mimic
protons by having high charge density and associating
with glycosidic oxygen atoms. The use of Csþ is best for
producing MALDI ions stable enough to be detected in
instruments, such as the FTMS, with long ion lifetimes.
That cation is large and interacts with several sites on the
oligosaccharide to minimize the destabilization of the glycosidic bond (Tseng et al., 1997). Activation energy barriers have been studied with semiquantative methods, and
have been found to be independent of alkali metal ions.
Those cleavages appear to be charge-remote, occurring
some distance from the charged site. Glycosidic-bond
FIGURE 2. Fragmentation of (a) protonated and (b) alkali-cationized glycosidic bonds (Modified from
Cancilla et al., 1996).
167
&
ZAIA
FIGURE 3. Three possible fragmentation pathways for metal-catio-
nized oligosaccharides. Reprinted with permission of Cancilla et al.
(1999). Copyright 1999 American Chemical Society.
cleavages appear to be favored under conditions where a
cation is associated with a glycosidic oxygen atom. Crossring cleavages occur under conditions where the cation
binds simultaneously to several oxygen atoms and the degree of destabilization of glycosidic linkages is minimized.
Analysis of protonated and natriated N-linked oligosaccharides by ESI Q-oTOF MS resulted in the conclusion
that the latter produces the more-informative product-ion
profiles (Harvey, 2000a). Cross-ring cleavages were observed to the reducing terminal GlcNAc residue and, at low
abundances, to antenna residues. The abundance of crossring cleavage ions diminished with increasing number of
HexNAc residues in the antennae; that result is consistent
with the idea that cross-ring cleavages are most abundant in
the absence of labile glycosidic bonds.
During CID of an oligosaccharide molecular ion, the
transfer of energy to rotational and vibrational modes
competes with glycosidic-bond cleavage (Mendonca et al.,
2003). The degree to which peralkylated oligosaccharides
undergo glycosidic bond cleavage increases with the size
of the alkyl group; i.e., methyl<ethyl<propyl<butyl<
pentyl. The larger alkyl groups are less free to rotate about
the glycosidic bonds, and therefore dissociate at a higher
rate to relieve steric and torsional strain. Certain rotational
states are inaccessible for large alkyl groups, so that transition states that lead to dissociation are ordered and give rise
to reduced energies of activation. The smaller derivatives
will have a less-ordered transition state and thus higher
entropy of activation. The anomeric configuration of (1–4)
linked disaccharides can also be distinguished by the CID
of peralkylated protonated ions (Mendonca et al., 2003).
Product ion-to-parent ion ratios are higher for the b-isomer
(cellobiose) than the a-isomer (maltose) for all alkyl
groups. The ratios reflect the differences in crowding
168
between a- and b-glycosidic linkages. The degree of
discrimination is highest for the largest alkyl group tested
(pentyl), reflecting the greater degree of conformational
restriction. The anomeric differentiation in (1–6)-linked
peralkyl derivatives is not observed due to the high conformational flexibility of this linkage.
Unambiguous determination of linkages of increasingly large oligosaccharides requires the coordination with
relatively more equivalents of metal ions, as shown with
ESI (Fura & Leary, 1993). In addition to the amount, the
choice of metal is important for large oligosaccharides.
Alkaline earth elements (Mg2þ and Ca2þ) should, when
coordinated with oligosaccharides, produce similar product-ion profiles as dilithiated compounds. Lithium has a
high charge-to-radius ratio and is similar to Mg2þ in that
both ions have a small radius and form stable compounds
with small anions. Calcium ions, with a larger radius, are
expected to coordinate more efficiently with branched
oligosaccharides. Note that [M þ Na]þ and [M þ K]þ ions
are isobaric with [M þ Mg H]þ and [M þ Ca–H]þ ions,
respectively. Thus, the use of isotopically pure 26Mg and
44
Ca may be necessary to establish that an observed adduct
is due to the divalent cation rather than to sodium or potassium. As the concentration of oligosaccharide decreases,
the optimal metal-to-oligosaccharide ratio increases. For
example, at 1.5 nmol/mL, the optimum ratio is 1:1, but at
12 pmol/mL, the ratio is 10:1 (Fura & Leary, 1993). Results
of the CID of coordinated trisaccharides were much clearer
for Ca2þ than for Mg2þ. Here, Ca2þ formed only a few
well-defined product ions, whereas the latter produced
complex patterns due to the presence of competing disaccharide pathways. The radius of Mg2þ (0.65 Å) is much
smaller than that of Ca2þ (0.99 Å) and is thus expected to
coordinate many more sites on the trisaccharide molecule.
Binding of Ca2þ is more selective in its sites of coordination, and thus produces simpler product-ion patterns.
Coordination with Mg2þ results in more abundant
cross-ring cleavages than do two equivalents of Liþ due to
the increased ability of the divalent ion to polarize and bind
tightly. Tighter binding leads to more cross-ring cleavages.
With larger oligosaccharides, Liþ induces product-ion patterns that are similar to those by Hþ, giving rise to only
glycosidic bond cleavages (Fura & Leary, 1993).
The abundance of natriated ions is maximized at high
cone voltages (Harvey, 2000a)—conditions that also result
in in-source fragmentation. Because of this factor, the
precursor-ion pattern is not necessarily reflective of the
glycan composition. With divalent metals (Ca2þ, Mg2þ,
Co2þ, Cu2þ, Mn2þ), the dominant species formed from
chloride salts were [M þ X]2þ (Harvey, 2001). The ability
of divalent metal ions to ionize neutral oligosaccharides
follows the order Ca2þ > Mg2þ > Mn2þ > Co2þ > Cu2þ.
The metal-adducted oligosaccharides resulted in singly
and doubly cationized product ions with abundances
MASS SPECTROMETRY OF OLIGOSACCHARIDES
highest for the Ca2þ ions. Because 1þ and 2þ product ions
are present simultaneously, the CID profiles are relatively
complex. The type of fragment ions observed for N-linked
oligosaccharides are similar to those observed with
natriated ions in that the only cross-ring cleavage ions
observed in abundance were 0,2A and 2,4A of the reducingterminus GlcNAc residue. The primary advantage of
divalent cations, the high relative abundance of the adducted ion, particularly for Ca2þ, is offset by the increased
complexity of the CID spectra. At the time of this writing,
Co2þ (Sible, Brimmer, & Leary, 1997) and Ca2þ (Harvey,
2001) are the most promising divalent metals for CID of
native oligosaccharides.
D. Tandem MS of Permethylated and
Peracetylated Oligosaccharides
The conversion of glycans to hydrophobic derivatives
enhances their signal strengths regardless of the ionization
technique used. The two most common derivatization
procedures involve peracetylation (Bourne et al., 1949) and
permethylation (Ciucanu & Kerek, 1984); both can be
carried out in high yield (Dell, 1990).
Permethylation has come to be the preferred method
because it results in a smaller mass increase and a greater
volatility (McNeil et al., 1982; Dell et al., 1983a,b).
Peracetylated derivatives may be detected in the positive
mode as [M þ Na]þ ions; CID fragmentation results in
preferential cleavages to the reducing side of HexNAc
residues (Dell et al., 1983). As with permethylated derivatives, [M þ H]þ ions from peracetylated oligosaccharides
fragment to produce little multiple-bond fragmentation
(Domon, Müller, & Richter, 1990). This result is an important advantage because the m/z values of such ions are,
in many cases, not distinguishable from those of primary
fragment ions generated from underivatized oligosaccharides. Therefore, internal fragment ions produce unique
mass values in permethylated and peracetylated oligosaccharide product-ion mass spectra. As a consequence, linear
and branched oligosaccharides are easily distinguished
for peracetylated [M þ H]þ ions despite the lack of crossring cleavage ions (Domon, Müller, & Richter, 1990).
Permethylation is often preferred over peracetylation; for
example, for the oligosaccharides with free hexosamine
amino groups that characterize the glycan core of glycoinositol phospholipid anchor. Those residues will be
rendered indistinguishable from those that are acetylated
if they are peracetylated. Permethylation produces a
quaternary ammonium cation from the free amine groups
that is differentiated by mass from the N-acetylhexosamine
residues, and also gives rise to unique CID fragmentation
(Baldwin et al., 1990).
Permethylated oligosaccharides are most often ionized
as [M þ Na]þ ions and produce reducing and non-reducing
&
terminal product ions with approximately equal abundances and preferential fragmentation to the reducing side
of HexNAc residues (Egge, Dell, & Von Nicolai, 1983).
Fragmentation of the substituents linked to the 3-position
of HexNAc residues is also preferred (Egge & PeterKatalinc, 1987). In general, ions that correspond to Neu5Ac
oxonium ions are abundant in fragmentation spectra of
sialylated oligosaccharides and glycoconjugates, in some
cases complicating the task of locating the position of
substitution because ions with glycosidic bonds to those
residues are in low abundance.
Fragmentation of [M þ H]þ generated from permethylated oligosaccharides also produces useful CID production mass spectra. Cleavage to the reducing side of HexNAc
is facile, and subsequent stages of MS are quite useful
for differentiating isomers. The protonated molecule ions
produce exclusively glycosidic-bond cleavage with lowenergy CID on a triple quadrupole instrument (Viseux, de
Hoffmann, & Domon, 1997). Although multiple bond fragmentation is observed, permethylation allows the ions to be
unambiguously identified on the basis of the exposed free
hydroxyl groups. HexNAc residues substituted at position3 undergo a specific elimination that identifies that type of
interglycosidic linkage.
Tandem mass spectrometric analysis with MSn of
permethylated N-linked oligosaccharides has been facilitated by removal of labile groups. The most labile groups
are Neu5Ac and HexNAc residues for typical oligosaccharides, and subsequent stages of MS proved information
on the non-reducing terminus and core structure that is not
obtained in the MS2 profile (Weiskopf, Vouros, & Harvey,
1997, 1998; Reinhold & Sheeley, 1998; Sheeley &
Reinhold, 1998; Viseux, de Hoffmann, & Domon, 1998).
Exposure of core structures produces a fragment-ion pattern upon subsequent CID that appears to be independent
of the history of formation of the ion. Figure 4 compares
product ions of m/z 1143.6, (GlcNAc)2(Man)3, generated
by MS4 of a sialylated biantennary structure and by MS3 of
asialo biantennary structure NA2 (the specific fragmentation routes are shown in the figure). The structure of the
non-reducing antennae may also be elucidated by MSn,
producing linkage information based on the presence of
cross-ring cleavages, and/or comparison to reference mass
spectra (Reinhold & Sheeley, 1998; Sheeley & Reinhold,
1998; Weiskopf, Vouros, & Harvey, 1998).
The protonated ions fragment at energies < natriated
ions, and the product-ion patterns contain different features. Permethylated oligosaccharides may be desalted
with reversed phase chromatography, resulting in their
ionization as protonated species (Laine et al., 1988, 1991;
Viseux, de Hoffmann, & Domon, 1997, 1998). For example, Figure 5 shows (a) the ESI mass spectrum of milk
oligosaccharide lacto-N-tetraose [Gal(b1-3)GlcNAc(b13)Gal(b1-4)Glc), LNT], (b) MS2 of the protonated ion,
169
&
ZAIA
FIGURE 4. MSn fingerprinting of N-linked oligosaccharide core for facile determination of sugar
branching and substitution: (a) MS4 of the glycan core of the disialo, biantennary human transferrin glycan,
m/z 1143.5; (b) MS3 of the glycans core of the asialo, galactosylated biantennary oligosaccharide (NA2)
from human fibrinogen, m/z 1143.7. Reprinted with permission from Weiskopf, Vouros, & Harvey (1998).
Copyright 1998 American Chemical Society.
and (c) MS2 of the natriated ion. The protonated ion
fragments to produce fragments that contain the nonreducing terminus. The natriated ion, on the other hand,
fragments to produce abundant Bm and Yn ions. The protonated ions undergo eliminative losses of substituents to
the 3-position of GlcNAc to produce an Em ion. Those ions
result from facile fragmentation of corresponding Bm ions,
as shown in Figure 6, and are useful to identify the mass of
the substituent on the 3-position. If the 3-position is
unsubstituted, then the loss of a methoxy group is observed.
If a monosaccharide residue is located on the 3-position,
then the Em ion will reflect that loss. The Em ions were not
observed for natriated ions. Fragmentation of protonated
permethylated oligosaccharides, therefore, appears to
have significant advantages over the natriated variants, by
virtue of the specific losses that indicate the pattern of
substitution of HexNAc residues. Unfortunately, those
advantages are offset by other problems, such as internalresidue rearrangements (see Section II.I).
An LC/MS approach has been developed for the online analysis of permethylated oligosaccharides with an ion
trap (Delaney & Vouros, 2001). The released oligosaccharides were reductively aminated with 2-aminobenzamide
to facilitate UV detection and permethylated. Ions were
subjected to data-dependent MS2 as they eluted from a
reversed phase HPLC column as sodium adducts. This
170
analysis entails the software-controlled detection of ions
followed by isolation and CID of those ions that meet
abundance or other exclusion criteria. The most abundant
product ion was automatically selected for MS3. The
HPLC-separation step carries the advantage that isomeric
structures are likely to be separated to some degree. As in
other QIT MSn analyses, the abundance of cross-ring
cleavages are quite low. The data are, therefore, most useful
as mass fingerprints, and the HPLC separation should,
therefore, facilitate the building of a library of reference
structures in a manner similar to that published (Viseux, de
Hoffmann, & Domon, 1998; Tseng, Hedrick, & Lebrilla,
1999).
E. Tandem MS of Reductively
Aminated Carbohydrates
The reducing-terminus aldehyde group of oligosaccharides
is easily reacted with alkylamines. As shown in Figure 7,
the amine forms a Schiff base with the aldehyde that cyclizes to form a glycosylamine. The Schiff base is usually
reduced to form a secondary amine because this amine is
more stable than a glycosylamine. Such a reductive
amination of glycans provides several benefits for MS
analysis. The derivative usually contains a chromophore to
enhance its chromatographic detectability. Derivatives that
MASS SPECTROMETRY OF OLIGOSACCHARIDES
&
FIGURE 6. Fragmentation of permethylated N-acetylhexosamine-con-
taining oligosaccharide B ions, glycosylated in the 3-position (a) and in
the 4-position (b). Reprinted with permission from Viseux, de Hoffmann,
& Domon (1997). Copyright 1997 American Chemical Society.
FIGURE 5. ESMS spectrum of permethylated LNT (a), and ESMS/MS
spectra of protonated (b), and natriated (c) molecular species measured at
collision offset voltages of 15 and 40 V, respectively. Reprinted with
permission from Viseux, de Hoffmann, & Domon (1997). Copyright
1997 American Chemical Society.
increase the hydrophobicity of the oligosaccharide will
typically increase signal intensity obtained from any
ionization technique (Wang et al., 1984; Carr et al., 1985;
Gillece-Castro & Burlingame, 1990; Poulter & Burlingame, 1990; Harvey, 2000b).
A number of amine groups have been used for
reductive amination in conjunction with ESI CID MS,
and their structures are shown in Figure 8 (Harvey, 2000b).
The choice of the amine influences the ionization
efficiency, and the effects are different with MALDI versus
ESI, as shown in Figure 9. Although the more-hydrophobic
derivatives tend to produce better ionization responses for
both ionization modes, the 2-aminoacridone derivative
produces a remarkable difference in response between the
two ionization techniques. That derivative is widely used in
fluorophore-assisted carbohydrate electrophoresis, and has
been used in mass spectrometric studies (Okafo et al.,
1997). It should be noted that the ESI measurements in
Figure 9 were made at flow rates of 50–200 nL/min, where
the droplet size is closer to high flow rate rather than nano
flow. It is, therefore, possible that the less hydrophobic
derivatives will produce stronger ion signals with nanospray, just as was observed with underivatized oligosaccharides (Bahr et al., 1997; Karas, Bahr, & Dulcks, 2000),
resulting from the increased surface activity from smaller
droplet size. The nature of the reductive amination derivative was found to have comparatively little effect on the
CID profiles of natriated ions, indicating that the cation
associates with the monosaccharide residues rather than
with the derivative (Harvey, 2000b). The profiles were also
similar to those of natriated native oligosaccharides. Crossring cleavages were present in low relative abundances,
requiring extended summation times to achieve an adequate signal-to-noise ratio for proper identification.
F. Discrimination of Monosaccharide Linkages
High-energy CID of cationized permethylated oligosaccharides furnishes information that identifies (2–3) versus
(2–6) linked Neu5Ac residues, as well as (1–3) versus (1–
4) linked hexose on a HexNAc residue (Lemoine et al.,
1991). Discrimination between Neu5Ac linkages is based
on the different affinities for sodium cations that gives rise
171
&
ZAIA
FIGURE 7. Reductive amination of oligosaccharides.
to dramatically different product-ion patterns. Discrimination between hexose linkages on HexNAc residues is based
on the facile elimination of the (1–3) linked group (Laine
et al., 1988; Laine, 1989). The abundance of structurally
useful cross-ring cleavage ions have been shown to depend
on the collision energy and the collision gas (Lemoine
et al., 1993).
The low-energy CID of the [M þ Na]þ ion generated
from permethylated oligosaccharides produces primarily
single-bond glycosidic-bond cleavage. At high energy
(4 keV), double-bond cross-ring cleavages predominated
in the product-ion mass spectrum, with single-bond glycosidic bond cleavages in low relative abundances. The use
of argon rather than helium as the collision gas also results
in more efficient fragmentation of the precursor-ion, and in
an enhancement of the abundance of cross-ring cleavage
FIGURE 8. Structures of common amines used with reductive amina-
tion. Reprinted by permission of Elsevier from Harvey (2000b).
172
ions. The results are consistent with the conclusion that the
cross-ring cleavage ions arise through charge-remote
fragmentation processes (Adams, 1990). Cationization at
different glycosidic linkages may give rise to a set of
charge-remote fragmentation reactions, the sum of which
is the product-ion spectrum acquired at high collision
energy (Lemoine et al., 1993).
The observations of characteristic peaks that differentiate linkage were originally made by several groups, as
summarized below.
1. Low-energy FAB CID mass spectra acquired for
[M þ H]þ ions generated from three trisaccharides,
Fuc(a1-X)GlcNAc(b1-3)Gal(b1-O-methyl),
where
X ¼ 3, 4, or 6, clearly show differences in the
abundance for Y2 ions at m/z 398 and B2 at m/z 350.
The resulting m/z 350/398 ratio indicates the Fuc
linkage position in the order of (1–6) > (1–4) > (1–3)
(Laine et al., 1988).
2. Direct stereochemical assignment of monosaccharide
units was made with low-energy CID of the glycosylated antibiotic papulacandin (Müller et al., 1988).
FIGURE 9. Comparison of MALDI (solid bars) and ESI (shaded bars)
signal strengths from the high mannose N-linked glycan (GlcNAc)2
(Man)5 derivatized by reductive amination with each of eight different
amines. Reprinted by permission of Elsevier from Harvey (2000b).
MASS SPECTROMETRY OF OLIGOSACCHARIDES
FIGURE 10. Structures of peracetylated monosaccharide oxonium
ions, m/z 331.
Chemical ionization was used to produce [M þ H]þ
ions for the acetylated molecule. The peracetylated
monosaccharide oxonium ion (see Fig. 10 for structures) was generated in the source, and was subjected to
isolation and CID fragmentation. The CID profile of
that ion was compared to the oxonium ion at the same
&
m/z value, generated from acetylated methyl glycosides
of Gal, Glc, and Man, respectively. The spectral pattern
for the acetylated Gal oxonium ion is easily distinguished from those of Glc and Man, based on the high
abundance of the product ion at m/z 127. The profiles of
acetylated Glc and Man are not distinguishable.
3. Low-energy CID product-ion mass spectra generated
for peracetylated disaccharide oxonium ions generated
from Glc(b1-X)Glc, where X ¼ 2, 3, 4, and 6 (Domon,
Müller, & Richter, 1989), show distinct patterns of
abundances, for product ions at m/z 109, 169, 211, 271,
331, 457, 559, and 619. In an attempt to make objective
distinctions among these spectral fingerprints, the
authors use similarity values that reflect vectors in ndimensional space, where n is the number of m/z
values.
FIGURE 11. Structures of peracetylated lacto-N-fucopentaoses, LNF-I, LNF-II, and LNF-III,
[M þ H]þ ¼ m/z 1484. The high-energy CID product ions diagnostic of each isomer are shown.
173
&
ZAIA
4. Low-energy CID of [M þ H]þ ions clearly distinguishes peracetylated isomeric lacto-N-fucopentaoses
(LNF, see Fig. 11 for structures) (Domon, Müller, &
Richter, 1990). The CID profile of the linear LNF-I
structure is distinguished from the branched structures
by the presence of a B2 ion at m/z 561. The branched
structures are distinguished from each other by the
presence of En ions (see Section II.D) at m/z 500 (from
LNF-II) and 558 (from LNF-III), formed by secondary
fragmentation of the branched B2 ion to lose the
substituent on the 3-position of the GlcNAc residue.
5. Distinct product-ion profiles were obtained from
isomeric glucose disaccharides with negative ion B/E
linked-scan product-ion MS (Garozzo et al., 1990).
Specifically, although the (b1–6) linkage is characterized by abundant product ions at m/z 251 (90), 281
(60), 311 (30) and 323 (18), only m/z 281 is
abundant in the profile of the (b1–4) linkage (the loss
from the precursor ion is indicated in parentheses). The
(b1–3) linkage displays abundant ions at m/z 161
(180) and 179 (162) that are absent or of very low
abundance for the other isomers. The (b1–2) linkage is
characterized by an abundant ion at m/z 263 (78).
Those losses can be used to determine linkages in the
production profiles generated from oligosaccharides.
6. Infrared MALDI desorption of isomeric disaccharides
results in distinctive fragmentation to the reducing ring
that is useful in assigning linkage (Spengler, Dolce, &
Cotter, 1990). As observed in the negative mode,
above, losses are observed that characterize each
linkage.
7. As mentioned in Section II.C.3, low-energy CID MS
of [M(Li) þ Li]þ ions generated from Glc(b1-X)Glc,
where X ¼ 4, 6, results in distinct product-ion patterns
from cross-ring cleavages (Zhou, Ogden, & Leary,
1990). Specifically, the (1–6) linked disaccharide
fragments to produce ions at m/z 235, 265, and 295
from cross-ring cleavage of the reducing-end residue.
The m/z 265 residue is absent in the profile of the (1–4)
linked disaccharide. It was also shown that linkage
patterns in linear glucose tetramers could be differ-
entiated based on those characteristic cross-ring
cleavage processes.
8. The formation of cross-ring cleavages to the reducingend residue with low-energy CID are dependent on
linkage (Hofmeister, Zhou, & Leary, 1991). The
disaccharide linkage isomers gentiobiose (b1,6), laminarbiose (b1,3), and sophorose (b1,2) are clearly
differentiated based on characteristic patterns of the
abundance of product ions at m/z 331 (–H2O), 289 (–
C2H4O2), 259 (–C3H6O3), 229 (–C4H8O4), 187
(–C6H10O5), and 169 (–C6H12O6), as indicated in
Figure 12.
9. High-energy CID product-ion mass spectra of sodiumducted methylated oligosaccharides has been shown
to discriminate between (2–3) and (2–6) linkages
of Neu5Ac (Lemoine et al., 1991), see Section V.D.
High-energy CID of permethylated Neu5Ac(a2-3)
Gal(b1–3)GlcNAc(b1–3) Gal(b1–4)Glc and Neu5Ac
(a2–6)Gal(b1–4)GlcNAc(b1–3)Gal(b1–4)Glc, respectively, showed differences in product-ion pattern
that is attributable to Neu5Ac linkage. Specifically, the
(2–3) linked isomer fragmented to form abundant ions
from cleavage of the glycosidic bond adjacent to the
Neu5Ac residue, corresponding to C1 (m/z 356) and Z4
(m/z 864). Both ions are absent in the spectrum of the
Neu5Ac (2–6) linked isomer.
Comparison of low- and high-energy CID of protonated
ions generated from glycoalkaloids indicates that production patterns are strongly dependent on the collision energy
(Claeys et al., 1996). The abundance of the ions that result
from Yn glycosidic bond cleavages are high with lowenergy CID, whereas those that result from cross-ring
cleavages, primarily 1,5Xn, are high with high-energy CID.
The high-energy mass spectra are more complex than those
acquired at low energy, but contain the structurally
valuable cross-ring cleavages. High-energy CID occurs
on a much shorter time-scale than low-energy CID, and is
much more likely to produce cross-ring cleavages from
a charge-remote mechanism. By contrast, low-energy
CID occurs on a longer time-scale, and involves multiple
FIGURE 12. Fragmentation of the [M þ Li]þ ion generated from gentiobiose, (a) cross-ring cleavages, (b)
glycosidic bond cleavages (Modified from Hofmeister, Zhou, & Leary, 1991).
174
MASS SPECTROMETRY OF OLIGOSACCHARIDES
collisions with gas molecules, resulting in cleavage only to
the most labile bonds. The question remains regarding how
best to produce cross-ring cleavages while minimizing the
extent to which multiple bond fragmentation occurs. Lowenergy CID instruments are by far the most readily available at the time of this writing.
High- and low-energy CID of high-mannose N-linked
oligosaccharides are compared in Figures 13 and 14, respectively. The high-energy CID data show nearly complete series of glycosidic-bond and cross-ring cleavages,
including those produced from cleavage of the branched
mannose residues (A4 and X3 ions). The most abundant
cross-ring cleavage ions observed in the low-energy CID
mass spectra of high mannose N-linked oligosaccharides
(Fig. 14) are those that occur to the reducing terminal
GlcNAc residue. Although other An ions are observed, the
abundances are significantly lower than the cross-ring
cleavage ions in the high-energy CID product ion mass
spectra. Those results demonstrate that high-energy CID
yields significantly more informative mass spectra than
low-energy data. At this time, however, there are no commercially available instruments that can practically accomplish high-energy CID on biologically relevant sample
quantities (low picomol quantities). Magnetic-sector
instruments offer high-energy CID, but scan far too slowly
&
to be of practical use. MALDI TOF/TOF instruments are
available, and theoretically produce high-energy CID.
Low-energy product ions are often observed, however, due
to metastable fragmentation (Yergey et al., 2002). The
development of high-pressure MALDI sources may allow
high-energy CID data to be obtained for oligosaccharides.
For that purpose, it would also be desirable to develop an
ESI source for the TOF/TOF analyzer.
Although protonated oligosaccharides fragment at lower
energies, natriated ions produce a higher yield of cross-ring
cleavage ions (Orlando, Bush, & Fenselau, 1990). Highenergy CID mass spectra were acquired on natriated ions
generated from MALDI, and were analyzed with a hybrid
magnetic sector orthogonal injection time-of-flight analyzer (Bordoli et al., 1994; Bateman et al., 1995). The
majority of the product ions observed in the tandem mass
spectra resulted from remote-site fragmentation. Highmannose N-linked sugars (Man9) produce a complete
series of Bm, Yn, and 1,5Xn ions, resulting in a highly
complex product-ion pattern (Harvey, Bateman, & Green,
1997). The abundances of multiple bond fragment ions are
low, and it appears that the product-ion mass spectra are
valuable. Because the spectra were acquired on native
molecules, the identities of some ions cannot be determined due to the possibilities of multiple bond fragmenta-
FIGURE 13. High-energy CID mass spectrum of the high-mannose N-linked sugar Man9GlcNAc2. The ion
labeled Man4 (m/z 671.3) is an internal fragment ion. Reprinted with permission from Harvey, Bateman, &
Green (1997)). Copyright 1997 John Wiley & Sons Limited.
175
&
ZAIA
FIGURE 14. Low-energy CID of high-mannose N-linked oligosaccharides. Reprinted with permission
from Harvey (2000a). Copyright 1997 John Wiley & Sons Limited.
tion. Natriated native complex N-linked oligosaccharides
produce abundant Bm ions from cleavage to the reducing
side of HexNAc residues, especially near the non-reducing
termini. Generally, the appearances of those mass spectra
were indicative of the oligosaccharide structures as fingerprints, but were not amenable to direct interpretation.
Although the abundances of cross-ring cleavage ions
in the MS2 stage are typically low with low-energy CID,
the use of higher-order MSn experiments enables more
detailed information to be produced on oligosaccharides.
The MS2 stage often serves to remove the most labile
saccharide substituents from complex oligosaccharide ions
to allow the selection and subsequent fragmentation of
less-labile portions of the molecule. Cross-ring cleavages
are in greater abundances in those subsequent MS stages,
allowing more detailed information on the nature of
underlying glycosidic linkages to be obtained (Weiskopf,
Vouros, & Harvey, 1998).
For complex N-linked oligosaccharides, the removal of
antennae and labile Fuc and Neu5Ac residues permits the
identification of the core pentasaccharide structure. Thus,
multiple MS stages may be used to expose a common core
structure exposed from different oligosaccharide molecules, resulting in identical product-ion patterns (Viseux,
de Hoffmann, & Domon, 1998; Weiskopf, Vouros, &
Harvey, 1998). Because of these observations of substructural fingerprints, several groups have proposed to use
multistage MS to virtually degrade complex oligosaccharides to common core structures; the spectra could be
176
searched against a structural library (Viseux, de Hoffmann,
& Domon, 1998; Tseng, Hedrick, & Lebrilla, 1999) or
interpreted with a computer program (Gaucher, Morrow, &
Leary, 2000; Ethier et al., 2002).
Studies on lithium- and sodium-cationized disaccharides have demonstrated conclusively that low-energy
CID results in glycosidic-bond and cross-ring cleavages
(Hofmeister, Zhou, & Leary, 1991; Asam & Glish, 1997).
Patterns of cross-ring cleavage ions are linkage-dependent,
as shown for lithium-cationized (1–2), (1–3), (1–4), and
(1–6) linked glucose disaccharides, as shown in Table 2
(Asam & Glish, 1997). The linkages in glucose oligomers
could also be determined, as shown in Figure 15 for
MS2 and MS3 data collected on lithium-cationized trimers
of isomaltose [Glc(a1–6)Glc(a1–6)Glc] and panose
[Glc(a1–6)Glc(a1–4)Glc] (Asam & Glish, 1997). The
MS2 data (panels a and c) show product-ion patterns at m/z
229, 259, and 289 that are consistent with (1–6) linkages,
referring to Table 2, for both isomers. The non-reducing
end linkage is indicated by ions at m/z 392, 422, and 452
that correspond to the same cross-ring cleavages with a
hexose residue attached.
Multistage MS of cobalt-coordinated milk oligosaccharides has been used to determine precise linkage positions
(Konig & Leary, 1998). This determination was shown for
the lacto-N-fucopentaoses (LNFP) I, II, III, and V, that were
all of composition (Gal)2(Glc)(GlcNAc)Fuc. During lowenergy CID, the metal-coordinated oligosaccharide ion
begins to dissociate with specific cross-ring cleavages from
MASS SPECTROMETRY OF OLIGOSACCHARIDES
&
TABLE 2. Lithium-cationized disaccharide MS/MS product ions and abundances [reprinted by permission of
Elsevier from Asam & Glish (1997)]
a
Intensity relative to base peak.
Cross-ring cleavage ion types in Domon–Costello nomenclature.
c
Galactosyl(1 ! 4)glucose, all others glucosyl(1 ! X)glucose.
b
FIGURE 15. (a) MS/MS spectrum of lithium-cationized isomaltotriose, (b) MS3 spectrum of the C2 ion of
lithium-cationized isomaltotriose, (c) MS/MS spectrum of lithium-cationized panose, and (d) MS3 spectrum
of the C2 ion of lithium-cationized panose. Note that the circular symbols in this figure depict MSn
transitions, and do not refer to monosaccharide residues. Reprinted by permission of Elsevier from Asam &
Glish (1997).
177
&
ZAIA
the reducing end, followed by non-reducing end cleavages
(Zhou, Ogden, & Leary, 1990; Hofmeister, Zhou, & Leary,
1991). Linkages are indicated by the presence of specific
neutral losses in the MS2 profiles, as shown in Table 3.
Multistage MS data were acquired on LNFP III and LNFP
V (Fig. 16), differing in the position of fucosylation and the
Gal-GlcNAc linkage (see figure for structures) (Konig &
Leary, 1998). The results are significant in that cross-ring
cleavages are produced in the MS2 stage despite the
presence of GlcNAc and Fuc residues; both are known to be
labile. For each stage of MS, cross-ring cleavages,
expressed as losses from the precursor ion, indicated the
linkage position of the reducing-terminus residue. Selection of the Cm ions in subsequent stages of MS allows other
linkages to be determined, as shown for the figure insets.
The MS5 product-ion spectrum of LNFP III corresponds to
the fragmentation of Gal(1–4)GlcNAc (Fig. 16d), in which
an ion is observed at m/z 280, which corresponds to the loss
of 161 from the precursor ion. The MS5 product-ion profile
of LNFP V corresponds to fragmentation of Gal(1–
3)GlcNAc (Fig. 16i), and features an ion at m/z 279,
produced by loss of 162 from the precursor ion. Convincing
evidence from stable isotope-labeling shows that its
formation occurs by a mechanism that is distinct from the
m/z 280 ion in Figure 16d (Konig & Leary, 1998), and the
cobalt complexes also differentiate GalGlcNAc linkages.
Although the cobalt complexes do not provide direct
evidence for the fucose linkages, dissociation of [M–H]
ions from the same solutions provided this information.
The usefulness of CID of cobalt complexes, however, is
limited by two factors: (1) the relatively high m/z of the
precursor ion, and (2) the difficulty in obtaining MS5 and
subsequent fragmentation steps to obtain the linkage
information on larger oligosaccharides. To get around the
fragmentation efficiency problem, it has been possible to
produce the necessary Cm ions in a single stage of MS2
(Konig & Leary, 1998). The MS3 experiments on each Cm
ion provided the same structural information as in the MSn
TABLE 3. MS2 results obtained from CID experiments of
[M þ Co–H]þ for the indicated disaccharides with specific neutral
losses depending on their linkage position [reprinted by permission
of Elsevier from Konig & Leary (1998)]
178
experiments shown above. That approach will be useful
to minimize sample consumption by reducing the number
of ion-isolation steps necessary to obtain the linkage
information.
Transition-metal ligand attachment has been used as a
derivatization method to improve oligosaccharide ionization efficiency (Gaucher & Leary, 1998; Desaire & Leary,
1999a,b; Gaucher, Pedersen, & Leary, 1999). Accordingly,
the synthesis of divalent metal-coordinated N-glycosides
has been shown to stabilize sialic residues and to provide
determination of linkages (Leavell & Leary, 2001). The
required N-glycosides are easily synthesized with diethylenetriamine, mixed with metal salts, and diluted prior to
ESI MS analysis (Gaucher, Pedersen, & Leary, 1999). The
structure of the complexes in which the metal is coordinated to the diethyenetriamine nitrogens at the reducing terminus is shown in Figure 17. Losses of sialic acid
were dramatically reduced by transition-metal complexation in the order of the Irving-Williams series: CoII<
NiII<CuII>ZnII. As shown in Figure 18, the pattern of
neutral losses from metal-ligated precursor ions differ for
(2–3) and (2–6) linked sialic acid linkages.
One problem with the methodology is that the metallation reaction results in several species for each trisaccharide, including protonated, natriated, and transitionmetallated ions, giving rise to a complex mass profile.
Also, the stabilization of sialic acid residues is believed to
involve deprotonation of the carboxyl group, and it is,
therefore, not clear how large a molecule will work with
this approach because non-reducing terminal Neu5Ac
residues will be distant from the derivatized reducing
terminus.
G. Gas-Phase Degradation of Oligosaccharides
Although permethylated oligosaccharides ionize readily
as sodium adducts, protonated ions may be observed
by dissolving samples in ammonium acetate (Viseux, de
Hoffmann, & Domon, 1998). Alternatively, permethylated
oligosaccharides may be desalted and/or fractionated with
reversed-phase chromatography or solid-phase extraction.
Multistage tandem MS of [M þ H]þ ions generated from
two permethylated isomeric milk oligosaccharides that
differ in structure by the linkage of a single GlcNAc residue
were observed, as shown in Figure 19. The data show the
production of distinct patterns, primarily due to the E2 ion
that results from facile elimination of the substituent on the
3-position of GlcNAc residues. The LNT oligosaccharide
contains a Gal(1–3)GlcNAc linkage. Loss of the Gal residue occurs through E2 fragmentation to form an ion at m/z
228, which is not detected in Figure 19a,b due to the fact
that it is below the ion-trap detection window. That ion is
detected in the MS3 stage acquired by selecting and
MASS SPECTROMETRY OF OLIGOSACCHARIDES
&
FIGURE 16. (a–e) MSn of LNFP III; (f–j) MSn of LNF PV. Reprinted by permission of Elsevier from
Konig & Leary (1998). (Continued on next page.)
179
&
ZAIA
FIGURE 16. (Continued)
fragmenting the B2 ion, as shown in Figure 19c,d. LNnT,
containing Gal(1–4)GlcNAc, fragments to produce an
abundant E2 ion at m/z 432 from the elimination of the
methoxy group on GlcNAc position 3. Those E2 ions are
shown in MS3 product-ion mass spectra acquired by
selecting and fragmenting the B2 ions for either isomer
(lower panels). In further work, it was shown that the
FIGURE 17. Structure of the Mn(II) N-glycoside complex. Reprinted
by permission of Elsevier from Leavell & Leary (2001).
180
pattern produced from the fragmentation of the B2 ion is the
same, regardless of the history of formation of the ion.
Specifically, 964 ! B2 ! produces the same pattern as
964 ! B3 ! B2 ! for a given isomer. Therefore, mass
spectrometric fragments generated from large oligosaccharides may be subjected to CID, and the resulting product
ion mass spectra compared directly with those of known
compounds, such as those in Figure 19. Thus, it is possible
to build a series of reference spectra to allow a comparison
with unknown oligosaccharides to facilitate interpretation.
It is clear that disaccharide reference product-ion
spectra allow the differentiation of (1–3), (1–4), and (1–6)
Hex-HexNAc linkages (Viseux, de Hoffmann, & Domon,
1998). The same approach was shown to be useful for the
differentiation of fucosylation linkage patterns in milk
oligosaccharides. The results are not as clear, however, for
sialylated oligosaccharides. The MS/MS product-ion
profiles of Neu5Ac(2–3)Gal(1–3)GlcNAc and Neu5Ac
(2–6)Gal(1–3)GlcNAc were used as references for the
MASS SPECTROMETRY OF OLIGOSACCHARIDES
&
FIGURE 18. MS3 of (a) [Co(II)/dien/a(2–3)-sialyllactose–H]þ and (b) [Co(II)/dien/a(2–6)-sialyllac-
tose–H]þ. Reprinted by permission of Elsevier from Leavell & Leary (2001).
determination of sialic acid linkages of milk oligosaccharides. As shown in Table 4, it was observed that (2–6)
linked oligosaccharide LST-c produced a pattern that
matched that of 6-SLN, a (2–6) linked compound used as a
reference; however, the pattern for (2–3) linked oligosaccharide LST-a did not match the corresponding reference
3-SLN. A more reliable alternative was to hydrolyze the
permethylated sialic residues under acid conditions, and
to perdeuteromethylate the exposed hydroxyl groups. The
resulting spectra may be directly compared to those of
asialo reference compounds; the deuterium mass-shifts
indicate the position of the sialylated residues.
H. Computer-Based Approaches for Interpretation
of Oligosaccharide Product-Ion Mass Spectra
An automated computer program has been described that
calculates theoretical PSD mass spectra from different
181
&
ZAIA
FIGURE 19. ESMSn of permethylated tetrasaccharides: MS2 spectra of [M þ H]þ ions at m/z 904 of LNT
(a) and LNnT (b) (collision energy 18%); MS3 spectra of B2 fragments at m/z 464 of LNT (c) and LNnT (d)
(collision energy 15%). Substructures derived from oligosaccharides are depicted with black symbols.
Reprinted with permission from Viseux, de Hoffmann, & Domon (1998). Copyright 1998 American
Chemical Society.
TABLE 4. Intensities of fragment ions generated from permethylated sialylated
subunits [reprinted with permission from Viseux, de Hoffmann, & Domon
(1998), Copyright 1998 American Chemical Society]
182
MASS SPECTROMETRY OF OLIGOSACCHARIDES
pyridylamidated N-linked oligosaccharide structures
(Mizuno et al., 1999). The computer program takes into
account the following fragmentation processes: (1) monosaccharide residue loss from the non-reducing termini;
(2) subsequent monosaccharide residue losses from the
aforementioned ions, generating a Yn ion series; (3) the
complementary Bm ions; (4) formation of internal fragment
ions from loss of pyridylamidated residues from the Yn
ions; (5) formation of internal fragments from Bm ions; and
(6) formation of Zn, as well as the complementary Cm ions.
The program assigns the monosaccharide compositions
of the losses observed from the precursor ion in acquired
PSD mass spectra, and thereby provides a useful interpretation tool. Simulated mass spectra are calculated for
candidate structures that may be manually matched against
the observed data. The program does not appear to differentiate the PSD mass spectra calculated for branching
isomers.
A computer-based approach for interpretation of
native oligosaccharide MSn data has been published
(Gaucher, Morrow, & Leary, 2000). That program, known
as STAT, calculates all possible monosaccharide compositions for a given precursor-ion and those for observed
product-ions. The output lists all possible structures that
may fit the precursor-ion mass and product-ion pattern. The
program is useful for oligosaccharides that contain up to 10
residues, and a long list of possible structures is generated
from the input data. A scoring system ranks the structures
in terms of the degree to which they match the data, and
the correct structure was found to be high-ranked. The
number of structural possibilities may be narrowed with
methylation analysis and/or biological data that eliminate
structures because they lack certain core structures. The
algorithm uses m/z values, and does not yet incorporate
abundance information. By incorporating abundances,
it may be possible to expand the discrimination of the
program.
Recent work has built on the STATalgorithm as applied
to the interpretation of product-ion mass spectra of Nlinked oligosaccharides (Ethier et al., 2002). In that case,
sugars are derivatized with phenyl-3-methyl-5-pyrazolone
(PMP) (Honda et al., 1989; Saba et al., 1999), and tandem
mass spectra were acquired with a Q-oTOF instrument.
The product-ion patterns of protonated precursor ions
generated by MALDI were dominated by Yn ions (Saba
et al., 1999).
I. Internal Residue Loss Rearrangements
of Oligosaccharide Ions During CID
High-energy CID of protonated native oligosaccharides
that contain (1–2) linkages showed an unexpected loss of
162 u from the precursor-ion that could only be explained
by loss of an internal residue (Kovacik et al., 1995). The
&
internal residue-loss ion does not appear in high-energy
CID mass spectra of natriated ions or of protonated peracetylated ions. Protonated native and protonated permethylated oligosaccharides of (1–6) Gal oligomers have
both been observed to produce abundant internal-residue
losses from certain Yn ions (Brüll et al., 1997). Although
a- and b-linked internal residues may both be lost,
peracetylated derivatives did not undergo the rearrangement. In further work, protonated, natriated, and deprotonated ions produced from native and permethylated
oligosaccharides were analyzed by high- and low-energy
CID MS/MS (Brüll et al., 1998). Because natriated and
deprotonated ions both failed to undergo internal residue
losses in either high- or low-energy modes of fragmentation, it appears that protons are necessary to trigger the
rearrangements. The rearrangements occur for a number of
linkages and monosaccharide compositions, including
HexNAc-containing oligosaccharides.
Protonated native sialyl Lewis tetrasaccharide alkyl
glycosides have been observed during low-energy CID to
undergo internal-residue rearrangements (also described as
sequence isomerization) through a migration of fucose
residues toward the non-reducing terminal sialic acid
residue (Ernst, Müller, & Richter, 1997). Those conclusions are based on the observation of fucosylated Bm ions
that are not expected, based on the well-characterized
known structures. The proposed mechanism for fucose
migration is shown in Figure 20. Analysis of the highest Bm
ion (B3 for sialyl Lewis structures) with MS3 resulted in
much lower abundances of rearrangement ions. Mass
spectrometric selection of a Bm ions corresponds to the
removal of the aglycon alkyl group, and seems to favor
better charge fixation, reduced proton mobility, and reduced rearrangement.
Protonated ions produced from reductively aminated
oligosaccharides undergo intramolecular fucose rearrangements towards the derivative (Harvey et al., 2000). The
rearrangement is seen in [M þ H]þ and [M þ 2H]2þ ions,
but is absent in [M þ Na]þ precursor ions. The fact that
rearrangement is only observed for protonated ions is
consistent with a charge-induced mechanism, and a proposal that involves protonation of the reducing terminal
derivative is shown in Figure 21.
The presence of rearrangements that isomerize the
sequences in protonated oligosaccharide tandem mass
spectra may lead to erroneous conclusions regarding the
presence of more than one isomer in a mixture. The use of
alkali or other metals is therefore advised when analyzing
unknown oligosaccharides structures. It may be feasible
to select appropriate ions, such as Bm, to minimize the
extent of rearrangement (Ernst, Müller, & Richter, 1997),
but this selection requires further investigation. Because
lithiated oligosaccharides produce product-ion patterns
that more closely resemble those produced from a
183
&
ZAIA
FIGURE 20. Proposed migration of (incipient) fucosyl C1 carbenium ion towards sialyl residue to yielding
irregular m/z 438 and 600 ions (paths 1 and 2, respectively) via sequence-rearranged [M þ H]þ intermediate.
Reprinted with permission of Elsevier from Ernst, Müller, & Richter (1997).
protonated ions (Fura & Leary, 1993), it is not clear
whether the use of this metal will prevent internal-residue
rearrangement. Until further investigation has been made,
it seems prudent to analyze natriated oligosaccharides to
avoid rearrangements.
J. Conclusions
At the time of this writing, the most promising means to
produce cross-ring cleavages by low-energy CID MS for
carbohydrate oligomers is through the use of transition-
FIGURE 21. Proposed mechanism for fucose migration during the fragmentation of 2-AB derivatized
20 FL. Reprinted from Harvey et al. (2002). Copyright 2002 American Chemical Society.
184
MASS SPECTROMETRY OF OLIGOSACCHARIDES
metal complexes in combination with MSn. It may be
possible to add appropriate metals to HPLC mobile phases
either pre- or post-column and to detect complexes that
would facilitate sequential identification of residues by
MSn (Kohler & Leary, 1995; Konig & Leary, 1998). The
observation that metals bind preferentially to alditols
(Tseng, Hedrick, & Lebrilla, 1999) bears further investigation in that context. The extent to which metal complexation and MSn will prove valuable depends largely on the
compound class. Linear oligosaccharides such as the glycosaminoglycans and those with a low degree of branching
such as milk oligosaccharides and some O-linked sugars
are likely to produce useful results with this approach.
Highly branched N-linked sugars are likely to be too
complex to produce data that can be directly interpreted.
The value of tandem MS of N-linked sugars appears to be
in classifying the type; i.e., high-mannose, hybrid, or
complex, and the degree of branching. The spectra cannot
be interpreted de novo because of the isomeric antenna
structures and the overall complexity. It is also possible to
make conclusions about the relative degree of branching
based on the appearance of the product-ion profiles. The
data do, however, appear to serve as a fingerprint for the
N-linked structures. In that light, it would be valuable to
construct libraries of MSn data on known N-linked structures, defining those that can be differentiated by mass
spectrometric degradation. As detailed in this section, this
work has already begun. Another development that may
further the analysis of oligosaccharides is the MALDI
TOF/TOF tandem mass spectrometer. That instrument is
theoretically able to produce cross-ring cleavages, and may
thus increase the information content of oligosaccharide
mass spectra by increasing the abundances of cross-ring
cleavages. The ions must be produced without metastable
fragmentation, however, or else all of the ion current will
go to more labile glycosidic bond cleavages. Because
metastable fragmentation has been a problem with MALDI
TOF/TOF of peptides (Yergey et al., 2002), carbohydrate
mass spectrometrists are practically limited to low-energy
CID at this time.
The approaches for mass spectrometric analysis of
carbohydrates are summarized in Figure 22. If the released
glycans are available in quantities of 5 mg or greater, then
permethylation is recommended. The procedure is quite
robust, as evidenced by its wide use, and it increases the
ionization responses from ESI and MALDI. In addition,
tandem mass spectra of permethylated carbohydrates are
more informative than those of native molecules because it
is possible to identify cleaved glycosidic bonds by virtue of
the free hydroxyl group created. This factor also allows one
to identify internal fragment ions because they will not be
isobaric with single-bond cleavage ions. The permethylated carbohydrates can be analyzed directly with ESI or
MALDI, to obtain semi-quantitative responses, or fractio-
&
FIGURE 22. Analytical pathways for mass spectrometric analysis of
carbohydrates.
nated with reversed phase HPLC to limit mixture complexity. The most fruitful approach to CID is to build an
understanding of the fragmentation of a library of known
structures by acquiring MSn data. The MSn profiles of
unknowns can be compared to that library to facilitate
interpretation.
In many cases, the quantities of released oligosaccharides will be below the level that is fruitful to attempt
permethylation. This limit is primarily due to losses associated with the comparatively large solution volumes
necessary during the extraction step that eliminates permethylation as a viable option for analysis of glycans
released from SDS–PAGE gel spots or from small pieces of
tissue. For such samples it is therefore necessary to gain as
much information as possible using the native structures.
One possibility for analysis is to separate the carbohydrate
mixture into neutral and acidic pools using an anion exchange cartridge. This serves to simplify the mixtures
entering the mass spectrometer, prevent overlapping
peaks, and eliminate suppression problems. Neutral carbohydrates would be analyzed as sodium (or other metal)
cations in the positive mode using either ESI or MALDI,
and acidic oligosaccharides would be analyzed in the
negative mode using either ionization technique. A second
possibility is to methyl esterify the unseparated glycans to
allow simultaneous ionization of all sugars with semiquantitative responses (Karlsson, Karlsson, & Hansson,
1995; Powell & Harvey, 1996). Analysis of all released
carbohydrates would be accomplished at once with minimal sample handling, an advantage for SDS–PAGE gel
spots. Tandem mass spectrometric analysis of neutral
carbohydrates is best carried out from metal cationized
ions in the positive mode or from deprotonated ions. It is
not likely that CID spectra will be directly interpretable,
185
&
ZAIA
and thus investigators are advised to build a library of
known structures from their biological system, and to use
the information and experience gained to interpret data
generated from unknowns.
III. ANALYZERS FOR MASS SPECTROMETRY
OF CARBOHYDRATES
A. Analysis of Permethylated Carbohydrates Using
High Temperature GC/MS
Capillary gas chromatography (GC) is a fast and sensitive
separation technique that is easily interfaced with MS.
Molecules eluting from the GC column are subjected to
electron impact (EI) ionization and mass spectra acquired.
These spectra are characterized by abundant fragment ions
and a low abundance of an ion corresponding to the intact
molecular ion. Although GC has been widely used to
separate monosaccharide residues after degradation (Lindberg & Lonngren, 1978; Geyer & Geyer, 1994), it has also
been used in the analysis of derivatized intact oligosaccharides (Hallgren & Lundblad, 1977; Fournet et al., 1980;
Nilsson & Zopf, 1982; Nilsson & Zopf, 1983) of up to 15
monosaccharide units in size (Wang, Matsuura, & Sweeley,
1983). The combination of GC with MS provides the
significant benefit that mixture components are separated
prior to being analyzed by EI MS. Because EI spectra
result in complex fragment-ion patterns, a high-resolution
separation technique, such as GC, is essential because,
under favorable conditions, only one mixture component
enters the MS source at a given time.
Permethylation of oligosaccharides is easily performed, using the solid sodium hydroxide method (Ciucanu &
Kerek, 1984), and that derivative is the most widely studied
by high-temperature GC/MS. As with all GC/MS methods,
the analysis is limited by the volatility of the analyte
molecules. Using thin-film capillary GC columns combined with high temperatures, the analysis of permethylated
oligosaccharides up to 2500 Da has been demonstrated
(Karlsson, Carlstedt, & Hansson, 1989). The EI fragments
of permethylated oligosaccharides correspond to Bm
ions, using standard nomenclature (Domon & Costello,
1988b). Although the precursor ion is often absent, the
pattern of fragment ions is often useful for differentiation
of structural isomers. This result is in sharp contrast to
ESI and MALDI, where the intact protonated molecule
ion is usually abundant and fragmentation of permethylated oligosaccharides during the ionization process
occurs to a much lower degree. It is, therefore, not
necessary to separate mixtures prior to analysis with
these techniques, and fragment ions may be produced by
mass isolation followed by collision-induced dissociation
(CID).
186
B. Analysis of Carbohydrates
with MALDI-TOF MS
Analysis of ions that result from metastable decay in the
post-source field-free region of a reflectron time-of-flight
mass spectrometer has been used to obtain structural information on biopolymers (Tang et al., 1988; Spengler,
Kirsch, & Kaufmann, 1991). That technique entails the
sequential lowering of the ion-reflector potential to focus
the metastable fragment ions. Now known as PSD, the
phenomenon has been observed with MALDI ionization
for sialylated glycopeptides (Huberty et al., 1993) as well
as oligosaccharides (Harvey et al., 1995; Yamagaki &
Nakanishi, 2000). Because MALDI ion signals produced
from oligosaccharides are enhanced by reductive amination (Takao et al., 1996) or permethylation (Perreault et al.,
1997), the effects of those modifications on the information
of MALDI-PSD spectra have been investigated. Reductively aminated oligosaccharides modified with benzylamine or 2-aminopyridine produce a full series of fragment
ions that contain the original reducing terminus (Lemoine,
Chirat, & Domon, 1996; Okamoto et al., 1997).
An optimized calibration procedure has been presented and used for structural characterization of permethylated oligosaccharides by PSD (Viseux, Costello, &
Domon, 1999). Cleavage of a(2–3)-linked sialic acids in
sialyllactoses has been observed with PSD to occur more
readily than cleavage of the a(2–6)-linkage (Yamagaki &
Nakanishi, 1999), allowing the isomers to be distinguished
based on B1 ion abundance. PSD has been combined with
sequential exoglycosidase treatment for oligosaccharide
microsequencing (Sato et al., 2000). Product-ion patterns
indicative of (1–3) versus (1–4) linkages at the nonreducing termini of 2-aminobenzamide-derivatized tetrasaccharides were observed in this work. The observation of
PSD fragment ions has also been used to differentiate
Lewis-type linkage isomers in trisaccharides (Yamagaki &
Nakanishi, 2000). Localizing a fixed positive charge at the
reducing end of carbohydrates has been accomplished by
reductive amination with benzylamine followed by N,Ndimethylation (Broberg, Broberg, & Duus, 2000). PSD
profiles have been shown to differ for trisaccharide linkage
isomers that derive from Lewis-type oligosaccharides
(Yamagaki & Nakanishi, 2000). The resulting MALDI
PSD spectra were considerably simpler than observed in
typical PSD spectra, and showed a series of reducingterminal fragment ions. PSD experiments have, therefore,
been demonstrated to be useful in the analysis of
oligosaccharides.
As with peptides, however, PSD is not the method of
choice for fragmentation of carbohydrates. The necessity
of stitching together spectra derived from several reflector
voltages makes calibration difficult. The number of laser
shots required to obtain a coverage of the entire m/z region
MASS SPECTROMETRY OF OLIGOSACCHARIDES
at sufficient signal-to-noise ratio is also quite high. Further,
the product-ion resolution is poor relative to that generated
by CID. The recent availability of MALDI sources for
analyzers such as the FT, Q-oTOF, and QIT enable far more
efficient production and detection of product ions than is
possible with PSD.
C. Analysis of Carbohydrates with
MALDI Q-oTOF MS
The development of MALDI ionization sources for
quadrupole-orthogonal time-of-flight (QoTOF) type instruments (Morris et al., 1996; Krutchinsky et al., 1998;
Loboda et al., 2000) has improved precursor-ion selection
and substantially higher product-ion resolution than with
MALDI-PSD TOF (Harvey et al., 2000). Using MALDI
QoTOF MS, it is possible to obtain a CID MS/MS production spectrum of a sulfated N-linked sugar, and to observe
Bm and Yn ions for the entire structure (Wheeler & Harvey,
2001). Spectra of sialylated glycans are reported to be
simpler for MALDI QoTOF MS than MALDI-TOF MS
due to the absence of metastable fragmentation (Harvey
et al., 2000). The analysis of singly charged ions is seen
as an advantage relative to that of multiply charged ions
generated by ESI (Harvey, 2000a) in that the spectra are
easier to interpret. Neutral and acidic glycans released from
glycoproteins separated by gel electrophoresis have been
derivatized and subjected to reductive amination with
3-acetylamino-6-acetylaminoacridine. Fragmentation observed with MALDI QoTOF MS/MS in the positive mode
showed that fragment ions that contain the derivatized
reducing terminus were predominant, facilitating the interpretation of the spectra (Hanrahan et al., 2001).
D. Analysis of Carbohydrates with
ESI Q-oTOF MS
The development of ESI QoTOF MS (Morris et al., 1997;
Shevchenko et al., 1997) has substantially increased the
sensitivity and resolution at which carbohydrate tandem
mass spectra can be obtained. Alkali-adducted ions can be
analyzed with ESI QoTOF MS (Harvey, 2000a). The use of
divalent cations, including Mg2þ, Ca2þ, Mn2þ, Co2þ, and
Cu2þ in ESI QoTOF MS/MS, has been studied and found to
induce more abundant ions than monovalent metal ions
(Harvey, 2001). Most observed fragment ions corresponded to glycosidic bond cleavage rather than cross-ring
cleavage. A new derivative for reductive amination, N-(2diethylamino)ethyl-4-aminobenzamide (Harvey, 2000c),
has been studied and produces more abundant ions for
carbohydrates than do other aromatic-amine derivatives
(Harvey, 2000b). Analysis of sialylated sugars with
negative-ion QoTOF MS/MS has shown that fragmentation is dramatically influenced by Neu5Ac linkage position
&
(Wheeler & Harvey, 2000). In that work, it was observed
that native deprotonated oligosaccharide ions fragment to
produce an ion at m/z 306, corresponding to a 0,4Am, from
cleavage across the ring to which the sialic acid is a2-6
linked. Because that ion is absent in the CID profiles of a23 linked isomers, the presence of that ion is diagnostic for
the a2-6 linkage.
E. Analysis of Carbohydrates with QIT MS
Over the last decade, tandem-in-time instruments, in which
the multiple stages of mass spectrometry are performed in
the same space, as in the FT MS (Marshall, Wang, & Ricca,
1985) and QIT MS (Johnson & Yost, 1985; March &
Hughes, 1989; March & Londry, 1995; March, 2000) have
become more common. Those tandem-in-time instruments
have inherently greater MS/MS efficiency due to their configuration, as compared to tandem-in-space instruments,
leading to greater sensitivity and selectivity.
The QIT MS has emerged as a remarkably sensitive
and selective instrument that is small, low in cost, and
capable of efficient MS/MS (Johnson et al., 1990; Todd,
1991; March, 1998; March, 2000). On commercially
available QIT MS instruments, it is possible to accurately
isolate and fragment a precursor ion, and to selectively
fragment only that m/z value, thus minimizing the extent to
internal fragment ions are formed. Because of the small
time necessary to acquire an MSn mass spectrum, it is
possible to couple the QIT MS to liquid chromatography to
gain more specific information about a sample. Newer
capabilities of commercially available QIT MS instruments include the ability to develop methods for intelligent, data-dependent scanning. Using ESI QIT MS, it is
possible to remove non-reducing terminal substituents that
suppress cross-ring and core-carbohydrate cleavage in the
tandem mass spectra of permethylated complex N-linked
oligosaccharides (Reinhold & Sheeley, 1998; Sheeley &
Reinhold, 1998; Weiskopf, Vouros, & Harvey, 1998). The
data produced from MS3 and MS4 stages allow sequence
and linkages to be obtained. Sequential stages of MS
have been used as virtual degradative steps to reduce the
structures of complex sialylated and fucosylated oligosaccharides to subunits, the spectra of which can be matched
against known standards (Viseux, de Hoffmann, & Domon,
1998). Mass spectrometric degradation of oligosaccharides
has been combined with a library of known structures to
characterize substructural motifs expressed by families
of N-linked glycoproteins (Tseng, Hedrick, & Lebrilla,
1999). Different glycan structures contain common substructural motifs that result in common features in tandem
mass spectrometric profiles. The tandem mass spectra of
known substructural motifs constitute a catalog library,
against which the data generated from unknowns are
searched.
187
&
ZAIA
F. Analysis of Glycoconjugates with FT MS
Detailed characterization of carbohydrate and glycolipid
structures often requires multi-stage decomposition (MSn),
making the application of trapped ion type instruments
such as the QIT and FTMS advantageous. The FTMS has
the considerable advantage over the QIT that ions are
detected with high resolution and mass accuracy. The
widespread use of FTMS in the analysis of biomolecules
has been limited, however, by the lack of well-developed
software to enable the facile isolation and fragmentation of
ions for MSn experiments. It is anticipated that instruments
will be introduced in the near future that will correct this
problem, and thereby allow broader use of FTMS. FTMS is
compatible with ESI and MALDI, and may be used with
several different dissociation techniques, including sustained off-resonance irradiation (SORI) CID (Gauthier,
Trautman, & Jacobson, 1991), infrared multiphoton dissociation (IRMPD) (Little et al., 1994), and electroncapture dissociation (ECD) (Zubarev, Kelleher, & McLafferty, 1998).
SORI-CID is a general method for the collisional
activation of ions trapped in an FTMS cell, and is frequently used for structural analysis of proteins and polypeptides (Guan, Marshall, & Wahl, 1994; Hofstadler et al.,
1994; Huang et al., 1994; Senko, Speir, & McLafferty,
1994; Wu et al., 1995; Little et al., 1996; Solouki
et al., 1996; Heck & Derrick, 1997; Kelleher et al., 1998;
Kelleher et al., 1999; Maier et al., 2000), carbohydrates
(Cancilla, Penn, & Lebrilla, 1998; Solouki et al., 1998;
Cancilla et al., 1999; Gaucher et al., 2000; Penn, Cancilla,
& Lebrilla, 2000), and oligonucleotides (Hettich &
Stemmler, 1996; Flora, Hannis, & Muddiman, 2001). With
SORI-CID, the use of many low-energy collisions has the
advantage of being a slow-heating method so that ions
fragment mainly through the lowest energy fragmentation
channel.
IRMPD, another low-energy fragmentation method,
has previously been shown to be an efficient and selective
fragmentation method for sequencing of proteins (Little
et al., 1994; Dufresne, Wood, & Hendrickson, 1998) and
DNA (Little & McLafferty, 1995; Little et al., 1996).
IRMPD has been used to analyze N-linked glycopeptides
with the result that extensive glycan fragmentation is observed with no ion from cleavage of the peptide backbone
(Håkansson et al., 2001). IRMPD has also been used to
analyze sulfated oligosaccharides derived from chondroitin sulfate and to result in more extensive losses of SO3 than
observed with SORI-CID (McClellan et al., 2002).
ECD has emerged as a new odd-electron fragmentation method for FT MS (Zubarev, Kelleher, & McLafferty,
1998), based on partial recombination of multiply protonated polypeptide molecules with thermal electrons
(Kruger et al., 1999). Peptide backbone cleavage occurs at
188
the Ca –N bond rather than at the amide linkage as occurs
with classical CID or IRMPD. The exothermic electron
capture reaction induces the formation of an atomic hydrogen H., the energetic transfer of which to the backbone
carbonyl permits the formation of primarily the zn.þ and
cnþ ions (Stensballe et al., 2000). Peptide fragmentation
with ECD typically cleaves all Ca –N bonds along the
peptide backbone, often yielding a complete sequence,
depending on the signal-to-noise ratio of the resulting
fragment ions. The notable exception to that rule is that
proline, being cyclic around the Ca –N bond, does not
generate fragment ions.
ECD is also the preferred method for the analysis of
modified peptides because backbone cleavages are usually
observed (Kelleher et al., 1999; Mirgorodskaya, Roepstorff, & Zubarev, 1999; Stensballe et al., 2000; Shi et al.,
2001). The use of ECD in the analysis of carbohydrates is
starting to be appreciated. It is well-accepted that sialic acid
(Neu5Ac) and fucose (Fuc) residues are readily lost during
low-energy CID, a fact which can make identification of
the branching location difficult. It is not known at this time
the extent to which ECD will allow carbohydrate branching
patterns to be determined. ECD has proven to be useful
in the analysis of glycopeptides, as will be discussed in
Section IV.C.
G. Conclusions
In summary, any of the modern mass analyzers may be used
in the analysis of oligosaccharides. For profiling oligosaccharide mixtures, MALDI has been observed to provide
more even ionization response as mass increases than does
ESI. ESI, on the other hand, is less likely to cause metastable fragmentation than in MALDI—an important concern for the analysis of fragile sulfated and/or sialylated
oligosaccharides. The use of high-pressure MALDI sources
to minimize the extent to which metastable fragmentation
of oligosaccharide samples occurs is recommended.
Conventional low-pressure MALDI-TOF is recommended for profiling oligosaccharide mixtures, provided
that measures are taken to address the fragility and ionization potentials of acidic oligosaccharides (see Section V).
Although MALDI-TOF is not recommended for sulfated
oligosaccharides, particularly those with more than one
sulfate group, due to extensive fragmentation in the source
(see Section VI.B.2.b), high-pressure MALDI has been
observed to be significantly gentler and is applicable to
sialylated glycoconjugates. The use of MALDI-QoTOF for
tandem MS of oligosaccharides is attractive, particularly
for those sample that have been methyl-esterified to
stabilize sialic acid linkages. It is clear that this geometry
will be sensitive enough to allow analysis of oligosaccharides released from SDS–PAGE gel spots.
MASS SPECTROMETRY OF OLIGOSACCHARIDES
Although the responses appear to diminish as oligosaccharide molecular weight increases, ESI is more appropriate for use with fragile compound classes. Analysis of
sialylated (see Section V.D.) and sulfated (see Section
VI.C.) oligosaccharides with negative-mode ESI allows
extremely useful CID product-ion patterns to be produced.
The QIT analyzer has been shown to be particularly useful
for oligosaccharides because of its efficiency in acquiring
multiple stages of MS fragmentation. This efficiency
allows fragment ions to be produced that are absent in
MS2 profiles, increasing the amount of structural information produced. The FTMS carries similar advantages as the
QIT but with higher resolution and mass accuracy. That
resolution will be particularly useful for the analysis of
multiply charged ions such as those that are produced from
polysulfated carbohydrates with ESI (see Section VI.C). At
the present time, however, the application of FTMS to
this and other biomolecular structural questions is limited
by the lack of software development in commercial
instruments.
The development of hybrid QIT-TOF and QIT-FT
mass spectrometers will benefit the analysis of oligosaccharides by increasing the resolution and mass accuracy of
fragment ions. That increase is particularly important for
polysulfated oligosaccharides because the precursor ions
are multiply charged. Although QIT-TOF instruments are
available with MALDI, problems with metastable fragmentation, similar to those observed with TOF/TOF instruments (Medzihradszky et al., 2000; Yergey et al.,
2002), are likely to limit their application to oligosaccharides. The development of high-pressure MALDI and ESI
sources is, therefore, important.
The MALDI TOF/TOF is the only modern commercial
instrument that is capable of high-energy CID, and may
prove to be very useful for the production of cross-ring
cleavage ions for oligosaccharides—provided that the ions
can be produced without extensive metastable fragmentation before the collision cell. Because data published at
the present time indicate that there is a problem with metastable fragmentation, resulting in low-energy product ions
(Medzihradszky et al., 2000; Yergey et al., 2002), the development of a high-pressure source for that instrument is
needed.
IV. TANDEM MASS SPECTROMETRY
OF GLYCOPEPTIDES
A. Ionization of Glycopeptides
Glycopeptides ionize poorly with FAB ionization due
to their high mass, hydrophilicity, and limited surface
activities in liquid matrices (Carr et al., 1990). Synthesis of
tBoc derivatives of glycopeptides was observed to increase
&
the FAB sensitivity over underivatized glycopeptides by a
substantial degree (derivatized molecular mass 1522.8)
(Medzihradszky et al., 1990). The identification of Nlinked glycosylation sites with FAB relied on the use of
PNGase F, an enzyme that releases N-linked glycans and
converts the Asn side chains to which they are bound to
Asp. Knowing the protein sequence, the one Da shift in
mass was used to determine the site of modification from
the FAB MS/MS spectra (Carr et al., 1989). It was possible
to HPLC-fractionate the components of a proteolytic digest
of a glycoprotein before and after glycosidase treatment
and to conduct FAB MS analysis of the isolated fractions
(Carr & Roberts, 1986). Endoglycosidase cleavage of
N-linked glycans in the presence of H218O introduces an
isotopic label into the Asp residue created by the enzymatic
hydrolysis reaction, the mass signature of which is useful
for identification purposes (Küster & Mann, 1999).
The development of ESI (Meng, Mann, & Fenn, 1988)
resulted in significantly better glycopeptide signal response than was observed with FAB. This improvement
allowed the development of in-source fragmentation and
precursor-ion scans for the detection of oxonium ions
generated from glycopeptides eluting from an HPLC
column (Conboy & Henion, 1992; Huddleston, Bean, &
Carr, 1993). Analysis of glycopeptides with ESI has been
widely applied. The in vitro glycosylation of mucin peptides has been followed with ESI MS to determine the mass
and composition of the product glycopeptides (Tetaert
et al., 1994). Detection of oxonium ions has been used to
provide a detailed insight into site-specific distribution of
N-linked groups on immunoglobulin A (Kragten et al.,
1995) and for the selective identification of glycopeptides
in complex glycoprotein proteolytic digests prior to
analysis with MALDI-TOF coupled with exoglycosidase
digestion (Yang et al., 1997; Colangelo et al., 1999).
An oxonium ion has been used to selectively identify
a sialyl Lewisx antigen on a1-acid glycoprotein (Dage,
Ackermann, & Halsall, 1998) (for more details, see Section
IV.B.1).
The usefulness of MALDI-TOF MS for analysis of
glycoproteins was recognized soon after its invention
(Overberg et al., 1990). The MALDI-TOF mass spectrum
of a1-acid glycoprotein exhibited a broader peak than
observed for a typical unmodified protein in the same mass
range, and it was suggested that this broadness reflected the
glycoform heterogeneity (Hillenkamp et al., 1991). The
contribution of prompt and metastable fragmentation has
been evaluated for human interferon-g and found not to
contribute significantly to the MALDI-TOF MS peak width
(Mortz et al., 1996). Thus, MALDI-TOF MS peak widths
appear to depend on the heterogeneity of glycosylation.
Differences in glycosylation between fetal and adult cartilage oligomeric matrix protein have been identified with
MALDI-TOF MS analysis of the intact 80 kDa glycopro189
&
ZAIA
teins as well as proteolytic glycopeptides (Zaia et al.,
1997). The acquisition of MALDI-TOF mass spectra in
conjunction with glycosidase treatment is quite a useful
technique to determine the total mass of carbohydrate on a
glycoprotein (Zaia et al., 1997; Colangelo et al., 1999;
Jacoby et al., 2000; Tarelli et al., 2000; Zaia et al., 2001).
Enzymatic deglycosylation of proteoglycans can also be
accomplished in conjunction with MALDI-TOF MS (Zaia
et al., 2000).
B. CID of Glycopeptides
1. Selective Identification of Glycopeptides
with Tandem MS
Tandem MS of glycopeptides results in the formation
of ions that are useful for recognition purposes (Conboy
& Henion, 1992; Carr, Huddleston, & Bean, 1993;
Huddleston, Bean, & Carr, 1993). As shown in Figure 23,
a complex biantennary N-linked glycopeptide fragments
to produce ions that result from neutral losses from the
precursor, corresponding to HexmHexNAcn. It is, therefore, possible in theory to identify glycopeptides in complex tryptic digest mixtures with constant neutral-loss
scans. Also observed in glycopeptide CID mass spectra
are oxonium ions that correspond to Hexþ (C6H11O5þ, m/z
163.06), HexNAcþ (C8H14NO5þ, m/z: 204.09), and
HexHexNAc (C14H24NO10þ, m/z: 366.14). Note that the
abundance of the Hexþ ion is often quite low. Thus, a scan
for the precursors of those oxonium ions would appear to be
useful to identify glycopeptides. Neutral-loss scans are
carried out most often on triple quadrupole analyzers by
scanning precursor ions while monitoring only those product ions that result from a constant-neutral loss. Precursor
ions scans are accomplished by scanning the precursor ions
while detecting only those product ions at a given m/z
value. Because it is possible for a fraction of unmodified
peptides to produce CID product ions at the same nominal
mass as the carbohydrate oxonium ions, the higher the resolution of the selecting quadrupole, the more selective the
measurement. The performances of triple-quadrupole and
Q-oTOF analyzers have been compared for precursor-ion
scans for phosphopeptides identification (Steen, Küster,
& Mann, 2001). The results, showing that greater selectivity was achievable with the Q-oTOF analyzer, are likely
to be applicable to the detection of oxonium ions for
glycopeptides.
The precursor-ion scanning approach has been applied
to the selective identification of glycopeptides with reversed phase LC/MS (Carr, Huddleston, & Bean, 1993).
That work demonstrated the use of precursor ions scans for
m/z 204 for the detection of fetuin glycopeptides. The
approach entails comparing the LC/MS total-ion chromatogram with that corresponding to the precursors of m/z
204, and clearly identifies the elution times and m/z values
of glycopeptides in the tryptic-digest mixture. Another
example of that approach is shown in Figure 24 for the
analysis of a recombinant human thrombomodulin tryptic
FIGURE 24. Peptide/glycopeptide map of recombinant human thromFIGURE 23. Product-ion MS/MS of m/z 878 [M þ 2H]2þ. The asterisk
(*) indicates ions formed from two bond cleavages. Reproduced with
permission from Huddleston, Bean, & Carr (1993). Copyright 1993
American Chemical Society.
190
bomodulin tryptic digest. (A) UV chromatogram at 206 nm, (B) total-ion
chromatogram produced by positive-ion LC/MS, (C) precursor-ion scan
for m/z 204 (HexNAcþ). Reproduced from Itoh et al. (2002) with
permission. Copyright 2002, Elsevier.
MASS SPECTROMETRY OF OLIGOSACCHARIDES
digest mixture (Itoh et al., 2002). The data show that the
glycopeptides are clearly identified, based on the m/z 204
precursor ion scan shown in (C). Precursor-ion scans have
also been used to identify glycosylated peptides in unseparated tryptic digest mixtures ionized directly with nano
ESI MS (Wilm, Neubauer, & Mann, 1996). Glycopeptides
present in proteolytic digests of glycoproteins have
also been selectively eluted from reversed phase HPLC
columns under neutral pH conditions, facilitating their
analysis with LC/MS (Ohta et al., 2001). That approach is
limited by the complexity of the mixture that may give rise
to suppression of glycopeptide ions.
2. CID of O-Linked Glycopeptides
High-energy tandem mass spectrometric analysis of a
monomannosylated glycopeptide showed peptide-backbone product ions that contained the carbohydrate group
(Medzihradszky et al., 1990). In the same work, the tandem
mass spectrum of tBoc-YGP-(GalGalNAc)T-P-(GalGalNAc)S-A was dominated by the fragmentation of the carbohydrate side chains rather than of the peptide backbone.
An ion that correspond to the unmodified peptide was,
however, observed. Analysis of that ion with a subsequent
stage of MS would serve to identify the peptide portion of
the molecule. High-energy tandem mass spectrometric
analysis of an O-(HexHexNAc)-modified glycopeptide indicated an abundant series of peptide-backbone product
ions with a single yn ion that contained the disaccharide and
indicated its site of modification (Medzihradszky et al.,
1996). Abundant and useful peptide-backbone product
ions were also observed for a peptide modified with two
(HexHexNAc) groups. High-energy tandem mass spectra
of O-linked glycopeptides contained relatively more abundant ions produced by glycosidic fragmentation and lessabundant ions produced by peptide-backbone fragmentation as the size and number of carbohydrate modifications
increased (Medzihradszky et al., 1996).
Low-energy tandem mass spectrometric analysis of
a peptide modified with a single glycan of composition
(NeuGc)2GalGalNAc resulted in abundant product ions
that corresponded to sequential losses of NeuGc, NeuGc,
Hex, and HexNAc, respectively, from the precursor ion
(Hirayama et al., 1998). The ion that corresponded to the
unmodified peptide was abundant, and would be amenable
to analysis with a subsequent MS stage if desired. Lowenergy tandem mass spectrometric analysis of synthetic
O-linked glycopeptides on a QoTOF instrument has shown
that product ions are generated in very low abundance from
a cleavage of the peptide backbone, a few of which contain the carbohydrate group (Alving, Paulsen, & PeterKatalinic, 1999). Extended data-accumulation times are
required to achieve a sufficient signal-to-noise ratio for
these ions. The data sufficed to identify the sites of
&
glycosylation in synthetic glycopeptides that contained one
or two glycans of up to tetrasaccharide size. Abundant ions
were observed that corresponded to the unmodified peptide
that could be subjected to an additional MS stage for the
purpose of identification in the case of an unknown. Monoand difucosylated O-linked glycopeptides produce abundant fragments that corresponded to loss of dHex from
the precursor ion on an ESI QoTOF instrument (Macek,
Hofsteenge, & Peter-Katalinic, 2001). Peptide-bond product ions that contained the dHex group, however, were
observed at very low abundances in the spectra. It appears
that, at low collision energy, the ion that corresponds to
loss of dHex from the precursor ion does not undergo
subsequent fragmentation. The results demonstrated that
tandem mass spectra with sufficiently high signal-to-noise
ratio can be used to assign sites of O-linked fucosylation at
low collision energies.
Given the difficulties with assigning O-linked glycosylation sites, chemical elimination of the glycans has been
used in conjunction with tandem MS. A procedure based on
b-elimination with hydroxylamine incorporates ammonia
into the Ser or Thr residue to which the glycans had been
attached (Rademaker et al., 1998). An alternative approach
that used alkylamines to release the O-linked glycans produced a larger mass difference between the modified and
unmodified Ser/Thr residues, facilitating the analysis
(Hanisch, Jovanovic, & Peter-Katalinic, 2001).
3. CID of N-Linked Glycopeptides
ESI triple-quadrupole tandem mass spectra of highmannose type N-linked glycopeptides result in abundant
ions that correspond to fragmentation of the glycan
(Bateman et al., 1998; Hirayama et al., 1998; Zeng et al.,
1999; Zhu et al., 2000; Hui et al., 2001). An example of a
CID profile obtained from a Man9 N-linked glycopeptide
is shown in Figure 25 (Nemeth et al., 2001). Because of
the lack of HexNAc residues in the antennae, the glycan
portion of the molecule fragments at a collision energy high
enough to allow peptide-backbone fragmentation to be
observed, as shown in the top panel. The ions labeled in the
spectrum are identified in Table 5.
The ion that corresponds to HexNAc-modified peptides can be generated in-source by using elevated declustering potential and analyzed using MS/MS. A more
useful approach, given the complexity of ESI mass spectra
generated from glycoprotein proteolytic digests, would be
to use MS3 on a QIT or FTMS instrument. The generation
of such data would identify the site of modification for the
high-mannose structure, and augment the information on
the glycan produced in the MS/MS spectrum. A glycopeptide modified with two high-mannose N-linked glycans
produces an abundant series of ions that correspond to
respective losses of Man and GlcNAc, and displays an
191
&
ZAIA
FIGURE 25. Nano ESI MS/MS analysis of a triply protonated Man9 N-linked glycopeptide (a) the mass
spectrum where peptide yn ions are indicated, and (b) the singly charged deconvoluted mass spectrum of
panel a where the sugar-loss ion series is indicated. Reproduced from Nemeth et al. (2001) with permission.
Copyright 2001 American Chemical Society.
abundant ion that corresponds to the peptide with two
GlcNAc residues (Zhu et al., 2000).
As shown in Figure 26, complex N-linked glycopeptides undergo fragmentation of labile HexNAc and Neu5Ac
residues. The result is that an abundant ion that corresponds
to the peptide þ GlcNAc is observed, but that the peptidebond cleavage ions are lower in abundance than observed
in CID mass spectra of high-mannose structures. This lack
TABLE 5. List of the observed product ions from the nano ESI MS/MS analysis of a Man9 glycopeptide [C12 þ 3H þ Man9GlcNAc2]3þ
[reproduced from Nemeth et al. (2001) with permission, Copyright 2001 American Chemical Society]
a
M, Mannose.
G, N-acetylglucosamine.
b
192
MASS SPECTROMETRY OF OLIGOSACCHARIDES
&
FIGURE 26. Product-ion scan of a trisialylated tri-antennary N-linked glycopeptide [M þ 4H]4þ ion.
Reprinted with permission from Ritchie et al. (2002). Copyright 2002, Elsevier.
of peptide fragmentation behavior makes it difficult to
analyze the oligosaccharide and peptide structures in an
unknown glycopeptide in a single measurement. The observation of an ion that corresponds to peptide þ GlcNAc
allows a variation on the oxonium-ion precursor-ion scan
for selective detection of glycopeptides. By scanning an
unseparated glycoprotein tryptic digest for precursors to
the peptide þ GlcNAc ion (corresponding to the oligosaccharide Y1 ion), it is possible to identify all glycoforms that
modify a given peptide (Ritchie et al., 2002).
C. Electron Capture Dissociation of Glycopeptides
ECD is a powerful tool for peptide sequencing, and allows
the determination of the location of post-translational
modifications like O-glycosylation (Mirgorodskaya, Roepstorff, & Zubarev, 1999), N-glycosylation (Håkansson
et al., 2001), gamma carboxylation (Kelleher et al., 1999),
and phosphorylation (Stensballe et al., 2000; Shi et al.,
2001). It is believed that because of the high H. affinity of
the electron-rich amide carbonyls, minimal fragmentation
occurs to post-translational modification groups. It has
been noticed, however, that disulfide bonds tend to
fragment preferentially with ECD (Zubarev et al., 1999).
The finding that ECD produces peptide-backbone fragmentation with minimal fragmentation of glycans is particularly significant because it allows the determination of
sites of glycosylation and the size of the glycan.
Given that ECD cleaves the peptide backbone without
fragmenting the glycan, it is of interest to use it in combination with IRMPD or SORI-CID. IRMPD of glycopeptides results in an extensive fragmentation of the glycan
with little or no peptide-bond fragmentation, and thus
produces data complementary to that of ECD, as shown in
Figure 27 (Håkansson et al., 2001). Cleavages of 11 of the
15 backbone amine position are observed in the ECD
profile (a), providing data that would be useful for peptide
database searching. The fragmentation observed in the
IRMPD mass spectrum of the same glycopeptide (b),
resulted exclusively from the glycan, and was sufficient to
determine that there are three branching sites in the structure. SORI-CID of glycopeptides also results in preferential fragmentation of the glycan, and has also been used
to complement ECD (Kjeldsen et al., 2003). Although the
combination of ECD with IRMPD or SORI-CID presently
requires a specially modified FTMS instrument and a high
level of operator skill, a combined fragmentation approach
is likely to be extremely useful for the characterization
of protein glycosylation once the technology becomes
commercially available.
D. Conclusions
Mass spectrometric analysis of glycopeptides entails careful consideration of the ionization and the fragmentation
characteristics of the glycan portion of those molecules.
The goal of the analysis is to obtain as much information on
the glycan and peptide structures as possible for a limited
quantity of material, such as might be produced from SDS–
PAGE gel spots. It is essential not to produce prompt or
metastable fragmentation from the ionization process, and
thus the use of ESI is indicated at the present time.
Although the use of high-pressure MALDI is likely also to
produce protonated molecule ions from glycopeptides
without metastable decay, it is essential that the sources be
tested with this compound class to determine that the
conditions are cool enough. Use of the negative mode will
maximize the stability of native sialylated glycopeptides
193
&
ZAIA
FIGURE 27. Comparison of ECD and IRMPD product ion profiles for a glycopeptide. (a) ECD FTICR
mass spectrum obtained from the triply protonated N-glycosylated peptide of m/z 1005.5; (b) IRMPD
FTICR mass spectrum (displayed in three segments) from the triply protonated N-glycosylated peptide of
m/z 1005.5; (c) Peptide sequence and glycan structure of the investigated N-glycosylated peptide. Top:
Dissociation sites for ECD. Bottom: Major dissociation sites for IRMPD. Reproduced from Håkansson et al.
(2001) with permission. Copyright 2001 American Chemical Society.
during the MALDI process. Methyl esterification of sialic
acid residues is recommended if the positive MALDI mode
will be used. Use of traditional low-pressure MALDI
sources in the negative linear mode will be useful for mass
profiling.
CID fragmentation of glycopeptides results in the
production of abundant ions from glycosidic bond and very
low abundance ions from peptide-bond fragmentation. It is
194
generally not possible to identify the modified amino acid
residues, although the number of possibilities in a given
peptide will be limited. Even the determination of the
peptide mass from the CID profile is not a straightforward
matter. Useful data are obtained from an extended summing of ESI or MALDI CID data to obtain adequate signalto-noise ratios for the low-abundance peptide-backbone
cleavage ions. The use of ECD, however, results in
MASS SPECTROMETRY OF OLIGOSACCHARIDES
abundant peptide-backbone cleavage ions with no fragmentation of the glycan. It is, therefore, possible to assign
the modified amino acid residues even for cases in which
the peptide is modified with several glycan structures.
Clearly, the most useful approach for the analysis of
glycopeptide is to acquire ECD and CID mass spectra on an
FTMS instrument. Alternatively, a combination of ECD
and IRMPD can be used.
V. MASS SPECTROMETRY
OF SIALYLATED GLYCOCONJUGATES
A. Permethylation of Sialylated Oligosaccharides
Permethylation has been used for 40 years as a technique
to improve the analytical properties of glycoconjugates
(Hakomori, 1964). The primary advantages of this derivatization step for mass spectrometry are that (1) the ionization responses for FAB, ESI, and MALDI are improved
relative to those of the native molecules, and (2) that the
product-ion mass spectra are much more informative relative to those of the underivatized sugars. These advantages must be balanced against the disadvantages of the
derivatization step itself. The derivatization chemistry
places a lower limit on the quantity of sample that can be
analyzed, and as a result does not appear to be appropriate
for analysis of samples available only in small quantities
such as from gel spots.
Use of the solid sodium hydroxide technique (Ciucanu
& Kerek, 1984) has been found to be effective to permethylate sialylated glycoconjugates, and has come to be widely
used (Dell et al., 1994; Perrault & Costello, 1994;
Reinhold, Chan, & Reinhold, 1994; Reinhold, Reinhold,
& Costello, 1995; Reinhold & Sheeley, 1998; Viseux, de
Hoffmann, & Domon, 1998). The technique involves
vortexing the dried sample in anhydrous dimethyl sulfoxide in the presence of powdered sodium hydroxide,
followed by an incubation at room temperature in the
presence of methyl iodide. The reaction is quenched by
the addition of water, and the permethylated molecules are
extracted into chloroform. Note that this procedure is not
recommended for sulfated oligosaccharides because they
are not soluble in chloroform. The short Hakomori method
is used for those cases (Dell et al., 1994). In addition, the
solid sodium hydroxide method is not recommended for
base-labile molecules.
Permethylated sialic residues are hydrolyzed under
basic conditions, and a modification to the permethylation
procedure is recommended (Barr et al., 1991). That modification entails back-extracting the permethylated mixture
with 30% acetic acid to minimize the chance of baseinduced hydrolysis (Reinhold & Sheeley, 1998; Viseux, de
Hoffmann, & Domon, 1998; Viseux, Costello, & Domon,
&
1999; Delaney & Vouros, 2001; Viseux et al., 2001). Baseinduced hydrolysis of permethylated sialylated oligosaccharides has been used to simplify the interpretation of
the product-ion mass spectra of those molecules (Viseux,
de Hoffmann, & Domon, 1998; Viseux, Costello, &
Domon, 1999). The hydrolyzed permethylated sialic acid
groups leave an open hydroxyl group that may be
perdeuteromethylated. The position of sialic acid substitution can be determined from the CID spectrum (see Section
V.D).
B. MALDI-MS of Sialylated Glycoconjugates
Early work on carbohydrate identification from proteolytic
digests of recombinant glycoproteins demonstrated that the
majority of ions observed in the linear mode for sialylated
glycoproteins correspond to metastable ions (Huberty et al.,
1993). The loss of most Neu5Ac residues is observed in the
reflector mode with a-cyano-4-hydroxycinnamic acid or
sinapinic acid MALDI matrices. The use of 2,6-dihydroxyacetophenone with ammonium citrate was subsequently
observed to diminish that fragmentation (Pitt & Gorman,
1996). A similar matrix, 2,4,6-trihydroxyacetophenone
(THAP) with ammonium citrate, was found to enable the
detection of sialylated N-linked sugars without decomposition and to facilitate analysis of sialylated glycopeptides (Papac, Wong, & Jones, 1996).
Given this metastable decay, acidic glycoconjugates
are best analyzed in the negative linear MALDI-TOF
mode. With careful use of threshold laser irradiance and a
cool matrix such as DHB, it is possible to detect ions that
correspond to intact acidic glycoconjugates. The ease with
which this detection is possible varies for different instrumental geometries and source designs, and it is advisable to
test sample preparation and source conditions in the negative linear mode with sialylated standards. Without the use
of such standards, it is difficult to determine whether ions
that differ by the mass of a sialic acid residue result from insource fragmentation or are present in solution (Kakehi
et al., 2001). Although neutral and acidic N-linked glycans
have been analyzed simultaneously in the positive mode
on a MALDI Q-oTOF instrument, it is not clear how
much in-source fragmentation of sialylated glycans occurs
(Hanrahan et al., 2001). A sialylated glycan has been
analyzed in the positive MALDI-TOF mode as a mixture of
[M þ H]þ, [M(Na) þ H]þ, and [M(K) þ H]þ ions without
reported loss of the sialic acid residue (Charlwood et al.,
1999). That observation is consistent with significant variability in the prevalence of in-source decay for sialylated
glycans among different instruments from different manufacturers. To simultaneously obtain equal ionization responses for neutral and acidic glycoconjugates, it is best to
modify either the MALDI matrix or to esterify the glycans
(see below).
195
&
ZAIA
1. Anionic Dopants for Analysis
of Sialylated Glycoconjugates
Neutral oligosaccharides are usually analyzed in the positive ion mode using a MALDI-TOF with adducts of alkali
cations such as Liþ, Naþ, or Kþ. Neutral and sialylated
oligosaccharides bind those cations differently, and thus
produce different ionization responses in positive-ion
MALDI-TOF MS. The addition of a low concentration of
H2SO4 to the matrix solution results in the observation of
[M þ H2SO4] ions for neutral oligosaccharides with a
harmane matrix (Wong et al., 1999). Sialylated oligosaccharides produce predominantly [M–H] ions under the
same conditions, and a mixture of neutral and acidic
oligosaccharides results in the observation of ions that
correspond to both oligosaccharides in the same spectrum.
In an extension of that work, alkylsulfonates have been
used as anionic dopants in place of H2SO4 (Wong, Wang, &
Lebrilla, 2000). Of the several alkylsulfonates and matrices
tested, the combination of sulfanilic acid with dihydroxyacetophenone was found to be most effective for the
analysis of mixtures of neutral and acidic oligosaccharides.
The ratio of intensity of a mixture that contained a neutral
and an acidic oligosaccharides was found to be linear,
with a correlation coefficient of 0.99, as the concentration
was varied, indicating that suppressive effects are minimized during the ionization process. It was possible to
follow the course of sialic acid hydrolysis, using the
combination of sulfanilic acid and dihydroxyacetophenone
matrix.
2. Methyl Esterification to Stabilize Sialic
Acid Residues
Conversion of sialic acid residues to methyl esters (Handa
& Nakamura, 1984) allows the detection of sialylated
oligosaccharides and ganglioside ions in the positive-ion
mode without metastable decay (Powell & Harvey, 1996).
The sialylated molecules are dissolved in dimethylsulfoxide, and incubated with methyl iodide at room temperature; MALDI analysis can be accomplished without
sample cleanup following removal of solvent. Sialic acids
can also be separated from neutral oligosaccharides with
anion exchange chromatography, and conveniently and efficiently converted to methyl esters while still bound by
eluting dimethylsulfoxide and methyl iodide (Karlsson,
Karlsson, & Hansson, 1995). Methyl esterification has
also been used to facilitate MALDI-TOF analysis of Nlinked glycans released by treatment of SDS–PAGE gel
spots with glycosidases (Küster et al., 1997). Once methylesterified, the glycans are passed through a gel-loader
pipette tip that is packed with C18, anion, and cation
exchange resins to remove salts and detergents prior to
MALDI MS. The acidic residues are rendered neutral
196
through esterification, allowing all released glycans to be
analyzed in a single MALDI-TOF mass spectrum. The
methodology has been applied to profile an a1-acid glycoprotein from different species with MALDI-TOF MS
(Küster et al., 1998).
Digests of pectin, a homogalacturonan, have been
analyzed as methyl esters, using MALDI-TOF with THAP
matrix (Körner et al., 1998), and the method has been used
to study enzyme specificities of pectin lyases (Körner et al.,
1999). Underivatized polygalacturonans derived from alginate have been analyzed with perfluorosulfonated ionomer
(Nafion)-coated probes for MALDI-TOF MS (Jacobs &
Dahlman, 2001). The immobilized sulfate groups evidently
scavenge sodium and potassium ions, resulting in the detection of [M–H] ions with little or no adduction and
much cleaner spectral profiles. Taken together, those
studies show how MALDI-TOF can be used to provide
useful information on acidic biopolymers.
3. Perbenzolylation to Stabilize Sialic
Acid Residues
Perbenzolylation (Daniel, 1987) has been shown to be
effective to stabilize sialic acid groups and to allow their
analysis with positive-mode MALDI-TOF MS (Chen et al.,
1999). Carbohydrates may be benzoylated in 50 pmol
quantities with benzoyl anhydride followed by purification
on a C-18 cartridge. The derivatization step is general in
nature, and will work on reduced and unreduced oligosaccharides. The spectra of benzoylated glycans that contain 2-3 linked sialic acid residues produce distinctly
different MALDI-TOF mass spectral patterns than those
that contain 2-6 linked residues. The former undergo internal lactonization of the carboxyl groups during the derivatization process, resulting in the addition of one fewer
benzoyl group than the 2-6 linked sialic acid. The technique may be useful to differentiate sialic acid linkages in
relatively pure samples. CID of perbenzylated cerebrosides
has also been shown to be useful to differentiate oligosaccharide isomers (Perreault & Costello, 1996).
4. High-Pressure MALDI
of Sialylated Glycoconjugates
The use of FTMS in the analysis of glycoconjugate mixtures carries the significant advantages that the highest
possible resolution and mass accuracies are obtained. The
use of MALDI with FTMS, as with other analyzers, is
advantageous because it has a lower sensitivity to contaminants than does ESI, and it is more easily applied to
samples from multiwell formats. In addition, MALDI is
more easily interfaced with surface techniques such as thinlayer chromatography and gel electrophoresis. Because
the MALDI source is thoroughly decoupled from the FT
MASS SPECTROMETRY OF OLIGOSACCHARIDES
analyzer, the resolution and mass accuracy do not depend
on the smoothness of the desorption surface. The quality of
MALDI-TOF spectra, on the other hand, is significantly
degraded by surface irregularities. The use of MALDI
FTMS in the analysis of fragile biomolecules has been
limited until recently by long-lived metastable decay of
MALDI ions prior to detection. Such decay is detrimental
to the analysis of mixtures becasue it is not possible to
determine whether only a given ion is an independent
species, or is one that is produced from fragmentation of
another ion.
MALDI of ions at elevated pressure has been
demonstrated to produce diminished metastable decay
(Krutchinsky et al., 1998; Laiko, Baldwin, & Burlingame,
2000; Loboda et al., 2000). As Figure 28 shows, sialylated
gangliosides are observed as de-sialylated ions with low
pressure MALDI FTMS (O’Connor & Costello, 2001).
Desorbing the ions using the same conditions but at
elevated pressure (1–10 mbar) results in the observation of
the intact sialylated ganglioside with ions produced from
losses of sialic acid residues in very low abundances. Those
results demonstrate that the thermal stabilization of vibrationally excited MALDI ions dramatically reduces the
degree of metastable fragmentation observed for fragile
sialylated glycoconjugates. Gangliosides with as many as
5 sialic acid residues have been observed by MALDI
FTMS with minimal metastable fragmentation (O’Connor,
Mirgorodskaya, & Costello, 2002). Those developments
will allow the direct profiling of glycolipids from thin-layer
chromatography plates, and will facilitate the analysis of
any sialylated glycoconjugate.
&
C. ESI MS of Sialylated Oligosaccharides
Given that ionization is dependent on the oligosaccharide
composition with respect to neutral and acidic residues,
modification of the carbohydrate structure is necessary to
provide a profile that reflects the relative quantities of each
mixture component. One way to equalize ionization responses is to remove sialic-acid residues enzymatically,
thereby rendering all oligosaccharide structures neutral,
assuming that sulfated residues are not present. That step
will simplify the mass profile due to the elimination of
sialic acid heterogeneity (Küster et al., 1997, 1998). To
preserve the sialic acid structure while producing a uniform
ionization response, it is necessary to neutralize the acidity
of the carboxyl groups by methyl ester formation. The
methods described in Sections III.A, V.A, and V.B.2 are
applicable to the ESI MS of sialylated oligosaccharides.
Permethylation also renders sialylated and neutral
oligosaccharides chemically equivalent. It is also possible
to permethylate sialylated sulfated glycans, although a
more involved procedure must be used (see Section VI.A)
(Dell et al., 1994). The permethylated oligosaccharides
produce stronger ion signals in ESI due to their increased
surface activities (Bahr et al., 1997; Cole, 1997; Karas,
Bahr, & Dulcks, 2000). The signals are also less sensitive to
the composition of the oligosaccharide. The ESI response
is such that quantitative analysis can be performed on
oligosaccharides released from recombinant glycoproteins
(Viseux et al., 2001). Quantitative characterization of
glycan mixtures has traditionally involved reducing-end
labeling with a chromophore such as 2-aminobenzamide
FIGURE 28. GangliosideGT1b desorbed from ATT matrix (a) without collision gas, and (b) with collision
gas. Peaks due to the presence of a lower homolog are marked LH. Reprinted with permission from
O’Connor & Costello (2001). Copyright 2001 John Wiley & Sons Limited.
197
&
ZAIA
TABLE 6. Relative intensities of the components identified by
mass spectrometric and HPLC analyses of the sCD4 glycan pool
[reprinted with permission from Viseux et al. (2001), Copyright
2001 American Chemical Society]
a
Symbols: B, biantennary N-glycan; F, fucosyl residue; S, sialyl
residue; MM, molecular mass.
FIGURE 29. HPLC chromatogram of the 2-aminobenzamide-deriva-
tized N-glycans (a) and electrospray mass spectrum of the permethylated
oligosaccharide derivatives (b) released from human R1-acid glycoprotein. Symbols: X includes biantennary (B), triantennary (T), and
tetraantennary (Q) N-glycans; F, fucosyl residue; S, sialyl residue; L,
lactosamine unit. Reprinted with permission from Viseux et al. (2001).
Copyright 2001 American Chemical Society.
followed by HPLC separation. Figure 29 compares the
HPLC chromatogram obtained from a1-acid glycoprotein
after labeling with 2-aminobenzamide with the ESI mass
spectrum obtained from the same sample after permethylation. The glycans are separated into peaks based on the
antenna structure and degree of sialylation. The ESI mass
spectrum of the permethylated mixture (lower panel)
demonstrates that significantly more detailed information
is obtained. Specifically, although the fucosylated glycoform is clearly resolved in the mass spectrum, it is not in the
chromatogram of the 2-aminobenzamide-labeled material.
As Table 6 shows, good agreement in quantitation was
obtained between the two methods. The coefficients of variation values obtained for the mass spectrometric method
were reported to be less th an 3% for most measurements.
In addition to an increased ionization efficiency and
desirable fragmentation patterns (Viseux, de Hoffmann, &
Domon, 1998), permethylated oligosaccharides may be
analyzed by reversed-phase HPLC. Permethylated oligo198
saccharides may be eluted on-line into an ESI mass
spectrometer with acetonitrile/water gradients that contain
1 mM sodium acetate to create stable sodium adducts
(Delaney & Vouros, 2001). The technique has been shown
to be applicable for the analysis of complex N-linked
glycans with or without a 2-aminobenzamide label. For
very complex mixtures, it is often not possible to detect
minor component in the absence of chromatographic
separation. Chromatographic separation simplifies the
complexity of the mixture that enters the instrument at
any moment, addressing this limitation. Thus, with on-line
separation, the problem of overlapping peaks in mass
spectra is minimized. An alternative is to remove sialic acid
residues enzymatically to reduce the complexity to the
point that minor components can be detected without
chromatographic separation (Viseux et al., 2001).
D. Tandem MS of Sialylated Oligosaccharides
Unlike their neutral counterparts, native sialylated
oligosaccharides produce abundant signals in negative
ionization. Using negative FAB MS, sialylated milk oligosaccharides displayed metastable fragmentation that
provided useful structural information (Egge & PeterKatalinc, 1987). A series of molecules with Neu5Ac at the
non-reducing termini display almost exclusively fragment
ions that contain that residue. This result is contrasted with
the fragmentation of [M–H] asialo carbohydrates and
glycoconjugates, in which reducing and non-reducing
terminal fragmentation is observed (Domon & Costello,
1988b). Fragmentation to the reducing terminal side of
HexNAc residues is facile in the product ion spectra of
[M–H] ions (Peter-Katalinic, 1994). Interestingly, when
Neu5Ac is present as a substituent of a mid-chain Gal
MASS SPECTROMETRY OF OLIGOSACCHARIDES
residue of a glycosphingolipid, most of the fragment ions
contain the acidic residue, indicating that fragmentation
centers around this locus (Egge & Peter-Katalinc, 1987;
Costello & Vath, 1990). By contrast, ionization of sialylated
glycoconjugates (glycosphingolipids) as [M þ H]þ ions
results in the preferential elimination of Neu5Ac (Egge &
Peter-Katalinc, 1987), and less informative mass spectra.
Permethylated glycosphingolipids fragment to produce
a pattern that identifies the presence of Neu5Ac, but may
not unequivocally locate the substitution position due the
high abundance of ions that correspond to the oxonium ion
(Costello & Vath, 1990).
Oligosaccharides that differ only in the linkage position of sialic acid groups may be differentiated based on
high-energy CID patterns. The natriated permethylated
derivatives of the pentasaccharide Neu5Ac(2–6)Gal(1–
4)GlcNAc(1–3)Gal(1–4)Glc produce abundant ions that
contain the nonreducing terminus and those that contain
the reducing terminus in low abundance (Lemoine et al.,
1991). The permethylated isomer with (2–3) linked
Neu5Ac, on the other hand, produces a pattern in which
the most abundant product ions contain the reducing
terminus. These observations suggest that the Neu5Ac(2–
6)Hex linkage results in a more stable sodium adduct,
resulting in charge-remote fragment ions that contain the
non-reducing terminus. The CID product-ion spectra of
natriated permethylated oligosaccharides was, therefore,
seen as an alternative to that of protonated ions that had
the potential to provide additional sensitivity to the linkage
structure.
E. Conclusions
For MS profiling of glycoconjugates and oligosaccharides,
it is important to consider that neutral and acidic molecules
will produce different ionization responses by MALDI or
ESI. In cases where it is desirable to simultaneously obtain
a mass profile of all components, chemical modifications
are necessary to obtain even ionization responses for all
components. Methyl esterification of sialic acid residues
renders them neutral, can be accomplished on a small
sample quantity (i.e., from glycans released from SDS
PAGE gel spots), and is applicable to MALDI and ESI.
Alternatively, the mixture of neutral and acidic molecules
can be analyzed in the negative MALDI mode, using an
anion dopant to equalize the ionization responses. For
many underivatized sialylated glycoconjugate or oligosaccharide mixtures, however, the complexity is such that
fractionation is required before acquiring mass spectra
to avoid overlapping peaks. Ion-exchange separation of
neutral and acidic molecules is straightforward and meets
the need of mixture simplification. After separation, the
neutral molecules may be analyzed with positive-mode
MALDI or ESI and the acidic molecules in the nega-
&
tive mode. Alternatively, the acidic pool can be methyl
esterified to facilitate its analysis in the positive mode.
Another approach is to permethylate both pools and
analyze them with MALDI or ESI with or without
reversed-phase fractionation. As described in Figure 22,
the choice of derivatization procedure depends on the
sample quantity. Methyl esterification is recommended
when analyzing glycan isolated from gel spots.
VI. MASS SPECTROMETRY OF SULFATED
OLIGOSACCHARIDES
Sulfated oligosaccharides have historically posed particular analytical challenges due to their acidity and lability.
Progress in the mass spectrometric analysis of sulfated
glycoconjugates has lagged behind that of other compound
classes, with serious implications for the understanding of
the biological activities of those structures. Because of
those challenges, the uses of mass spectrometry in the
analysis of sulfated carbohydrates have been limited in
scope.
A. Derivatization
Sulfate groups are not compatible with the solid sodium
hydroxide method of permethylation (see sections III.A
and VIII.A) due to their lability under basic conditions
and difficulties associated with the extraction step. Those
molecules must be permethylated with the Hakomori
method (Hakomori, 1964; Dell et al., 1994), a procedure
that leaves the sulfate groups underivatized. The molecules
may be analyzed directly, or subjected to further derivatization to produce fully derivatized glycans. In the latter
scheme, the sulfate groups are chemically hydrolyzed, and
the free hydroxyl groups formed are permethylated with
trideuteromethyl iodide with the solid sodium hydroxide
method (Viseux, de Hoffmann, & Domon, 1998; Viseux,
Costello, & Domon, 1999). The positions of the sulfate
groups are indicated by the masses of the trideuteromethyl
groups. In addition, the glycans are no longer acidic, and
will ionize equally well with permethylated derivatives of
neutral and sialylated molecules, enabling quantitative
mass spectrometric profiling; see Section V.C.
B. Ionization Methods
1. Fast Atom Bombardment
By nature, sulfated oligosaccharides are easily ionized on
account of the acidity of the sulfate groups. Nonetheless,
this class has proven difficult to analyze because of the
fragility of the sulfate groups during the ionization and
dissociation processes. As summarized in Table 7, FAB has
199
&
ZAIA
TABLE 7. Summary of the use of FAB MS for the analysis of sulfated oligosaccharides
(Continued)
200
MASS SPECTROMETRY OF OLIGOSACCHARIDES
&
TABLE 7. (Continued)
a
Abundant ions corresponding to losses of 102 u (NaSO3) from the precursor ion were observed.
Ion produced from glucosidic-bond cleavages observed in the FAB mass spectrum.
c
A mixture of natriated precursor ions was observed corresponding to [Mn(Na)–H].
d
No losses of 80 u (SO3) or 102 u (NaSO3) were observed from the precursor ion.
e
Derivatized with a lipid group using reductive amination.
f
Abundant losses of 80 u (SO3) were observed from the precursor ion.
g
A glycosphingolipid.
h
FAB MS of a mixture of peracetylated oligosaccharides in which the sulfate group is not modified.
i
The oligosaccharide was permethylated, leaving the sulfate groups unmodified.
j
Losses of KSO3 from the precursor ion were observed.
b
been used to ionize sulfated carbohydrates with sensitivities in the 5-50 nmol range. Typically, the molecules are
observed as sodium adducts; the number of associated
sodium ions increases with the number of sulfate groups.
For a monosulfated disaccharide, it was possible to observe
ions in the positive and negative modes (Carr & Reinhold,
1984). In the positive mode, a mixture of [M(nNa) þ H]þ
ions were detected, where n ¼ 2, 3. Abundant ions that
corresponded to the loss of 102 u, which corresponded to
the loss of NaSO3 with proton replacement, were observed
in addition to glycosidic-bond cleavages. In the negative
mode, [M–H] and [M(Na)–H] were observed, with no
losses of 102 u, in addition to ions that corresponded to
glycosidic bond cleavages. For disulfated tetramers, losses
of 102 u were observed in the negative mode in addition to
glycosidic bond cleavages. Those data demonstrate that
the FAB ionization process is energetic enough to break
covalent bonds in carbohydrates, and that glycosidic bond
fragmentation and losses of NaSO3 occur under similar
conditions.
Work on monosulfated mucin oligosaccharides shows
that positive ionization results in [M(Na) þ H]þ ions with
concomitant losses of NaSO3 with proton replacement
(Mawhinney & Chance, 1994). Similar results are obtained
from disulfated disaccharides in the positive FAB ionization mode (Reinhold et al., 1987). Singly sulfated mucin
oligosaccharides may be detected in the negative FAB
ionization mode as [M–H] ions, where glycosidic-bond
cleavages are observed with no losses of sulfate-containing
moieties (Strecker et al., 1987). Disulfated mucin oligosaccharides may be observed as [M(Na) þ H]þ ions in the
positive mode, where losses of 102 u and glycosidic bond
cleavages are observed. The ionization of polysulfated
oligosaccharides can be accomplished in the negative ion
FAB mode, resulting in polysulfated protonated molecule
ions, losses of NaSO3, and glycosidic bond cleavages (see
Table 7 for references).
The FAB sensitivity of sulfated oligosaccharides can
be enhanced through permethylation or peracetylation
(Dell et al., 1988; Khoo et al., 1993; Karlsson, Karlsson, &
Hansson, 1996), although losses of NaSO3 are still
observed. Polysulfated oligosaccharides as lipid conjugates can be ionized at [M–H] ions, resulting in losses of
80 u (SO3), in addition to glycosidic-bond cleavages (Chai,
Rosankiewicz, & Lawson, 1995). In some cases, it is not
possible to determine whether losses correspond to 80 or
102 u, due to the presence of a distribution of [M(nNa)]–
H] ions (Reinhold et al., 1987; Dell et al., 1988; Mallis
et al., 1989; Linhardt et al., 1992). In other cases, it is clear
that a loss of 102 u occurs from the protonated molecule ion
(Mawhinney & Chance, 1994). For the purposes of this
discussion, losses observed in the spectra of polysodiated
201
&
ZAIA
sulfated oligosaccharide protonated molecule ions will be
considered to correspond to 102 u, NaSO3 with proton
replacement. A tetrasulfated CS octamer was desalted
against ammonium salts with gel filtration, and was
analyzed with FAB MS in the negative mode, resulting in
an ion that corresponded to [M–3(SO3)–H] (Chai,
Kogelberg, & Lawson, 1996). The addition of increasing
quantities of NaHCO3 to the sample resulted in successively decreasing abundances of ions produced from losses
of NaSO3. Ultimately, it was possible to observe a series of
[M(nNa)–H] ions, where n ¼ 3–6, with no losses of
NaSO3 or SO3. Those results are consistent with the
conclusion that pairing sulfate groups with sodium ions
increases their stability during ionization.
2. MALDI
a. MALDI of Sulfated Peptides. The Use of MALDI for
the analysis of sulfated peptides is reviewed here because
the findings are instructive for the analysis of sulfated
oligosaccharides. Early studies on sulfated peptides with
FAB ionization suggested that positive ions generated from
those molecules undergo facile losses of SO3, with the
result that the peptide may not appear to be sulfated
(Gibson & Cohen, 1990). Monosulfated peptides are
observed as [M–H] ions in the negative FAB ionization
mode, indicating its use for such molecules. Disulfated
peptides, however, are observed as a mixture of [M–H]
and [M–SO3 –H] ions by negative FAB ionization
(Yagami, Kitagawa, & Futaki, 1995). For such ions, one
of the sulfate groups is protonated and one deprotonated,
and the results are consistent with the loss of the deprotonated sulfate group during the ionization process. It is
possible to detect a sulfated peptide in the negative mode
with MALDI-TOF MS as an [M–H] ion, and to observe
a combination of [M–SO3 þ H]þ and [M þ H]þ in the
positive mode (Talbo & Roepstorff, 1993; Zaia et al.,
2001), an effect that can be used to identify sulfated peptides in digest mixtures. The degree to which this metastable decay occurs is dependant on the mass spectrometer
source design and on the MALDI matrix used. Although,
when using a conventional MALDI-TOF source in the
positive mode, no [M þ H]þ ion was detected for a sulfated conotoxin peptide, the use of atmospheric-pressure
MALDI ionization resulted in the observation of that ion in
addition to those that resulted from SO3 loss (Wolfender
et al., 1999; Laiko, Baldwin, & Burlingame, 2000). In the
linear negative mode, however, the [M–H] ion was 5–10fold more abundant than the [M–SO3 –H] ion. In the
negative reflectron mode, the two ions were equal in
abundance, showing that metastable fragmentation occurs
to a significant degree. The analysis of sulfated peptides is,
therefore, best undertaken in the negative MALDI mode,
and the use of atmospheric pressure sources is also war202
ranted. Using atmospheric-pressure MALDI on an ion trap,
ions that correspond to the loss of SO3 were observed in the
positive mode, and intact [M–H] ions were observed in
the negative mode (Moyer et al., 2002).
b. Direct MALDI of Sulfated Oligosaccharides.
MALDI-TOF MS has been used to detect monosulfated
mucin tetrasaccharide alditols as [M(2Na) þ H]þ ions with
either 2,5-dihydroxybenzoic acid or 3-aminoquinoline
matrices (LoGuidice et al., 1997). The result implies that
monosulfated carbohydrates are amenable to MALDI
ionization in the positive mode in the sodium-adducted
form. In-gel PNGase F digestion has been used to release
N-linked glycans from bovine thyroid-stimulating hormone (Wheeler & Harvey, 2001). The neutral oligosaccharides were separated from the sulfated and sialylated
oligosaccharides with a porous graphitized carbon column,
and the fractions were analyzed with MALDI-TOF MS.
The acidic fractions were analyzed with D-arabinosazone
as a matrix, and the spectra exhibited abundant [M þ Na]þ
ions that correspond to mono- and di-sulfated biantennary
N-linked structures. Significantly, several ions that correspond to sulfated glycans were observed with no losses of
80 u (SO3) or 102 u (NaSO3 with proton displacement),
indicating that metastable decay does not occur under the
conditions used.
Sulfated N-linked glycans of bovine peripheral myelin
have been analyzed with a trihydroxyacetophenone matrix
in the negative MALDI mode (Gallego et al., 2001). The
results showed that a mixture of [M–H] and [M(Na)–H]
ions were observed with little or no losses of 80 or 102 u.
Sulfated N-linked glycans have also been analyzed effectively by negative ion MALDI-TOF with a DHB matrix
(Zamze et al., 2001). Chondroitin sulfate solutions digested
with chondroitinase ABC or hyaluronidase have been
analyzed by MALDI-TOF with a DHB matrix (Schiller
et al., 1999). The results for those polysulfated oligosaccharides are similar to those obtained on sulfated oligosaccharides with FAB ionization in that losses of 102 u
(NaSO3 with proton displacement) are observed from
many of the ions in the spectra. The results are ambiguous
because they are consistent with either the presence of a
substantial population of undersulfated oligosaccharides
in the enzymatic digest mixtures or mass spectrometric
fragmentation of the ions.
To summarize, traditional low-pressure MALDI can
be used for monosulfated and, provided care is taken to use
cation adduction, disulfated oligosaccharides. With more
than two sulfate groups present, the use of MALDI is not
recommended due to abundant metastable losses. Highpressure MALDI has been used to suppress fragmentation
of fragile molecules such as polysialylated gangliosides
(O’Connor & Costello, 2001; O’Connor, Mirgorodskaya,
& Costello, 2002). The extent to which high-pressure
MASS SPECTROMETRY OF OLIGOSACCHARIDES
&
MALDI may be useful to analyze polysulfated carbohydrates is not clear at this time.
c. Use of Basic Peptides for MALDI of Polysulfated
Oligosaccharides. Polysulfated carbohydrates are detected only in very low abundance with MALDI-TOF, with
extensive losses of SO3 and HSO4 (Juhasz & Biemann,
1995). A vastly improved signal is obtained when the
polysulfated oligosaccharide is mixed 1:1 with a basic
peptide of sequence (RG)n, with abundant peaks that correspond to the complex. The abundance of the ions produced from losses of SO3 were reported to be dependent on
the matrix used; the use of 3-hydroxypicolinic acid (3HPA) results in very low abundances for those losses.
Further, it was shown that equimolar mixtures of a heparin
oligosaccharide and a basic protein, such as angiogenin,
produce ions that correspond to the complex in MALDITOF mass spectra acquired from 3-HPA. That work led to
marked progress in the analysis of heparin sequences by
determining the molecular weight of complexes before and
after chemical or enzymatic degradation. Complexes between heparin oligosaccharides and the peptide (RG)19R
were ionized from caffeic acid, and the data used to show
that the heparin-degrading enzyme heparinase II acts by an
endolytic mechanism (Rhomberg et al., 1998). The same
conditions were used to show that heparinase I cleaves by
a predominantly exolytic mechanism (Ernst et al., 1998).
Derivatization of heparin oligosaccharides to reducingterminal semicarbazones has been used to help distinguish
products of enzymatic degradation that contain the original
reducing- and non-reducing-termini (Rhomberg et al.,
1998).
Figure 30 illustrates the use of basic peptide complexation to monitor the course of a heparin-lyase digestion
of a mixture of heparin-like glycosaminoglycan (HLGAG)
oligosaccharides as a function of time (Rhomberg et al.,
1998). Sucrose octasulfate (SOS) was used as an internal
standard, relative to which the abundances of the HLGAG
oligosaccharides were measured to obtain a semi-quantitative digestion profile. HLGAG oligosaccharides H1 and
T5 with SOS were digested with heparin lyase I, and
aliquots were removed at time points and diluted with
matrix solution to stop the reaction. Caffeic acid was used
as the matrix, and the solution contained (RG)15 at an
approximately 2:1 molar excess over HLGAG oligosaccharides. The results are impressive in that the degradation
of H1 to produce T7 can be followed with a relatively
simple procedure that involves no derivatization or
chromatographic steps. The masses of the HLGAGs are
determined by subtracting the observed m/z value of
the peptide [M þ H]þ ion (labeled P in Figure 30) from
those observed for the peptide-HLGAG complexes.
Those results are significant because they show that
MALDI-TOF, with its ease of sample handling and well-
FIGURE 30. (A) MALDI mass spectrum of a mixture of substrates H1
and T5, as well as sucrose octasulfate (SOS), after a 40-min digestion
with heparinase I. (B) Peak ratios of oligosaccharides H1, T5, and T7
relative to SOS during a 2-hr heparinase I digest. Reprinted with
permission from Rhomberg et al. (1998)). Copyright 1998 National
Academy of Sciences, USA.
developed instrumentation, can be used in the mass analysis
of HLGAGs. The technique is limited because the use of a
complexing peptide requires a great deal of skill and patience, making the experiment time-consuming. In addition, the fact that the charge is carried by the complexing
peptide makes it impossible to conduct CID of the HLGAG
ions. To gain sequence information requires the use of
chemical or enzymatic degradation and the use of pure
oligosaccharides. It is possible to mass-analyze HLGAGs
without the use of complexing molecules with negative ESI
(Chai et al., 1998; Pope et al., 2001) and to obtain sequence
information with CID (Zaia & Costello, 2003), see Section
VI.C.6.b.
A detailed notation system to represent HLGAG sequences has been described, and was used in combination
with chemical and enzymatic degradation steps to determine sequences of oligosaccharides involved in anticoagulation, cell growth, and differentiation (Venkataraman
203
&
ZAIA
et al., 1999b). That approach was used to sequence a 3-Osulfate-containing antithrombin III-binding HLGAG decamer with a caffeic acid matrix and (RG)19R as the basic
peptide (Shriver et al., 2000). It was also shown with
MALDI-TOF MS that glucosamine 3-O sulfation at the
reducing end of a glycosidic linkage imparts resistance to
glycosidic cleavage by heparinases I, II, and III (Shriver
et al., 2000). Recently, MALDI-TOF MS combined with
the semicarbazide derivatization scheme and heparinase
digestion have been used in the structural characterization
of a 3-O-sulfated octasaccharide that binds to herpes
simplex type I glycoprotein D (Liu et al., 2002). The use
of basic-peptide complexation is also applicable to the
analysis of a complex low molecular weight heparin fraction that consisted of hexamers and octamers with varying
numbers of sulfate groups (Sturiale, Naggi, & Torri, 2001).
d. MALDI Analysis of Protein-Sulfated Oligosaccharide
Complexes. The addition of a heparin decasaccharide to
solutions of fibroblastic growth factor (FGF) types 1 or 2
results in marked changes to the MALDI-TOF mass
spectra obtained from those proteins (Venkataraman et al.,
1999a). It was possible to detect complexes between the
protein and decamer, and no such complexes were observed
for a heparin hexamer. Whereas FGF-1 forms 1:1 and 2:1
protein:decasaccharide complexes with MALDI-TOF
MS, FGF-2 forms 1:1, 2:1, 3:1, and 4:1 complexes under
the same conditions. HLGAG tetrasaccharides have been
shown to be sufficient to induce FGF-1 dimerization, and
the effects of different protein-ligand ratios on complex
formation have been shown (Sturiale, Naggi, & Torri,
2001). It has recently been demonstrated with MALDITOF MS that FGF-1 dimerization requires only one 6-Osulfate group heparin tetrasaccharide (Guerrini et al.,
2002). Those results demonstrate the utility of MALDI
for the analysis of growth factor-heparin complexes, and
produce important new data to provide insight on FGF
biochemistry.
Figure 31 shows a MALDI-TOF MS study of the
stoichiometry of the binding of a growth factor (fibroblastic
growth factor type 2, FGF2) with its receptor (FGF receptor, FGFR2) as a function of the presence of HLGAGs
(Kwan et al., 2001). FGF, FGFR2, and a heparin lyasederived oligosaccharide were mixed in 10 mM sodium
phosphate buffer, and were allowed to equilibrate at 48C for
30 min. The sample was spotted onto the MALDI target
with 1 mL sinapinic acid solution (saturated in 50%
acetonitrile). After drying, the sample was washed with
water. In the absence of HLGAG, FGF and FGFR2 both
form complexes with a 1:1 stoichiometry (Fig. 31a). In the
presence of HLGAG, a detectable 2:2:1 FGF:FGFR2:
HLGAG complex is formed, indicating that the oligosaccharide mediates formation of a unique binding stoichiometry (Fig. 31b). The inset shows the effect of the addition
204
FIGURE 31. Competitive binding of dFGF2 to FGFR2. (A) MALDI
mass spectrometric profile of a mixture of wild-type FGF2 and the
ectodomain of FGFR2. Observed in the mass spectrum are the [M þ H]þ
ions for an FGF2 dimer (m/z 30,214) and trimer (m/z 45,132), and FGFR2
monomer (m/z 24,888) and dimer (m/z 49,572), and a 1:1 FGF2FGFR2
complex (m/z 39,896). The theoretical molecular masses for FGF2 and
FGFR2 are 15,114 and 24,864 Da, respectively. (B) Mass spectrum of the
FGF2/FGFR2 mixture in the presence of a homogeneous HLGAG
decasaccharide. Addition of a decasaccharide (Deca) to FGF2/FGFR2
promoted the formation of a 2:2 FGF2FGFR2 complex with an observed
[M þ H]þ ion at m/z 82,750 (with the decasaccharide) or m/z 79,872
(without the decasaccharide). The [M þ H]þ ions for two dimeric FGFR2
species were also observed. The first (at m/z 49,592) represents the apo
complex, and the second (at m/z 52,474) is a 2:1 FGFR2decasaccharide
complex. Inset: Mass spectrum of dFGF2 added to the decasaccharide/
FGF2/FGFR2 mixture shown above. Three high molecular mass
complexes were observed: a 2:2 FGF2FGFR2 complex with or without
the decasaccharide and a 1:2 dFGF2FGFR2 complex without the
decasaccharide. Reprinted with permission from Kwan et al. (2001).
Copyright 2001 American Society for Biochemistry and Molecular
Biology.
of a genetic construct of FGF (dFGF) that corresponds to a
dimerized molecule with a histidine tag. The dFGF construct binds to two molecules of FGFRs in the presence of
the HLGAG, and forms a peak at slightly higher mass than
observed for the 2:2:1 FGF:FGFR2:HLGAG complex. The
data were used to support the conclusion that HLGAGs
facilitate FGF and/or FGFR2 oligomerization. The results
show that MALDI-TOF can be used to determine multiprotein binding stoichiometries. It is interesting that the
matrix conditions would be expected to weaken complexes
and that, nonetheless, complexes are observed. Although it
MASS SPECTROMETRY OF OLIGOSACCHARIDES
may be possible to obtain more biologically relevant results
with nano ESI and aqueous, neutral pH conditions, such
an experiment is more technically challenging than the
MALDI-TOF MS.
3. ESI of Sulfated Oligosaccharides
Desalted CS oligosaccharides have been analyzed from a
0.5 nM ammonium acetate solution with negative mode
ESI (Takagaki et al., 1992), where the most abundant
charge state for the oligosaccharide ions corresponded to
one charge per sulfate group. For example, an abundant
[M–4H]4 ion was observed for the tetrasulfated CS
octamers with [M–3H]3 and [M–5H]5 ions in much
lower abundances. Ions that correspond to losses of SO3
were in very low abundances, and it was possible to analyze
up to CS tetradecasaccharides. The same conditions were
used to obtain negative ESI of disulfated Lewisx trisaccharides, which results in abundant [M(Na)–H] and [M–
2H]2 ions with no losses of SO3 (Ii et al., 1995). Heparin
oligosaccharides generated by heparin-lyase digestion
have been analyzed with negative ESI as ammonium salts
(Chai et al., 1998). Gel filtration-purified oligosaccharides
were infused in organic solvent with or without 0.5 mM
ammonium bicarbonate. In the absence of ammonium ions,
the mass spectra of heparin oligosaccharides were quite
complex due to multiple adduction with sodium ions. In the
presence of ammonium ions, the profiles were much less
complex, due to a marked reduction in the abundances of
ions that correspond to sodium adducts of the oligosaccharides. It was possible to acquire data on heparin
fractions up to a decasaccharide size and of CS oligosaccharides up to tetradecamer size (Beeson et al., 1998; Chai,
Beeson, & Lawson, 2002).
Negative-ion ESI mass spectra were acquired on chromatographically purified oligosaccharides derived from
acharan sulfate, a glycosaminoglycan from the giant
African snail that is related to heparin and HS (Kim et al.,
1998). The spectra were acquired from 50% acetonitrile,
0.05% ammonium hydroxide, and showed abundant ions
that correspond to [M–2H]2 and [M(Na)–H]2 for the
disulfated tetrasaccharide fraction. The degree of sodium
adduction was observed to decrease as the concentration of
the analyte was decreased, and to increase with the size of
the analyte, and data were acquired on up to octasaccharides. The same conditions have been used to analyze
synthetic heparin oligosaccharides (Yu et al., 2000) and
those derived from dermatan sulfate (Yang et al., 2000).
Nano-ESI MS (Wilm & Mann, 1994) has been applied to
the analysis of synthetic heparins, using a modification of
the above conditions (Pope et al., 2001). The oligosaccharides were sprayed from a solution of 0.2% ammonium
hydroxide, 0.2% imidazole, and 50% acetonitrile from
pulled glass capillaries, using a titanium wire to provide
&
electrical contact. The results showed that abundant deprotonated ions are produced; those ions that corresponded
to sodium, potassium, or imidazole adduction had a lower
abundance.
Although it is clear that sulfated oligosaccharides may
be ionized without in-source fragmentation with negative
ESI, the use of basic peptides is an alternative method.
Polysulfated carbohydrates such as sucrose octasulfate
were mixed with a basic peptide in 1:5 (carbohydrate:peptide) molar ratios, and negative-ion ESI mass spectra were
acquired from 50% acetonitrile solutions (Siegel et al.,
1997). The results for sucrose octasulfate in the absence
of basic peptide showed a complex pattern of ions that
resulted from varying degrees of sodiation and desulfation.
In the presence of basic peptide, abundant ions that corresponded to [M þ peptide-4H]4 and [M þ peptide–
3H]3 ions were observed with natriated ions in very
low abundance and no in-source fragmentation of sulfate
groups. A number of peptides that contained a high percentage of lysine and arginine residues were found to form
complexes, the choice of which was found empirically for
each polysulfated carbohydrate tested.
Recently, the analysis of sucrose octasulfate (SOS) has
been accomplished in the positive ESI mode through the
use of counterions in the spray solvent (Gunay et al., 2003).
The results show that, although a clear [M(8Na) þ Na]þ ion
is observed in the positive mode with no fragmentation,
extensive fragmentation was observed for the [M(7Na)–
H] ion in the negative mode. Note that, using this nomenclature, the protons displaced by the cation in parentheses
is understood, see Section II.B. Pairing of SOS with
tetramethylammonium cation resulted in the detection of
[M(8(CH3)4N) þ (CH3)4N]þ, in which all eight sulfate
groups are deprotonated and paired with a cation, and an
additional cation provides the net positive charge. In the
negative mode, an ion that corresponds to [M(7(CH3)4N)–
H] was detected in addition to abundant fragment ions.
Similar trends were observed for SOS paired with tetrapropylammonium and tetramethylphosphonium cations.
The best results for ionization of SOS were obtained in
10 mM tetraethylammonium hydroxide, resulting in the
observation of a single ion that corresponded to [M(8R) þ
R]þ in the positive mode and [M(7R)–H] in the negative
mode, where R ¼ tetraethylammonium cation. The sensitivity was 10-fold higher in the positive mode. Those
results demonstrate that sulfate groups may be stabilized
with cations to allow their detection as singly charged ions
with ESI. Although it is not clear at this time why more
fragmentation is observed in the negative mode, it will be
important to study the fragmentation patterns that result
from CID of those ions.
The use of ESI to analyze of polysulfated oligosaccharide is recommended because it results in little or no insource fragmentation—provided that a low desolvation
205
&
ZAIA
energy is used. MALDI is considerably more energetic,
and is, therefore, applicable to molecules for which the
sulfate groups are relatively stable: negatively charged
monosulfated oligosaccharides or sodium-adducted disulfated oligosaccharides (Wheeler & Harvey, 2001).
4. On-Line Separation Systems
for Sulfated Carbohydrates
An interface has been optimized to allow direct capillary
electrophoresis measurements on oligosaccharides produced from enzymatic digestion of heparins (Duteil et al.,
1999). Separations were performed using either ammonia/
ammonium acetate pH 9.2 or acetic acid/ammonium
acetate pH 3.5; both are compatible with ESI. The two
conditions resulted on a reversal of the elution order for a
series of 8 heparin disaccharides. Capillary electrophoresis
was coupled with negative ESI by using a sheath liquid of
2 mM triethylamine, 3 mM ammonium formate in 50%
acetonitrile, resulting in the observation of abundant deprotonated protonated molecule ions, no in-source fragmentation, and low-abundance triethylamine adducts.
On-line CE/MS data were acquired for heparin disaccharide and tetrasaccharide mixtures from a heparin-lyase
digestion. Pressure-assisted capillary electrophoresis has
been interfaced with an ESI ion trap mass spectrometer,
using a sheath liquid of 5 mM formic acid pH 3.20 in
2-propanol (Ruiz-Calero et al., 2001). The capillary
electrophoresis electrolyte consisted of 30 mM formic
acid, pH 3.2, and the separation was assisted with 0.5 psi of
nitrogen gas. Using those conditions, it was possible to
resolve the eight heparin disaccharides in a 13-min electrophoretic separation, and it was possible to analyze disaccharides produced from enzymatic depolymerization of
heparins. Off-line CE has been used to separate oligosaccharides produced from the chondroitinase digestion of
chondroitin sulfate samples for ESI MS and MS/MS
analysis (Zamfir et al., 2002).
On-line LC/MS has been used to analyze sulfated
N-linked oligosaccharides, using a graphitized carbon
column (Kawasaki et al., 2001). The oligosaccharides were
eluted with a water/acetonitrile gradient in the presence of
5 mM ammonium acetate, and were detected with negative
or positive ESI. An ion-pairing reversed-phase chromatographic method has recently been described for on-line LC/
MS of heparins as well as other acidic carbohydrates
(Kuberan et al., 2002). Dibutylamine is used in the mobile
phase in the presence of acetic acid, where dibutyammonium ions pair with the negatively charged carbohydrates
to allow fractionation on a C-18 column. A mixture of
deprotonated and dibutylammonium-adducted ions were
observed with these conditions to analyze a heparin pentasaccharide. Significantly, no sulfate losses were observed.
That technique seems appropriate for on-line LC/MS
206
determination of sulfated oligosaccharide masses, and is
appropriate for miniaturization to maximize sensitivity. It
should be noted that this method is most appropriate when
an instrument can be dedicated to HLGAG analysis due to
difficulties in removing the ion-pairing agent.
Size-exclusion chromatography (SEC) has been used
on-line with ESI MS to desalt and separate mixtures of
sulfated GAG oligosaccharides (Zaia & Costello, 2001). A
commercial column (Amersham Biosciences Superdex
Peptide) is available that separates di-, tetra-, hexa-, and
octa-saccharides in a 20-min time-period. The method is
quite convenient because contaminating salts are effectively removed; at the same time, the complexity of the
oligosaccharide mixture that enters the instrument at any
moment is limited. Solvent conditions of 0.1 M ammonium
acetate, 10% methanol have been found to work for all
types of sulfated GAG oligosaccharides.
C. Tandem MS of Sulfated Oligosaccharides
1. Lessons from CID of Sulfated Peptides
High-energy tandem mass spectrometric analysis of
tyrosine-sulfated peptides as [M–H] ions with FAB
ionization results in abundant product ions from peptidebackbone cleavage and an ion that corresponds to [M–
SO3 –H] in moderate abundance (Gibson & Cohen,
1990). A sulfated an ion is observed that definitively places
the sulfated tyrosine group in the context of the peptide
sequence. It was also reported that losses of SO3 from the
precursor are relatively more abundant than losses of HPO3
from peptides that contain phosphotyrosine. Using positive
FAB ionization, ions produced from the loss of all sulfate
groups form the base peaks in CID mass spectra generated
from peptides that contain multiple tyrosine-sulfate residues (Yagami, Kitagawa, & Futaki, 1995). In the negative
FAB mode, a trisulfated peptide produced a ladder of ions
that corresponded to [M–H], [M–SO3 –H], and [M–
2(SO3)–H] with the absence of a [M–3(SO3)–H] ion.
Those results are consistent with the conclusion that protonated sulfate groups are significantly more labile than
deprotonated sulfate groups. Positively charged ions generated from phosphopeptides undergo minimal losses of
HPO3 during CID on a Q-oTOF mass spectrometer with
ESI (Nemeth-Cawley, Karnik, & Rouse, 2001). Under the
same conditions, sulfated peptides undergo losses of SO3;
the number of losses correspond to the number of sulfate
groups that modify the peptide. By increasing the collision
energy, it was possible to remove all sulfate groups and to
induce fragmentation of the peptide backbone. The method
is useful to differentiate isobaric phosphate and sulfate
groups based on CID fragmentation. Carbohydrates are
known to undergo backbone fragmentation at lower collision energies than peptides, and the challenge is to find
MASS SPECTROMETRY OF OLIGOSACCHARIDES
conditions where such ions are formed in preference over
losses of SO3 so that as much structural information as
possible can be produced from the tandem mass spectra.
2. Tandem MS of Mono- and
Di-Sulfated Oligosaccharides
From the discussion of ionization of sulfated oligosaccharides (see Section VI.B), the analysis of those molecules is
best carried out with negative ESI. Using that technique,
sulfated N-linked oligosaccharides are likely to ionize
with one charge per sulfate group, a condition where those
groups are least labile. Using negative ESI, CID mass
spectra of monosulfated N-linked oligosaccharides show
product ions that always contain the sulfate group. For a
biantennary core fucosylated lactosamine structure, the
product-ion pattern was consistent with sulfation of one of
the two GlcNAc residues, as shown in Figure 32 (Geyer
&
et al., 2001). The results show a relatively complete series
of Bm and Yn ions. Although the B5 ion is the most abundant
in the spectrum, other cleavages to the reducing side of
GlcNAc are absent. This result is consistent with the
proton-mediated weakening of the HexNAc glycosidic
bond for protonated ions. The absence of mobile protons
results in the stabilization of those linkages, and allow other
cleavages to occur in the MS2 stage.
Sulfoglycolipids have been analyzed in negative- and
positive-mode MS. Positive-mode product-ion mass spectra were not useful because the abundant losses of SO3 from
the precursor ion resulted in the loss of information on the
position of sulfation. In the negative mode, however, all
fragment ions contained the SO3 group and were thus useful
for structural determination (Kushi, Handa, & Ishizuka,
1985; Egge & Peter-Katalinc, 1987). Useful fragmentation
was also obtained from CID of [M–H] ions generated
from disulfated glycosphingolipids.
FIGURE 32. Negative-ion nano-ESI-IT-MS spectra of a sulfated N-linked glycan present on neural cell
adhesion molecule from newborn mouse brain. (A) MS2 profile of the [M–H] ion at m/z 1944.0; (B) MS3
profile of the Y5b ion at m/z 1781.7 (Geyer et al., 2001).
207
&
ZAIA
3. Precursor-Ion and Neutral-Loss Scans
for Sulfated Glycoconjugates
The sphingolipid sulfatide consists of a 3-sulfogaltosyl
group linked to a ceramide nucleus. Product-ion tandem
mass spectra of these species with negative ESI showed
the formation of structurally significant product ions in the
absence of losses of SO3. In addition, an ion at m/z 97,
[HSO4], was observed in all sulfatide product-ion spectra
(Marbois et al., 2000). Precursor-ion scans for m/z 97 were
used to identify ions produced from brain extracts that
contained sulfate and that were likely to correspond to
sulfatide (Hsu, Bohrer, & Turk, 1998). Precursor-ion scans
have been used to identify and quantify glycosphingolipids
from kidney tissue (Sandhoff et al., 2002). Significantly,
the m/z 97 ion characterizes O-sulfate groups and is absent
for sulfone groups (giving rise to .SO3, m/z 80). Quantification was linear over more than two orders of magnitude, a
range that was 70-times greater than that for densitometric
scanning.
Precursor-ion scanning has been used in the structural
analysis of sulfated N-linked oligosaccharides (Kawasaki
et al., 2001). In that work, a complex mixture of N-linked
oligosaccharide alditols released from erythropoietin was
analyzed with negative-ion triple-quadrupole ESI LC/MS
with graphitized carbon chromatography. Sulfated oligosaccharides were identified from mass chromatograms
of m/z 1035 (sulfated triantennary, trisialylated, 3-), m/z
1254 (sulfated tetraantennary tetrasialylated, 3-), m/z 1281
(disulfated tetraantennary, tetrasialylated, 3-), and other
sulfated N-linked ions. The results provide useful information on the presence of sulfated species in the mixtures,
and show the value of negative-ion ESI for the analysis
of sulfated oligosaccharides. Because the sulfate groups
are charged, they are stable and can be analyzed with
confidence.
Sulfated peptides generated by the tryptic digestion of
recombinant factor VIII produce abundant doubly charged
ions in positive-ESI reversed-phase LC/MS analyses
(Severs et al., 1999). Detected m/z values for several ions
were consistent with modification with up to three sulfate
groups. Scans for precursor ions that lose SO3 (80 Da) were
used to identify peptides that eluted from the column that
are modified with sulfate. Because the ions were doubly
charged, this identification entailed scanning for a 40 u
neutral loss.
4. Determination of Positional Sulfation
Isomers in GAG Disaccharides
Negative-ion FAB CID mass-analyzed ion kinetic-energy
spectra (MIKES) have been used to analyze monosulfated
disaccharides generated from CS (Lamb et al., 1992). As
summarized in Table 8, the results show that [M–H] ion
208
generated from those molecules undergo fragmentation to
produce abundant ions from glycosidic-bond and crossring cleavages with no ions that correspond to losses of
SO3. The residue location of the sulfate group could easily
be determined by the m/z values of the ions produced
from glycosidic-bond cleavage. Mass spectrometric differentiation of positional sulfation isomers was demonstrated for [M–H] ions of isomeric disaccharides that
differ only in the position of sulfation on the GalNAc
residues, DUA(1,3)GalNAc4S and DUA(1,3)GalNAc6S.
The MIKES spectra showed that the 4-sulfated isomer
produced an abundant ion at m/z 300 that corresponded to
Y1, and that an abundant ion at m/z 282 that corresponded to
Z1 was produced from the 6-sulfated disaccharide. The
product-ion spectra are very distinct in that the Z1 ion is
absent for the 4-sulfated isomer and the Y1 ion is absent for
the 6-sulfated isomer.
MIKES spectra of [M–H] ions generated from
DUA2S(1,3)GalNAc and DUA2S(1,4)GlcNAc were very
similar. Spectra of DUA(1,3)GalNAc6S and DUA(1,4)GlcNAc6S were quite distinct, the latter produced an abundant
Y1 ion in addition to cross-ring cleavage ions that are
absent in spectra of the former. It is significant that MIKES
spectra of the [M(Na)–H] ions generated from those
disaccharides produced different product-ion patterns.
DUA(1,3)GalNAc4S and DUA(1,3)GalNAc6S both produced abundant Z1 and Z1(Na) ions, and the two isomers
differ in the extent to which cross-ring fragmentation
occurs.
Positive ions generated from sulfated GAG disaccharides with FAB have also been analyzed with tandem MS (Ii
et al., 1994). As summarized in Table 9, for [M(2Na) þ H]þ
ions, it was not possible to distinguish DUA(1,3)GalNAc4S
from DUA(1,3)GalNAc6S based on the product-ion pattern. The differences between DUA2S(1,3)GalNAc and the
other two isomers was not very distinct due to abundant
losses of SO3 from B1 and C1 ions. Those three compounds
were easily distinguished, however, from tandem mass
spectra generated from [M(3Na) þ H]þ ions. An ion that
corresponded to Z1(2Na) was much more abundant in the
spectrum of DUA(1,3)GalNAc6S than in DUA(1,3)GalNAc4S, and abundant B1(3Na) and C1(3Na) ions distinguished the DUA2S(1,3)GalNAc isomer. Abundant losses
of 98 u (H2SO4) were observed from the latter ions.
The disulfated isomers DUA2S(1,3)GalNAc4S and
DUA2S(1,3)GalNAc6S could not be distinguished from
CID of [M(3Na) þ H]þ ions, and losses of SO3 from
the Y1(3Na) made it difficult to distinguish the DUA(1,3)GalNAc4S,6S isomer (see Table 9). All three disulfated
isomers, however, were easily distinguished by CID of the
[M(4Na) þ H]þ ions.
CID product-ion spectra have proven to be generally
useful to determine the number of sulfate groups that
modify heparin disaccharides (Ii et al., 1995b). In
MASS SPECTROMETRY OF OLIGOSACCHARIDES
&
TABLE 8. Summary of product ions generated from monosulfated GAG disaccharides that are useful for distinguishing
isomers
a
Ions analyzed by MIKES.
Spectra of the isomeric compounds were not distinguishable.
c
Ions analyzed by CID MS/MS.
d
Ion contains sulfate and is generated by ring cleavage.
e
Ions produced from subsequent fragmentation of Z1.
f
[Na2HSO4]þ.
g
[Na2HSO4]þ.
b
particular, it is possible to differentiate DUA2S(1,4)GlcN
from DUA(1,4)GlcNS and DUA(1,4)GlcN6S by CID of
[M(2Na) þ H]þ ions. As Table 9 shows, the information
obtained depends on the number of sodium adducts in the
precursor ion. In general, the B1 and C1 ions are observed in
low relative abundance, but are useful to determine whether
the DUroA residue is sulfated. Cross-ring cleavage ions
0,2
A2 and 2,4A2 are often abundant in the positive-ion CID
tandem mass spectra of disulfated heparin disaccharides,
are useful to assign the position of sulfation on the GlcNAc
residue, and may be useful in the case of a rare GlcNS3S or
GlcNAc3S sulfation pattern.
5. Tandem Mass Spectrometric Quantification
of GAG Disaccharides
A stereochemical differentiation of hexosamine monosaccharide residues has been demonstrated with ESI QIT
MS (Desaire & Leary, 1999b). In that work, it was demonstrated that cobalt–diaminopropane complexes of the
hexosamine monosaccharides glucosamine, galactosamine, and mannosamine produced unique product-ion
spectra. The relative abundances of ions produced from
pure standards were characterized, and were used to
quantify two- and three-component mixtures. Using that
209
method, an adaptation of the superposition principle of
quantifying oil distillates (Roboz, 1968), a system of equations is developed, one for each mixture component. That
method has been applied to the quantification of CS
disaccharides from mixtures derived from a chondroitinase
digestion (Desaire & Leary, 2000). DUroAGalNAc4S
(DDi-4S) produces an abundant Y1 ion at m/z 300, and
DUroAGalNAc6S (DDi-6S) produces an abundant Z1 ion
at m/z 282 (Linhardt et al., 1992). Using ESI QIT,
DUroA2SGalNAc (DDiUA2S) was observed to produce
an ion at m/z 237 that corresponded to B1 (Desaire, Sirich,
& Leary, 2001). That disaccharide was used as an internal
standard for the quantification of mixtures of DDi-4S and
DDi-6S produced from the digestion of CS chains with
chondroitinase enzymes. For that purpose, CID production mass spectra were first collected individually for the
three pure disaccharides under the same instrumental
conditions. The CID product-ion spectrum of a 1:1:1
DDiUA2S:DDi-4S:DDi-6S standard mixture was acquired,
and the per cent total ion abundances for m/z 237, 282, and
300 were determined. The abundance of each product ion
was expressed as contributions from each of the three
standard disaccharides, as shown in Table 10 and Figure 33.
Normalization factors (A, B, C) were calculated by
substituting the observed percent total ion abundances for
m/z 237, 282, and 300 in the equation. Internal standard
DDiUA2S was added to unknown mixtures of DDi-4S and
DDi-6S, the CID product-ion tandem mass spectra were
acquired, and the percent total ion abundances for m/z 237,
TABLE 10. Contributions of the three distinguishing product ions
for analysis of CS disaccharides, m/z 237, 282, and 300 [reprinted
with permission from Desaire, Sirich, & Leary (2001), Copyright
2001 American Chemical Society]
b
a
Loss of SO3 and a neutral fragment, corresponding to 0,2X0 (C4H7NO2), from the precursor ion.
[OHCCH2OSO3].
ZAIA
TABLE 9. Summary of product ions generated from disulfated GAG disaccharides that are useful for distinguishing isomers
&
a
The contribution is an average of four measurements.
210
MASS SPECTROMETRY OF OLIGOSACCHARIDES
FIGURE 33. System of equations used to quantify disaccharides.
Reprinted with permission from Desaire, Sirich, & Leary (2001).
Copyright 2001 American Chemical Society.
282, and 300 were substituted in the same system of
equations. The values calculated for A, B, and C were normalized with those values calculated from a 1:1:1 mixture
to produce the ratios of the three disaccharides present in
the mixture. That method was validated with a series of
standard mixtures, and used to quantify disaccharides produced from enzyme digests (Desaire, Sirich, & Leary,
2001).
Negative-ion FAB tandem mass spectrometry has been
used to structurally analyze CS tetrasaccharide tri- and
tetrasulfated monoclonal antibody antigenic determinants
as potassium salts (Ii et al., 1995a). That work has been
extended to negative-ionization ESI with QIT MS (Desaire
& Leary, 2000; Desaire, Sirich, & Leary, 2001), triplequadrupole (Zaia & Costello, 2001), QoTOF (Zaia,
McClellan, & Costello, 2001), and FT (McClellan et al.,
2002) instruments to show that the tandem mass spectra
can be used to determine sulfation ratios in mixtures produced by chondroitin lyases without any derivatization or
purification.
6. Tandem Mass Spectrometric Analysis
of GAG Oligosaccharides
It has been shown that CID MS/MS of alkali-adducted
chondroitin sulfate tetramers ionized by negative FAB
results in abundant glycosidic bond cleavages with no
losses of SO3 (Ii et al., 1995a). Heparin disaccharides that
contain three sulfate groups can be ionized as singly
charged negative ions, and fragmented by CID with
abundant glycosidic-bond and cross-ring cleavages with
minimal SO3 losses, provided that at least two sodium
adducts are present in the negative singly charged molecular ion (Ii et al., 1995b). Singly charged trisulfated
&
disaccharides can also be dissociated in the positive ionization mode with tetra-sodium adducts with low to moderate
losses of sulfate from the precursor ion (Li et al., 1995).
Those results, and data on sulfated peptides (Yagami et al.,
2000), support the conclusion that sulfate groups are least
labile when charged, and that dissociation of negatively
charged ions results in the greatest number of charged
sulfate groups with the fewest metal-cation adducts.
The fact that ESI results in more highly charged ions
than observed for FAB ionization or MALDI has important
implications for the analysis of highly sulfated carbohydrates. Heparin oligosaccharides generated by heparinlyase digestion can be ionized with negative ESI as
multiply charged deprotonated ions from solutions that
contain ammonium salts (Chai et al., 1998; Desaire &
Leary, 2000; McClellan et al., 2002) without SO3 losses
provided that mild desolvation conditions were used.
Abundant sodium-adducted ions are observed with negative ESI in the absence of ammonium ions (Chai et al.,
1998; Kim et al., 1998). Dissociation of singly charged
monosulfated (Ruiz-Calero et al., 2001) or doubly charged
disulfated heparin disaccharides (Pope et al., 2001) resulted in abundant ions from glycosidic-bond and crossring cleavages. Dissociation of a doubly charged, trisulfated, heparin disaccharide results in an abundant loss of
one SO3 molecule from the precursor, in addition to
glycosidic and cross-ring cleavages (Ruiz-Calero et al.,
2001). In summary, SO3 losses are minimized and
glycosidic-bond and cross-ring cleavages are maximized,
when the negative charge on the precursor ion equals the
number of sulfate groups.
a. CS Oligosaccharides. Negative ESI of CS oligosaccharides results in a mixture of charge states, the most
abundant corresponds to one charge per sulfate group
(Zaia, McClellan, & Costello, 2001). To ionize GAGs, it is
best to use 30% organic solutions modified with either
ammonium acetate or ammonium hydroxide at a low
percent (0.1%). Although the exact charge-state distribution will depend to some extent on the instrument-source
design, the most abundant ion will contain one charge per
sulfate group. For oligosaccharides produced by chondroitinase enzymes that act by eliminative mechanism
to produce a D-unsaturated residue at the nonreducing
terminus, the charge state with all sulfates charged corresponds to m/z 458. Thus, a mixture of those oligosaccharides will result in an overlapping isotopic cluster at that
value. Therefore, it is necessary to either separate the oligosaccharides prior to MS analysis, or to choose a different,
non-degenerate, charge state. CS oligosaccharides generated by testicular hyaluronidase, an enzyme that acts by a
hydrolytic mechanism, contain a saturated non-reducing
terminus, and do not produce degenerate m/z values. That
enzyme, however, contains transglycosylase activity, and
211
&
ZAIA
its use for sequence analysis of biological samples is not
advised.
Tandem mass spectrometric analysis of D-unsaturated
CS oligosaccharides for which all sulfates are charged
produces a product-ion pattern as shown in Figure 34.
Note that the abundance of the precursor ion was kept
high to minimize the extent to which multiple bond fragmentation occurs. The product-ions correspond to Bm and
Yn cleavages to odd-numbered glycosidic bonds, with no
cleavage to even-numbered bonds. Inspection of CID
product-ion mass spectra of mixtures of 4- and 6-sulfated
CS octamers demonstrates that the abundance of the Bm
ions depends on the position of GalNAc sulfation; see
Figure 35.
To gain further information on how product-ion
abundances vary according to the position of sulfation,
CID mass spectra were obtained on various 4- and 6-
FIGURE 35. Bar-graph representations of percent total ion abundances
for Bm ions produced from [M–4H]4 ions CSA/CSC octasaccharides
mixtures, ratios as given: (a) tetrasulfated octamer; (b) sodium
borodeuteride-reduced tetrasulfated octamer. For symbol definitions,
see Figure 34. Reprinted with permission from Zaia, McClellan, &
Costello (2001). Copyright 2001 American Chemical Society.
FIGURE 34. Negative-ion ESI CID product-ion spectra acquired at
collision energy 17.5 V, (a) CSA tetrasulfated octamer [M–4H]4 and
(b) CSC tetrasulfated octamer [M–4H]4. The symbolic structure and
product-ion cleavages are shown in (c). Reprinted with permission from
Zaia, McClellan, & Costello (2001). Copyright 2001 American
Chemical Society.
212
sulfated CS oligosaccharide mixture percentages. As
shown in Figure 35, the abundance of Bm ions diminishes
with the percent of 4-sulfated oligosaccharides in the
mixture; those data are consistent with the abundances of
those ions that indicate the sulfation position of adjacent
GalNAc residues (Zaia, McClellan, & Costello, 2001).
Evidently, the glycosidic bonds for 4-sulfated GalNAc
oligosaccharides are more labile than are those that are
6-sulfated, resulting in more-abundant Bm ions. Therefore,
it is expected that the relative abundances of those product
ions will be useful in determining the position of sulfation
of adjacent GalNAc residues. To prove this point, it will
be necessary to synthesize CS oligosaccharides with mixed
4- and 6-sulfation in defined positions.
The CS oligosaccharide product-ion pattern varies
markedly with charge state, as summarized in Figure 36
(McClellan et al., 2002). For CS hexamers, D(UroA)3
(GalNAc)3(SO3)3, CID of the 2 charge state results in Bm
and Yn ions from cleavage of only the even-numbered
glycosidic bonds. Those results are consistent with the
generation of information that is complementary to CID of
the 3 charge state for hexamers. The CID product-ion
MASS SPECTROMETRY OF OLIGOSACCHARIDES
&
FIGURE 36. The effect of charge state on the product-ion profile generated from CS oligosaccharides. The
top spectrum shows the CID profile obtained from SORI CID of the [M–3H]3 ion generated from a CS
hexamer, and the bottom shows that for the [M–2H]2 ion. The fragmentation diagrams are shown to the
right of each spectrum (McClellan et al., 2002).
profile for hexamer 1 ions results in even-numbered Bm
and Yn ions, with intramolecular sulfate transfer observed.
It appears that the ions with a 1 charge state may adopt a
compact gas-phase conformation that allows intramolecular sulfate transfer to occur. The 2 and 3 charge states
are likely to adopt an extended conformation, thus
separating potential acceptor sites for an intramolecular
sulfate transfer.
The observed product-ion pattern for ions in which all
sulfate groups are charged (top spectrum in Fig. 36) in
which fragmentation occurs to the reducing side of GlcA
residues is in contrast to the fragmentation observed for
deprotonated HexNAc-containing oligosaccharides, in
which abundant fragmentation is observed to the reducing
side of HexNAc residues (see Section II.C.2). Protonated
and alkali-cationized ions produced from oligosaccharides
have, in fact, been observed to undergo facile cleavages to
the reducing side of HexNAc residues in the positive mode
for native and peralkylated ions, see Sections II.C.1, II.C.3,
and II.D. The cases with low-energy CID, where facile
fragmentation of HexNAc residue is not observed, are
those in which a transition metal is tightly associated with
the reducing end to yield charge-remote fragmentation
(Konig & Leary, 1998). When CS oligosaccharide ions in
which one of the sulfate groups is protonated undergo CID
(bottom spectrum in Fig. 36), abundant fragmentation to
the reducing side of GalNAc residues is observed that
corresponds to even-numbered Bm and Yn ions (McClellan
et al., 2002). Those observations may be explained based
on the availability of mobile protons for ions in which
one of the sulfate groups is not charged. During the
multiple-collision low-energy CID process, it is likely that
the –SO3H proton migrates to a GalNAc nitrogen atom,
and is, therefore, in close proximity to the glycosidic bond
on the reducing side, destabilizing it to the point that it is
the most labile bond in the ion, resulting in preferential
production of even-numbered Bm and Yn ions. This result
is consistent with fragmentation of fully charged CS
oligosaccharides (all sulfate groups are deprotonated and
therefore charged) being a charge-remote process despite
the low-energy fragmentation regime. This process is
analogous to what is seen when transition metals are coordinated to an oligosaccharide, immobilizing charge and
resulting in fragmentation directed by the metal center
(Fura & Leary, 1993; Konig & Leary, 1998), where
HexNAc fragmentation is suppressed. Care must be taken
in setting the collision voltages for tandem mass spectrometric analysis of CS oligosaccharides to avoid overfragmentation of the molecules (Zamfir et al., 2002). The
multiply charged ions generated from GAG oligosaccharides are unique among the classes of biomolecules because
of the high density of charge per residue. This charge
density dramatically changes the CID properties such that
bonds are destabilized to approximately the same degree as
the molecular mass increases. Thus, the diminishing yield
of CID product ions that is observed for peptides is not
expected to limit the fragmentation of GAGs to so great a
degree.
213
&
ZAIA
b. Heparin/HS Oligosaccharides. Negative ESI has been
used to ionize heparin oligosaccharides for tandem mass
spectrometric analysis, and has been found to be useful in
quantification (Pope et al., 2001). Tandem mass spectrometric analysis has also been shown to produce abundant
glycosidic bond cleavages despite the high degree of
sulfation of heparins (Zaia & Costello, 2003). CID production mass spectra generated from the [M–6H]6 ion of an
octasulfated pentamer results in abundant ions from glycosidic bond cleavage—as shown in Figure 37 (Zaia &
Costello, 2003). The data demonstrate that abundant
glycosidic bond cleavages can be obtained, provided that
most of the sulfate groups are charged. Repulsions from
such charges are likely to destabilize glycosidic bond, and
to make their rupture energetically favorable with respect
to losses of SO3 from the precursor ion. Glycosidicbond cleavage is also favorable for [M(3Ca)–4H]4 ions
generated from the same oligosaccharide, as shown in
Figure 38 (Zaia & Costello, 2003). Note that the nomenclature used assumes the protons displaced by the metal
cations, and shows only the loss of protons that confer
charge on the ion (see Section V.B). More information is
produced from the [M(3Ca)–4H]4 CID mass spectrum
than from that of the [M–6H]6 ion in the form of
glycosidic-bond and cross-ring cleavages. The use of ESI
MS for the analysis of highly sulfated oligosaccharides is,
therefore, indicated from the point of view of obtaining
ions without in-source fragmentation, and of obtaining
tandem mass spectra that indicate patterns of bond cleavages that indicate sequence.
FIGURE 37. Product-ion tandem mass spectrum of the [M–6H]6 ion
generated from an octasulfated synthetic heparin pentamer, using
negative ESI QIT MS. Reprinted with permission from Zaia & Costello
(2003). Copyright 2003 American Chemical Society.
214
FIGURE 38. Negative-ion ESI tandem mass spectrum of a calcium-
complexed synthetic heparin pentamer [M(3Ca)–4H]4, acquired on a
Q-oTOF instrument. Reprinted with permission from Zaia & Costello
(2003). Copyright 2003 American Chemical Society.
D. Conclusions
The most important factor for the mass spectrometric
analysis of sulfated oligosaccharides is the charge environment of the sulfate groups. Sulfate groups are quite
labile when protonated, but become substantially less so
when either deprotonated or paired with a metal cation. The
decrease in lability is such that losses of SO3 from the
precursor ion, either in-source or during CID, become less
energetically favorable than cleavage of glycosidic bonds.
When ions formed correspond to one negative charge
per sulfate group, abundant intact protonated molecule
ions result, and abundant glycosidic bond cleavages
are observed in CID mass spectra. ESI results in the
observation of such charge states, and is, therefore, the
choice for analysis of sulfated oligosaccharides. For sulfated N- and O-linked oligosaccharides, the CID product
ions all contain the sulfate group, facilitating the identification of the modified residue. Although chondroitin sulfate and keratan sulfate GAGs contain more sulfate groups
per monosaccharide unit, the most abundant ions formed
with ESI correspond to one charge per sulfate. Because
oligosaccharides derived from those GAGs produce
abundant glycosidic-bond product ions with losses of
SO3 in very low abundances, the product-ion patterns are
readily interpreted in terms of the number of sulfate groups
on each monosaccharide residue. The abundance of the
MASS SPECTROMETRY OF OLIGOSACCHARIDES
glycosidic-bond product ions also indicate the position of
sulfation (GalNAc 4- or 6-) for CS oligosaccharides.
ESI MS is also useful for the analysis of the HLGAGs,
a class that contains the highest degree of sulfation of any
biomolecule. Provided that gentle desolvation conditions
are used in the source, intact deprotonated protonated
molecule ions can be observed with no losses of SO3. It is
not, however, possible to produce ions in which all sulfate
groups are deprotonated, due to charge-charge repulsion.
The high density of charge on these ions appears to destabilize glycosidic bonds, and to make their rupture energetically favorable with respect to losses of SO3. Therefore
deprotonated HLGAG ions produce both types of fragments that result from CID. Interpretation of those production profiles entails calculating neutral-mass equivalents for
each product ion, and identifying pairs that sum to the mass
of the neutral intact molecule. Such pairs will be produced
without losses of SO3 and will be useful for determining
structure. It is necessary to obtain CID MS on all observed
charge states, and to combine the useful structural data. An
alternative is to use metal cations to stabilize sulfate groups
without increasing the ionic charge density. Divalent calcium ions have been shown to be useful for that purpose,
producing abundant glycosidic bond as well as some crossring cleavages. That approach seems likely to be the most
generally useful because complete or nearly complete
fragmentation data are obtained in a single CID spectrum.
VII. OVERALL CONCLUSIONS
Although the analysis of complex oligosaccharides remains a challenge, recent research has outlined clear
approaches with modern mass spectrometry. Mixtures
released from a glycoprotein or other biological source
may be effectively profiled, provided that steps are taken to
make the oligosaccharides chemically equivalent, thereby
insuring equal ionization responses. Such steps include
chromatographic separation, peralkylation, and methyl
esterification. Peralkylated ions produce the most informative tandem mass spectra with regard to the branching
structure, by virtue of the unique mass created by bond
scission. For analysis of sample quantities below a few
micrograms, however, permethylation is not recommended
due to sample losses. Many researchers are thus investigating mass spectrometric analysis of native or minimally
derivatized samples. Although protonated native ions produce informative product-ion mass spectra, that precursorion form is not recommended because of the possibility of
internal-residue rearrangements.
At this time, direct interpretation of branched oligosaccharide tandem mass spectra does not seem feasible.
Rather, cleavages to labile residues in branced oligosaccharides (Neu5Ac, Fuc and HexNAc) expose core
&
structures; subsequent stages of fragmentation of those
structures provide useful structural fingerprints. Such data
will be most useful when compared to a library of tandem
mass spectra generated from known oligosaccharide structures. In that way, oligosaccharide substructures released
by gas-phase degradation in the mass spectrometer may be
identified by a comparision with library spectra. Although
the theoretical basis for that approach has been described in
a series of publications (Konig & Leary, 1998; Sheeley &
Reinhold, 1998; Viseux, de Hoffmann, & Domon, 1998;
Weiskopf, Vouros, & Harvey, 1998; Tseng, Hedrick, &
Lebrilla, 1999), more development is needed. Specifically,
bioinformatics tools must be developed to allow the facile
comparison of the tandem mass spectrometric data from
unknown and known structures. This process will entail a
collaboration among glycoscience research groups to
produce the libraries and to conduct detailed studies of
the elucidation of unknowns to establish the validity of the
gas-phase degradation and library-searching techniques.
The public availability of such libraries is likely to result
in a dramatic progress in the ability to correlate the expression of specific carbohydrate structures with biological
functions.
VIII. ABBREVIATIONS
CID
CS
dHex
ECD
ESI
FAB
FT
Fuc
GAG
Gal
GalNAc
Glc
GlcA
GlcNAc
Hex
HexA
DHexA
HexNAc
HS
HLGAG
IRMPD
KS
MALDI
Man
Neu5Ac
NeuGc
collision-induced dissociation
chondroitin sulfate
deoxyhexose residue
electron capture dissociation
electrospray ionization
fast atom bombardment
Fourier transform
fucose residue
glycosaminoglycan
galactose residue
N-acetylgalactosamine residue
glucose residue
glucuronic acid residue
N-acetylglucosamine
hexose residue
hexuronic acid residue
4,5-unsaturated hexuronic acid residue
N-acetylhexosamine residue
heparan sulfate
heparin-like glycosaminoglycan
infrared multiphoton dissociation
keratan sulfate
matrix assisted laser desorption/ionization
mannose residue
N-acetylneuraminic acid residue
N-glycolyl neuraminic acid
215
&
ZAIA
PSD
SDS–PAGE
SORI
TOF
QIT
post-source decay
sodium dodecyl sulfate–polyacrylamide
gel electrophoresis
selective off-resonance irradiation
time-of-flight
quadrupole ion trap
REFERENCES
Adams J. 1990. Charge-remote fragmentations—Analytical
applications and fundamental-studies. Mass Spectrom Rev
9:141–186.
Alving K, Paulsen H, Peter-Katalinic J. 1999. Characterization of
O-glycosylation sites in MUC2 glycopeptides by nanoelectrospray QTOF mass spectrometry. J Mass Spectrom 34:
395–407.
Asam MR, Glish GL. 1997. Tandem mass spectrometry of alkali
cationized polysaccharides in a quadrupole ion trap. J Am
Soc Mass Spectrom 8:987–995.
Bahr U, Pfenninger A, Karas M, Stahl B. 1997. High-sensitivity
analysis of neutral underivatized oligosaccharides by nanoelectrospray mass spectrometry. Anal Chem 69:4530–4535.
Baldwin MA, Stahl N, Reinders LG, Gibson BW, Prusiner SB,
Burlingame AL. 1990. Permethylation and tandem mass
spectrometry of oligosaccharides having free hexosamine:
Analysis of glycoinisitol phospholipid anchor glycan from
the scrapie prion protein. Anal Biochem 191:174–182.
Barr JR, Anumula KR, Vettese MB, Taylor PB, Carr SA. 1991.
Structural classification of carbohydrates in glycoproteins
by mass spectrometry and high-performance anion-exchange chromatography. Anal Biochem 192:181–192.
Bateman RH, Green MR, Scott G, Clayton E. 1995. A combined
magnetic sector-time-of-flight mass-spectrometer for structural determination studies by tandem mass-spectrometry.
Rapid Commun Mass Spectrom 9:1227–1233.
Bateman K, White R, Yaguchi M, Thibault P. 1998. Characterization of protein glycoforms by capillary-zone electrophoresis nanoelectrospray mass spectrometry. J Chromatogr
A 794:327–344.
Beeson JG, Chai W, Rogerson SJ, Lawson AM, Brown GV. 1998.
Inhibition of binding of malaria-infected erythrocytes by a
tetradecasaccharide fraction from chondroitin sulfate A.
Infect Immun 66:3397–3402.
Biemann K, Martin SA. 1987. Mass spectrometric determination
of the amino acid sequence of peptides and proteins. Mass
Spectrom Rev 6:1–76.
Bordoli RS, Howes K, Vickers RG, Bateman RH, Harvey DJ.
1994. Matrix-assisted laser-desorption mass-spectrometry
on a magnetic-sector. Instrument fitted with an array
detector. Rapid Commun Mass Spectrom 8:585–589.
Bourne EJ, Stacey M, Tatlow JC, Tedder JM. 1949. Studies on
trifluoroacetic acid. Part I. Trifluoroacetic anhydride as a
promotor of ester formation between hydroxy-compounds
and carboxylic acids. J Chem Soc 2976–2979.
216
Broberg S, Broberg A, Duus JO. 2000. Matrix-assisted laser
desorption/ionization time-of-flight mass spectrometry of
oligosaccharides derivatized by reductive amination and
N,N-dimethylation. Rapid Commun Mass Spectrom 14:
1801–1805.
Brüll LP, Heerma W, ThomasOates J, Haverkamp J, Kovacik V,
Kovac P. 1997. Loss of internal 1 ! 6 substituted monosaccharide residues from underivatized and per-O-methylated trisaccharides. J Am Soc Mass Spectrom 8:43–49.
Brüll LP, Kovacik V, Thomas-Oates JE, Heerma W, Haverkamp J.
1998. Sodium-cationized oligosaccharides do not appear to
undergo ‘internal residue loss’ rearrangement processes on
tandem mass spectrometry. Rapid Commun Mass Spectrom
12:1520–1532.
Burlingame AL, Boyd RK, Gaskell SJ. 1994. Mass spectrometry.
Anal Chem 66:634R–683R.
Cancilla MT, Penn SG, Lebrilla CB. 1998. Alkaline degradation
of oligosaccharides coupled with matrix-assisted laser
desorption/ionization Fourier transform mass spectrometry:
A method for sequencing oligosaccharides. Analytical
Chemistry 70:663–672.
Cancilla MT, Penn SG, Carroll JA, Lebrilla CB. 1996.
Coordination of alkali metals to oligosaccharides dictates
fragmentation behavior in matrix assisted laser desorption
ionization Fourier transform mass spectrometry. J Am Chem
Soc 118:6736–6745.
Cancilla MT, Wong AW, Voss LR, Lebrilla CB. 1999.
Fragmentation reactions in the mass spectrometry analysis
of neutral oligosaccharides. Anal Chem 71:3206–3218.
Carr SA, Huddleston MJ, Bean MF. 1993. Selective identification
and differentiation of N- and O-linked oligosaccharides in
glycoproteins by liquid chromatography-mass spectrometry. Protein Sci 2:183–196.
Carr SA, Reinhold VN. 1984. Structural characterization of
sulfated glycosaminoglycans by fast atom bombardment
mass spectrometry: Application to chondroitin sulfate.
J Carbohydr Chem 3:381–401.
Carr SA, Roberts GD. 1986. Carbohydrate mapping by mass
spectrometry: A novel method for identifying attachment
sites of Asn-linked sugars in glycoproteins. Anal Biochem
157:396–406.
Carr SA, Reinhold VN, Green BN, Hass JR. 1985. Enhancement
of structural information in FAB ionized carbohydrate
samples by neutral gas collision. Biomed Mass Spectrom
12:288–295.
Carr SA, Hemling ME, Folena-Wasserman G, Sweet RW,
Anumula K, Barr JR, Huddleston MJ, Taylor P. 1989.
Protein and carbohydrate structural analysis of a recombinant soluble CD4 receptor by mass spectrometry. J Biol
Chem 264:21286–21295.
Carr SA, Barr JR, Roberts GD, Anumula KR, Taylor PB. 1990.
Identification of attachment sites and structural classes of
asparagine- linked carbohydrates in glycoproteins. Methods
Enzymol 193:501–518.
Chai W, Beeson JG, Lawson AM. 2002. The structural motif in
chondroitin sulfate for adhesion of Plasmodium falciparum-
MASS SPECTROMETRY OF OLIGOSACCHARIDES
infected erythrocytes comprises disaccharide units of 4-Osulfated and non-sulfated N-acetylgalactosamine linked to
glucuronic acid. J Biol Chem 277:22438–22446.
Chai W, Kogelberg H, Lawson AM. 1996. Generation and
structural characterization of a range of unmodified chondroitin sulfate oligosaccharide fragments. Anal Biochem
237:88–102.
Chai W, Rosankiewicz JR, Lawson AM. 1995. TLC-LSIMS of
neoglycolipids of glycosaminoglycan disaccharides and of
oxymercuration cleavage products of heparin fragments that
contain unsaturated uronic acid. Carbohydr Res 269:111–
124.
Chai W, Luo J, Lim CK, Lawson AM. 1998. Characterization of
heparin oligosaccharide mixtures as ammonium salts using
electrospray mass spectrometry. Anal Chem 70:2060–
2066.
Chance DL, Mawhinney TP. 1996. Disulfated oligosaccharides
derived from tracheobronchial mucous glycoproteins of a
patient suffering from cystic fibrosis. Carbohydr Res 295:
157–177.
Charlwood J, Tolson D, Dwek M, Camilleri P. 1999. A detailed
analysis of neutral and acidic carbohydrates in human milk.
Anal Biochem 273:261–277.
Chen P, Werner-Zwanziger U, Wiesler D, Pagel M, Novotny MV.
1999. Mass spectrometric analysis of benzoylated sialooligosaccharides and differentiation of terminal alpha 2 ! 3
and alpha 2 ! 6 sialogalactosylated linkages at subpicomole levels. Anal Chem 71:4969–4973.
Ciucanu I, Kerek F. 1984. A simple and rapid method for the
permethylation of carbohydrates. Carbohydr Res 131:209–
217.
Claeys M, VandenHeuvel H, Chen S, Derrick PJ, Mellon FA,
Price KR. 1996. Comparison of high- and low-energy
collision-induced. Dissociation tandem mass spectrometry
in the analysis of glycoalkaloids and their aglycons. J Am
Soc Mass Spectrom 7:173–181.
Colangelo J, Licon V, Benen J, Visser J, Bergmann C, Orlando R.
1999. Characterization of the N-linked glycosylation site of
recombinant pectate lyase. Rapid Commun Mass Spectrom
13:2382–2387.
Cole RB. 1997. Electrospray ionization mass spectrometry:
Fundamentals, Instrumentation and Applications. New York,
NY: John Wiley, Inc.
Conboy JJ, Henion JD. 1992. The determination of glycopeptides
by liquid-chromatography mass-spectrometry with collision-induced dissociation. J Am Soc Mass Spectrom 3:804–
814.
Costello CE, Juhasz P, Perreault H. 1994. New mass spectral
approaches to ganglioside structure determinations. Prog
Brain Res 101:45–61.
Costello CE, Vath JE. 1990. Tandem mass spectrometry of
glycolipids. Methods Enzymol 193:738–768.
Dage JL, Ackermann BL, Halsall HB. 1998. Site localization of
sialyl Lewis(x) antigen on alpha1-acid glycoprotein by high
performance liquid chromatography-electrospray mass
spectrometry. Glycobiology 8:755–760.
&
Daniel PF. 1987. Separation of benzoylated oligosaccharides by
reversed-phase high-pressure liquid chromatography: Application to high-mannose type oligosaccharides. Methods
Enzymol 138:94–116.
Delaney J, Vouros P. 2001. Liquid chromatography ion trap mass
spectrometric analysis of oligosaccharides using permethylated derivatives. Rapid Commun Mass Spectrom 15:325–
334.
Dell A. 1990. Preparation and desorption mass spectrometry of
permethyl and peracetyl derivatives of oligosaccharides.
Methods Enzymol 193:647–660.
Dell A, Morris HR, Egge H, von Nicolai H, Strecker G. 1983a.
Fast-atom bombardment mass spectrometry for carbohydrate structure determination. Carbohydr Res 115:41–52.
Dell A, Oates JE, Morris HR, Egge H. 1983b. Structure
determination of carbohydrates and glycosphingolipids by
fast atom bombardment mass spectrometry. Int J Mass
Spectrom Ion Phys 46:415–418.
Dell A, Rogers M, Thomas-Oates J, Huckerby T, Sanderson P,
Nieduszynski I. 1988. Fast-atom-bombardment mass-spectrometric strategies for sequencing sulfated oligosaccharides. Carbohydr Res 179:7–19.
Dell A, Reason AJ, Khoo KH, Panico M, McDowell RA, Morris
HR. 1994. Mass spectrometry of carbohydrate-containing
biopolymers. Methods Enzymol 230:108–132.
Derappe C, Bauvy C, Lemonnier M, Lhermitte M, Platzer N,
Egge H, Peter-Katalinc J, van Halbeek H, Kamerling JP,
Vliegenthart JF. 1986. Determination of the structure of a
fucose-containing trisaccharide monophosphate isolated
from human pregnancy urine. Carbohydr Res 150:273–
284.
Desaire H, Sirich TL, Leary JA. 2001. Evidence of block and
randomly sequenced chondroitin polysaccharides: sequential enzymatic digestion and quantification using ion trap
tandem mass spectrometry. Anal Chem 73:3513–3520.
Desaire H, Leary JA. 1999a. Differentiation of diastereomeric Nacetylhexosamine monosaccharides using ion trap tandem
mass spectrometry. Anal Chem 71:1997–2002.
Desaire H, Leary JA. 1999b. Multicomponent quantification of
diastereomeric hexosamine monosaccharides using ion trap
tandem mass spectrometry. Anal Chem 71:4142–4147.
Desaire H, Leary J. 2000. Detection and quantification of the
sulfated disaccharides in chondoritin sulfate by electrospray
tandem mass spectrometry. J Am Soc Mass Spectrom
11:916–920.
Domon B, Müller DR, Richter WJ. 1989. Identification of
interglycosidic linkages and sugar constituents of larger
glycosides by tandem mass spectrometry. Organic Mass
Spectrom 24:357–359.
Domon B, Müller DR, Richter WJ. 1990. High performance
tandem mass spectrometry for sequence, branching and
interglycosidic linkage analysis of peracetylated oligosaccharides. Biomed Environ Mass Spectrom 19:390–392.
Domon B, Costello CE. 1988a. Structure elucidation of glycosphingolipids and gangliosides using high-performance
tandem mass spectrometry. Biochemistry 27:1534–1543.
217
&
ZAIA
Domon B, Costello CE. 1988b. A systematic nomenclature for
carbohydrate fragmentations in FAB-MS/MS spectra of
glycoconjugates. Glycoconjugate J 5:397–409.
Duffin KL, Welply JK, Huang E, Henion JD. 1992. Characterization of N-linked oligosaccharides by electrospray and
tandem mass spectrometry. Anal Chem 64:1440–1448.
Dufresne CP, Wood TD, Hendrickson CL. 1998. High-resolution
electrospray ionization Fourier transform mass spectrometry with infrared multiphoton dissociation of glucokinase
from bacillus stearothermophilus. J Am Soc Mass Spectrom
9:1222–1225.
Duteil S, Gareil P, Girault S, Mallet A, Feve C, Siret L. 1999.
Identification of heparin oligosaccharides by direct coupling
of capillary electrophoresis/ionspray-mass spectrometry.
Rapid Commun Mass Spectrom 13:1889–1898.
Egge H, Dell A, Von Nicolai H. 1983. Fucose containing
oligosaccharides from human milk. I. Separation and identification of new constituents. Arch Biochem Biophys 224:
235–253.
Egge H, Peter-Katalinc J. 1987. Fast atom bombardment mass
spectrometry for structural elucidation of glycoconjugates.
Mass Spectrom Rev 6:331–393.
Ernst B, Müller DR, Richter WJ. 1997. False sugar sequence ions
in electrospray tandem mass spectrometry of underivatized
sialyl-Lewis-type oligosaccharides. Int J Mass Spectrom Ion
Proc 160:283–290.
Ernst S, Rhomberg AJ, Biemann K, Sasisekharan R. 1998. Direct
evidence for a predominantly exolytic processive mechanism for depolymerization of heparin-like glycosaminoglycans by heparinase I. Proc Natl Acad Sci USA 95:4182–
4187.
Ethier M, Saba JA, Ens W, Standing KG, Perreault H. 2002.
Automated structural assignment of derivatized complex Nlinked oligosaccharides from tandem mass spectra. Rapid
Commun Mass Spectrom 16:1743–1754.
Fenn JB, Mann M, Meng CK, Wong SF. 1990. Electrospray
ionization—Principles and practice. Mass Spectrom Rev
9:37–70.
Flora JW, Hannis JC, Muddiman DC. 2001. High-mass accuracy
of product ions produced by SORI-CID using a dual
electrospray ionization source coupled with FTICR mass
spectrometry. Anal Chem 73:1247–1251.
Fournet B, Dhalluin JM, Strecker G, Montreuil J, Bosso C,
Defaye J. 1980. Gas-liquid chromatography and mass spectrometry of oligosaccharides obtained by partial acetolysis
of glycans of glycoproteins. Anal Biochem 108:35–56.
Fura A, Leary JA. 1993. Differentiation of Ca2þ-coordinated and
Mg2þ-coordinated branched trisaccharide isomers—An
electrospray-ionization and tandem mass-spectrometry
study. Anal Chem 65:2805–2811.
Gallego RG, Blanco JLJ, Thijssen-van Zuylen C, Gotfredsen CH,
Voshol H, Duus JO, Schachner M, Vliegenthart JFG. 2001.
Epitope diversity of N-glycans from bovine peripheral
myelin glycoprotein P0 revealed by mass spectrometry and
nano probe magic angle spinning H-1 NMR spectroscopy.
J Biol Chem 276:30834–30844.
218
Gangl ET, Annan MM, Spooner N, Vouros P. 2001. Reduction of
signal suppression effects in ESI-MS using a nanosplitting
device. Anal Chem 73:5635–5644.
Garozzo D, Giuffrida M, Impallomeni G, Ballistreri A,
Montaudo G. 1990. Determination of linkage position and
identification of the reducing end in linear oligosaccharides
by negative-ion fast atom bombardment mass-spectrometry.
Anal Chem 62:279–286.
Gaucher SP, Morrow J, Leary JA. 2000. STAT: A saccharide
topology analysis tool used in combination with tandem
mass spectrometry. Anal Chem 72:2331–2336.
Gaucher SP, Pedersen SF, Leary JA. 1999. Stereospecific
synthesis and characterization of aminoglycoside ligands
from diethylenetriamine. J Org Chem 64:4012–4015.
Gaucher SP, Leary JA. 1998. Stereochemical differentiation
of mannose, glucose, galactose, and talose using zinc(II)
diethylenetriamine and ESI-ion trap mass spectrometry.
Anal Chem 70:3009–3014.
Gaucher SP, Cancilla MT, Phillips NJ, Gibson BW, Leary JA.
2000. Mass spectral characterization of lipooligosaccharides from Haemophilus influenzae 2019. Biochemistry 39:
12406–12414.
Gauthier JW, Trautman TR, Jacobson DB. 1991. Sustained offresonance irradiation for collision-activated dissociation
involving Fourier transform mass spectrometry. Collisionactivated dissosiation technique that emulates infrared
multiphoton dissociation. Analytica Chimica Acta 246:
211–225.
Geyer R, Geyer H. 1994. Saccharide linkage analysis using
methylation and other techniques. Methods Enzymol 230:
86–108.
Geyer H, Bahr U, Liedtke S, Schachner M, Geyer R. 2001. Core
structures of polysialylated glycans present in neural cell
adhesion molecule from newborn mouse brain. Eur J Biochem 268:6587–6599.
Gibson BW, Cohen P. 1990. Liquid secondary ion mass
spectrometry of phosphorylated and sulfated peptides and
proteins. Methods Enzymol 193:480–501.
Gillece-Castro BL, Burlingame AL. 1990. Oligosaccharide
characterization with high-energy collision-induced dissociation mass spectrometry. Methods Enzymol 193:689–
712.
Guan S, Marshall AG, Wahl MC. 1994. MS/MS with high
detection efficiency and mass resolving power for product
ions in Fourier transform ion cyclotron resonance mass
spectrometry. Anal Chem 66:1363–1367.
Guerrini M, Agulles T, Bisio A, Hricovini M, Lay L, Naggi A,
Poletti L, Sturiale L, Torri G, Casu B. 2002. Minimal
heparin/heparan sulfate sequences for binding to fibroblast
growth factor-1. Biochem Biophys Res Commun 292:222–
230.
Gunay NS, Tadano-Aritomi K, Toida T, Ishizuka I,
Linhardt R. 2003. Evaluation of counterions for electrospray ionization mass spectral analysis of a highly sulfated
carbohydrate, sucrose octasulfate. Anal Chem 75:3226–
3231.
MASS SPECTROMETRY OF OLIGOSACCHARIDES
Gunnarson A. 1987. N- and O-alkylation of glycoconjugates and
polysaccharides by solid base dimethyl sulphoxide/alkyl
iodide. Glycoconj J 4:239–245.
Hakomori S-I. 1964. Rapid permethylation of glycolipid, and
polysaccharide catalyzed by methylsulfinyl carbanion in
dimethyl sulfoxide. J Biochem 55:205–208.
Hallgren P, Lundblad A. 1977. Structural analysis of nine
oligosaccharides isolated from the urine of a blood group O,
nonsecretor, woman during pregnancy and lactation. J Biol
Chem 252:1014–1022.
Handa S, Nakamura K. 1984. Modification of sialic acid carboxyl
group of ganglioside. J Biochem (Tokyo) 95:1323–1329.
Hanisch FG, Jovanovic M, Peter-Katalinic J. 2001. Glycoprotein
identification and localization of O-glycosylation sites by
mass spectrometric analysis of deglycosylated/alkylaminylated peptide fragments. Anal Biochem 290:47–59.
Hanrahan S, Charlwood J, Tyldesley R, Langridge J, Bordoli R,
Bateman R, Camilleri P. 2001. Facile sequencing of
oligosaccharides by matrix-assisted laser desorption/ionisation on a hybrid quadrupole orthogonal acceleration time-offlight mass spectrometer. Rapid Commun Mass Spectrom
15:1141–1151.
Harvey DJ. 1993. Quantitative aspects of the matrix-assisted
laser-desorption mass-spectrometry of complex oligosaccharides. Rapid Commun Mass Spectrom 7:614–619.
Harvey DJ. 2000a. Collision-induced fragmentation of underivatized N-linked carbohydrates ionized by electrospray.
J Mass Spectrom 35:1178–1190.
Harvey DJ. 2000b. Electrospray mass spectrometry and fragmentation of N-linked carbohydrates derivatized at the
reducing terminus. J Am Soc Mass Spectrom 11:900–915.
Harvey DJ. 2000c. N-(2-diethylamino)ethyl-4-aminobenzamide
derivative for high sensitivity mass spectrometric detection
and structure determination of N-linked carbohydrates.
Rapid Commun Mass Spectrom 14:862–871.
Harvey DJ. 2001. Ionization and collision-induced fragmentation
of N-linked and related carbohydrates using divalent cations.
J Am Soc Mass Spectrom 12:926–937.
Harvey DJ, Bateman RH, Green MR. 1997. High-energy
collision-induced fragmentation of complex oligosaccharides ionized by matrix-assisted laser desorption/ionization
mass spectrometry. J Mass Spectrom 32:167–187.
Harvey DJ, Naven TJP, Küster B, Bateman RH, Green MR,
Critchley G. 1995. Comparison of fragmentation modes for
the structural determination of complex oligosaccharides
ionized by matrix-assisted laser. Rapid Commun Mass
Spectrom 9:1556–1561.
Harvey DJ, Bateman RH, Bordoli RS, Tyldesley R. 2000.
Ionisation and fragmentation of complex glycans with a
quadrupole time-of-flight mass spectrometer fitted with a
matrix-assisted laser desorption/ionisation ion source.
Rapid Commun Mass Spectrom 14:2135–2142.
Harvey DJ, Mattu TS, Wormald MR, Royle L, Dwek RA, Rudd
PM. 2002. ‘‘Internal residue loss’’: Rearrangements occurring during the fragmentation of carbohydrates derivatized
at the reducing terminus. Anal Chem 74:734–740.
&
Heck AJR, Derrick PJ. 1997. Ultrahigh mass accuracy in isotopeselective collision-induced dissociation using correlated
sweep excitation and sustained off-resonance irradiation—
A Fourier transform ion cyclotron resonance mass spectrometry case study on the [M þ 2H](2þ) ion of Bradykinin.
Anal Chem 69:3603–3607.
Hettich RL, Stemmler EA. 1996. Investigation of oligonucleotide
fragmentation with matrix-assisted laser desorption/ionization Fourier-transform mass spectrometry and sustained offresonance irradiation. Rapid Commun Mass Spectrom 10:
321–327.
Hillenkamp F, Karas M, Beavis RC, Chait BT. 1991. Matrixassisted laser desorption/ionization mass spectrometry of
biopolymers. Anal Chem 63:1193A–1203A.
Hirayama K, Yuji R, Yamada N, Kato K, Arata Y, Shimada I.
1998. Complete and rapid peptide and glycopeptide mapping of mouse monoclonal antibody by LC/MS/MS using
ion trap mass spectrometry. Anal Chem 70:2718–2725.
Hofmeister GE, Zhou Z, Leary JA. 1991. Linkage position
determination in lithium-cationized disaccharides—Tandem mass-spectrometry and semiempirical calculations.
J Am Chem Soc 113:5964–5970.
Hofstadler SA, Wahl JH, Bakhtiar R, Anderson GA, Bruce JE,
Smith RD. 1994. Capillary electrophoresis Fourier transform
ion cyclotron resonance mass spectrometry with sustained
off-resonance irradiation for the characterization of protein
and peptide mixtures. J Am Soc Mass Spectrom 5:894–899.
Honda S, Akao E, Suzuki S, Okuda M, Kakehi K, Nakamura J.
1989. High-performance liquid-chromatography of reducing carbohydrates as strongly ultraviolet-absorbing and
electrochemically sensitive 1-phenyl-3-methyl-5-pyrazolone derivatives. Anal Biochem 180:351–357.
Hsu FF, Bohrer A, Turk J. 1998. Electrospray ionization tandem
mass spectrometric analysis of sulfatide. Determination of
fragmentation patterns and characterization of molecular
species expressed in brain and in pancreatic islets. Biochim
Biophys Acta 1392:202–216.
Huang Y, Pasa-Tolic L, Guan S, Marshall AG. 1994. Collisioninduced dissociation for mass spectrometric analysis of
biopolymers: High-resolution Fourier transform ion cyclotron resonance MS4. Anal Chem 66:4385–4389.
Huberty MC, Vath JE, Yu W, Martin SA. 1993. Site-specific
carbohydrate identification in recombinant proteins using
MALD-TOF MS. Anal Chem 65:2791–2800.
Huddleston MJ, Bean MF, Carr SA. 1993. Collisional fragmentation of glycopeptides by electrospray ionization LC/MS and
LC/MS/MS: Methods for selective detection of glycopeptides in protein digests. Anal Chem 65:877–884.
Hui JP, Lanthier P, White TC, McHugh SG, Yaguchi M, Roy R,
Thibault P. 2001. Characterization of cellobiohydrolase I
(Cel7A) glycoforms from extracts of Trichoderma reesei
using capillary isoelectric focusing and electrospray mass
spectrometry. J Chromatogr B Biomed Sci Appl 752:349–
368.
Håkansson K, Cooper HJ, Emmett MR, Costello CE, Marshall
AG, Nilsson CL. 2001. Electron capture dissociation and
219
&
ZAIA
Infrared multiphoton dissociation MS/MS of an N-glycosylated tryptic peptide to yield complementary sequence information. Anal Chem 73:4530–4536.
Ii T, Okuda S, Hirano T, Ohashi M. 1994. Tandem mass
spectrometry for characterization of unsaturated disaccharides from chondroitin sulfate, dermatan sulfate and hyaluronan. Glycoconj J 11:123–132.
Ii T, Ohashi Y, Nunomura S, Ogawa T, Nagai Y. 1995. Fast atom
bombardment and electrospray ionization tandem mass
spectrometry of sulfated Lewis(x) trisaccharides. J Biochem
(Tokyo) 118:526–533.
Ii T, Kubota M, Hirano T, Ohashi M, Yoshida K, Suzuki S. 1995a.
FAB CID-MS/MS characterization of tetrasaccharide triand tetrasulfate derived from the antigenic determinant
recognized by the anti-chondroitin sulfate monoclonal
antibody MO-225. Glycoconj J 12:282–289.
Ii T, Kubota M, Okuda S, Hirano T, Ohashi M. 1995b. Negativeion fast atom bombardment tandem mass spectrometry
for characterization of sulfated unsaturated disaccharides
from heparin and heparan sulfate. Glycoconj J 12:162–
172.
Itoh S, Kawasaki N, Ohta M, Hayakawa T. 2002. Structural
analysis of a glycoprotein by liquid chromatography-mass
spectrometry and liquid chromatography with tandem mass
spectrometry—Application to recombinant human thrombomodulin. J Chromatogr A 978:141–152.
Jacobs A, Dahlman O. 2001. Enhancement of the quality of
MALDI mass spectra of highly acidic oligosaccharides by
using a nafion-coated probe. Anal Chem 73:405–410.
Jacoby ES, Kicman AT, Laidler P, Iles RK. 2000. Determination
of the glycoforms of human chorionic gonadotropin betacore fragment by matrix-assisted laser desorption/ionization
time-of-flight mass spectrometry. Clin Chem 46:1796–
1803.
Johnson JV, Yost RA. 1985. Tandem mass spectrometry for trace
analysis. Anal Chem 57:758A–768A.
Johnson JV, Yost RA, Kelley PE, Bradford DC. 1990. Tandem-inspace and tandem-in-time mass spectrometry: Triple quadrupoles and quadrupole ion traps. Anal Chem 62:2162–
2172.
Juhasz P, Biemann K. 1995. Utility of non-covalent complexes in
the matrix-assisted laser desorption ionization mass spectrometry of heparin-derived oligosaccharides. Carbohydr Res
270:131–147.
Kakehi K, Kinoshita M, Kawakami D, Tanaka J, Sei K, Endo K,
Oda Y, Iwaki M, Masuko T. 2001. Capillary electrophoresis
of sialic acid-containing glycoprotein. Effect of the heterogeneity of carbohydrate chains on glycoform separation
using an alpha(1)-acid glycoprotein as a model. Anal Chem
73:2640–2647.
Karas M, Bahr U, Dulcks T. 2000. Nano-electrospray ionization
mass spectrometry: Addressing analytical problems beyond
routine. Fresenius J Anal Chem 366:669–676.
Karlsson H, Carlstedt I, Hansson GC. 1989. The use of gas
chromatography and gas chromatography-mass spectrometry for the characterization of permethylated oligosac-
220
charides with molecular mass up to 2300. Anal Biochem
182:438–446.
Karlsson NG, Karlsson H, Hansson GC. 1995. Strategy for the
investigation of O-linked oligosaccharides from mucins
based on the separation into neutral, sialic acid- and sulfatecontaining species. Glycoconj J 12:69–76.
Karlsson NG, Karlsson H, Hansson GC. 1996. Sulphated mucin
oligosaccharides from porcine small intestine analysed by
four-sector tandem mass spectrometry. J Mass Spectrom
31:560–572.
Kawasaki N, Haishima Y, Ohta M, Itoh S, Hyuga M, Hyuga S,
Hayakawa T. 2001. Structural analysis of sulfated N-linked
oligosaccharides in erythropoietin. Glycobiology 11:1043–
1049.
Kelleher NL, Taylor SV, Grannis D, Kinsland C, Chiu HJ,
Begley TP, McLafferty FW. 1998. Efficient sequence
analysis of the six gene products (7–74 kDa) from the
Escherichia coli thiamin biosynthetic operon by tandem
high-resolution mass spectrometry. Protein Sci 7:1796–
1801.
Kelleher NL, Zubarev RA, Bush K, Furie B, Furie BC,
McLafferty FW, Walsh CT. 1999. Localization of labile
posttranslational modifications by electron capture dissociation: The case of gamma-carboxyglutamic acid. Anal Chem
71:4250–4253.
Khoo KH, Morris HR, McDowell RA, Dell A, Maccarana M,
Lindahl U. 1993. FABMS/derivatisation strategies for the
analysis of heparin-derived oligosaccharides. Carbohydr
Res 244:205–223.
Kim YS, Ahn MY, Wu SJ, Kim DH, Toida T, Teesch LM, Park Y,
Yu G, Lin J, Linhardt RJ. 1998. Determination of the
structure of oligosaccharides prepared from acharan sulfate.
Glycobiology 8:869–877.
Kinoshita A, Yamada S, Haslam SM, Morris HR, Dell A,
Sugahara K. 1997. Novel tetrasaccharides isolated from
squid cartilage chondroitin sulfate E contain unusual sulfated disaccharide units GlcA(3-O-sulfate)beta1-3GalNAc(6-O-sulfate) or GlcA(3-O-sulfate)beta1-3GalNAc.
J Biol Chem 272:19656–19665.
Kinoshita A, Yamada S, Haslam SM, Morris HR, Dell A,
Sugahara K. 2001. Isolation and structural determination of
novel sulfated hexasaccharides from squid cartilage chondroitin sulfate E that exhibits neuroregulatory activities.
Biochemistry 40:12654–12665.
Kjeldsen F, Haselmann KF, Budnik BA, Sorensen ES, Zubarev
RA. 2003. Complete characterization of posttranslational
modification sites in the bovine milk protein PP3 by tandem
mass spectrometry with electron capture dissociation as the
last stage. Anal Chem 75:2355–2361.
Kohler M, Leary JA. 1995. LC/MS/MS of carbohydrates with
postcolumn addition of metal chlorides using a triaxial
electrospray probe. Anal Chem 67:3501–3508.
Konig S, Leary JA. 1998. Evidence for linkage position determination in cobalt coordinated pentasaccharides using ion
trap mass spectrometry. J Am Soc Mass Spectrom 9:1125–
1134.
MASS SPECTROMETRY OF OLIGOSACCHARIDES
Kovacik V, Hirsch J, Kovac P, Heerma W, Thomasoates J,
Haverkamp J. 1995. Oligosaccharide characterization
using collision-induced dissociation fast-atom-bombardment mass-spectrometry—Evidence for internal monosaccharide residue loss. J Mass Spectrom 30:949–958.
Kragten EA, Bergwerff AA, vanOostrum J, Müller DR, Richter
WJ. 1995. Site-specific analysis of the N-glycans on murine
polymeric immunoglobulin a using liquid chromatography
electrospray mass spectrometry. J Mass Spectrom 30:1679–
1686.
Kruger NA, Zubarev RA, Horn DM, McLafferty FW. 1999.
Electron capture dissociation of multiply charged peptide
cations. Int J Mass Spectrom 185/186/187:787–793.
Krutchinsky AN, Loboda AV, Spicer VL, Dworschak R, Ens W,
Standing KG. 1998. Orthogonal injection of matrix-assisted
laser desorption/ionization ions into a time-of-flight spectrometer through a collisional damping interface. Rapid
Commun Mass Spectrom 12:508–518.
Kuberan B, Lech M, Zhang L, Wu ZL, Beeler DL, Rosenberg R.
2002. Analysis of heparan sulfate oligosaccharides with ion
pair-reverse phase capillary high performance liquid
chromatography—Microelectrospray ionization time-offlight mass Spectrometry. J Am Chem Soc 124:8707–
8718.
Kushi Y, Handa S, Ishizuka I. 1985. Secondary ion mass
spectrometry for sulfoglycolipids: application of negative
ion detection. J Biochem (Tokyo) 97:419–428.
Kwan CP, Venkataraman G, Shriver Z, Raman R, Liu D, Qi Y,
Varticovski L, Sasisekharan R. 2001. Probing fibroblast
growth factor dimerization and role of heparin-like glycosaminoglycans in modulating dimerization and signaling.
J Biol Chem 276:23421–23429.
Körner R, Limberg G, Mikkelsen JD, Roepstorff P. 1998.
Characterization of enzymatic pectin digests by matrixassisted laser desorption/ionization mass spectrometry.
J Mass Spectrom 33:836–842.
Körner R, Limberg G, Christensen TM, Mikkelsen JD,
Roepstorff P. 1999. Sequencing of partially methylesterified oligogalacturonates by tandem mass spectrometry
and its use to determine pectinase specificities. Anal Chem
71:1421–1427.
Küster B, Mann M. 1999. 18O-labeling of N-glycosylation sites to
improve the identification of gel-separated glycoproteins
using peptide mass mapping and database searching. Anal
Chem 71:1431–1440.
Küster B, Wheeler SF, Hunter AP, Dwek RA, Harvey DJ. 1997.
Sequencing of N-linked oligosaccharides directly from
protein gels: In-gel deglycosylation followed by matrixassisted laser desorption/ionization mass spectrometry and
normal-phase high-performance liquid chromatography.
Anal Biochem 250:82–101.
Küster B, Hunter AP, Wheeler SF, Dwek RA, Harvey DJ. 1998.
Structural determination of N-linked carbohydrates by
matrix-assisted laser desorption/ionization-mass spectrometry following enzymatic release within sodium dodecyl
sulphate-polyacrylamide electrophoresis gels: Application
&
to species-specific glycosylation of alpha1-acid glycoprotein. Electrophoresis 19:1950–1959.
Laiko VV, Baldwin MA, Burlingame AL. 2000. Atmospheric
pressure matrix assisted laser desorption/ionization mass
spectrometry. Anal Chem 72:652–657.
Laiko VV, Moyer SC, Cotter RJ. 2000. Atmospheric pressure
MALDI/ion trap mass spectrometry. Anal Chem 72:5239–
5243.
Laine RA. 1989. Tandem mass spectrometry of oligosaccharides.
Methods Enzymol 179:157–164.
Laine RA, Pamidimukkala KM, French AD, Hall RW, Abbas SA,
Jain RK, Matta KL. 1988. Linkage position in oligosaccharides by fast atom bombardment ionization, collisionactivated dissociation, tandem mass-spectrometry and
molecular modeling—L-Fucosylp-(alpha-1-]3)-D-N-acetylD -glucosamylp-(beta-1-]3)- D -galactosylp-(beta-1-OMethyl), L-fucosylp-(alpha-1-]4)-D-N-acetyl-D-glucosaminylp-(beta-1-]3)-D-galactosylp-(beta-1-O-methyl), L-fucosylp-(alpha-1-]6)-D-N-acetyl-D-glucosaminylp-(beta-1-]3)D-galactosylp-(beta-1-O-methyl)-. J Am Chem Soc 110:
6931–6939.
Laine RA, Yoon E, Mahier TJ, Abbas SA, de Lappe B, Jain R,
Matta K. 1991. Non-reducing terminal linkage position
determination in intact and permethylated synthetic oligosaccharides having a penultimate amino sugar: Fast atom
bombardment ionization, collisional-induced dissociation
and tandem mass spectrometry. Biol Mass Spectrom 20:
505–514.
Lamb DJ, Wang HM, Mallis LM, Linhardt RJ. 1992. Negative ion
fast-atom bombardment tandem mass spectrometry to
determine sulfate and linkage position in glycosaminoglycan derived disaccharides. J Am Soc Mass Spectrom 3:797–
803.
Leavell MD, Leary JA. 2001. Stabilization and linkage analysis
of metal-ligated sialic acid containing oligosaccharides.
J Am Soc Mass Spectrom 12:528–536.
Lemoine J, Chirat F, Domon B. 1996. Structural analysis of derivatized oligosaccharides using post-source decay matrixassisted laser desorption ionization mass spectrometry.
J Mass Spectrom 31:908–912.
Lemoine J, Strecker G, Leroy Y, Fournet B, Ricart G. 1991.
Collisional-activation tandem mass spectrometry of sodium
adduct ions of methylated oligosaccharides: sequence
analysis and discrimination between alpha-NeuAc-(2–3)
and alpha-NeuAc-(2–6) linkages. Carbohydr Res 221:209–
217.
Lemoine J, Fournet B, Despeyroux D, Jennings KR, Rosenberg
R, Dehoffmann E. 1993. Collision-induced dissociation of
alkali-metal cationized and permethylated oligosaccharides—Influence of the collision energy and of the collision
gas for the assignment of linkage position. J Am Soc Mass
Spectrom 4:197–203.
Li T, Kubota M, Okuda S, Hirano T, Ohashi M. 1995. Positive-ion
fast atom bombardment tandem mass spectrometry for
characterization of sulfated disaccharides from heparin and
heparansulfate. Eur Mass Spectrom 1:11–22.
221
&
ZAIA
Lindberg B, Lonngren J. 1978. Methylation analysis of complex
carbohydrates: general procedure and application for sequence analysis. Methods Enzymol 50:3–33.
Linhardt RJ, Wang HM, Loganthan D, Lamb DJ, Mallis LM.
1992. Analysis of glycosaminoglycan-derived oligosaccharides using fast atom bombardment mass spectrometry.
Carbohydr Res 225:137–145.
Little DP, McLafferty FW. 1995. Sequencing 50-Mer DNAs
using electrospray tandem mass spectrometry and complementary fragmentation methods. J Am Chem Soc 117:
6783–6784.
Little DP, Speir JP, Senko MW, Oconnor PB, McLafferty FW.
1994. Infrared multiphoton dissociation of large multiply
charged ions for biomolecule sequencing. Anal Chem
66:2809–2815.
Little DP, Aaserud DJ, Valaskovic GA, McLafferty FW. 1996.
Sequence information from 42–108-Mer DNAs (complete
for a 50-Mer) by tandem mass spectrometry. J Am Chem Soc
118:9352–9359.
Liu J, Shriver Z, Pope M, Thorp SC, Duncan MB, Copeland RJ,
Raska CS, Yoshida K, Eisenberg RJ, Cohen G, Linhardt RJ,
Sasisekharan R. 2002. Characterization of a heparan sulfate
octasaccharide that binds to herpes simplex viral type 1
glycoprotein D. J Biol Chem 277:33456–33467.
Loboda AV, Krutchinsky AN, Bromirski M, Ens W, Standing KG.
2000. A tandem quadrupole/time-of-flight mass spectrometer with a matrix-assisted laser desorption/ionization
source: Design and performance. Rapid Commun Mass
Spectrom 14:1047–1057.
LoGuidice JM, Herz H, Lamblin G, Plancke Y, Roussel P,
Lhermitte M. 1997. Structures of sulfated oligosaccharides
isolated from the respiratory mucins of a non-secretor (O,
Le(aþb–)) patient suffering from chronic bronchitis.
Glycoconj J 14:113–125.
Macek B, Hofsteenge J, Peter-Katalinic J. 2001. Direct
determination of glycosylation sites in O-fucosylated
glycopeptides using nano-electrospray quadrupole timeof-flight mass spectrometry. Rapid Commun Mass Spectrom
15:771–777.
Maier CS, Yan XG, Harder ME, Schimerlik MI, Deinzer ML,
Pasa-Tolic L, Smith RD. 2000. Electrospray ionization
Fourier transform ion cyclotron resonance mass spectrometric analysis of the recombinant human macrophage
colony stimulating factor beta and derivatives. J Am Soc
Mass Spectrom 11:237–243.
Mallis LM, Wang HM, Loganathan D, Linhardt RJ. 1989.
Sequence analysis of highly sulfated, heparin-derived
oligosaccharides using fast atom bombardment mass spectrometry. Anal Chem 61:1453–1458.
Mann M, Wilm M. 1994. Error-tolerant identification of peptides
in sequence databases by peptide sequence tags. Anal Chem
66:4390–4399.
Marbois BN, Faull KF, Fluharty AL, Raval-Fernandes S, Rome
LH. 2000. Analysis of sulfatide from rat cerebellum and
multiple sclerosis white matter by negative ion electrospray
mass spectrometry. Biochim Biophys Acta 1484:59–70.
222
March RE. 1998. Quadrupole ion trap mass spectrometry:
Theory, simulation, recent developments and applications.
Rapid Commun Mass Spectrom 12:1543–1554.
March RE. 2000. Quadrupole ion trap mass spectrometry: Aview
at the turn of the century. Int J Mass Spectrom 200:285–312.
March RE, Hughes HE. 1989. Quadrupole storage mass
spectrometry. New York, NY: John Wiley & Sons, Inc.
March RE, Londry FA. 1995. Theory of quadrupole mass
spectrometry. In: March RE, Todd JFJ, editors. Practical
aspects of ion trap mass spectrometry. Vol. 1. New York, NY:
CRC Press, Inc. pp 25–48.
Marshall AG, Wang T-CL, Ricca TL. 1985. Tailored excitation
for Fourier transform ion cyclotron resonance mass spectrometry. J Am Chem Soc 107:7893–7897.
Mawhinney TP, Chance DL. 1994. Structural elucidation by fastatom-bombardment mass-spectrometry of multisulfated
oligosaccharides isolated from human respiratory mucous
glycoproteins. J Carbohydr Chem 13:825–840.
McClellan JM, Costello CE, O’Connor PB, Zaia J. 2002.
Influence of charge state on product ion mass spectra and
the determination of 4S/6S sulfation sequence of chondroitin sulfate oligosaccharides. Anal Chem 74:3760–3771.
McNeil M, Darvill AG, Aman P, Franzen LE, Albersheim P.
1982. Structural analysis of complex carbohydrates using
high-performance liquid chromatography, gas chromatography, and mass spectrometry. Methods Enzymol 83:3–45.
Medzihradszky KF, Gillece-Castro BL, Settineri CA, Townsend
RR, Masiarz FR, Burlingame AL. 1990. Structure determination of O-linked glycopeptides by tandem mass spectrometry. Biomed Environ Mass Spectrom 19:777–781.
Medzihradszky K, Gillece-Castro B, Townsend R, Burlingame
A, Hardy M. 1996. Structural elucidation of O-linked
glycopeptides by high energy collision-induced dissociation. J Am Soc Mass Spectrom 7:319–328.
Medzihradszky KF, Campbell JM, Baldwin MA, Falick AM,
Juhasz P, Vestal ML, Burlingame AL. 2000. The characteristics of peptide collision-induced dissociation using a highperformance MALDI-TOF/TOF tandem mass spectrometer.
Anal Chem 72:552–558.
Mendonca S, Cole RB, Zhu JH, Cai Y, French AD, Johnson GP,
Laine RA. 2003. Incremented alkyl derivatives enhance
collision induced glycosidic bond cleavage in mass spectrometry of disaccharides. J Am Soc Mass Spectrom 14:63–78.
Meng CK, Mann M, Fenn JB. 1988. Of protons or proteins. Z Phys
D 10:361–368.
Mirgorodskaya E, Roepstorff P, Zubarev RA. 1999. Localization
of O-glycosylation sites in peptides by electron capture
dissociation in a Fourier transform mass spectrometer. Anal
Chem 71:4431–4436.
Mizuno Y, Sasagawa T, Dohmae N, Takio K. 1999. An automated
interpretation of MALDI/TOF postsource decay spectra of
oligosaccharides. 1. Automated peak assignment. Anal
Chem 71:4764–4771.
Morris HR, Paxton T, Dell A, Langhorne J, Berg M, Bordoli RS,
Hoyes J, Bateman RH. 1996. High sensitivity collisionallyactivated decomposition tandem mass spectrometry on a
MASS SPECTROMETRY OF OLIGOSACCHARIDES
novel quadrupole/orthogonal-acceleration time-of-flight
mass spectrometer. Rapid Commun Mass Spectrom 10:
889–896.
Morris HR, Paxton T, Panico M, McDowell R, Dell A. 1997. A
novel geometry mass spectrometer, the Q-TOF, for lowfemtomole/attomole-range biopolymer sequencing. J Protein Chem 16:469–479.
Mortz E, Sareneva T, Julkunen I, Roepstorff P. 1996. Does
matrix-assisted laser desorption ionization mass spectrometry allow analysis of carbohydrate heterogeneity in
glycoproteins? A study of natural human interferon-gamma.
J Mass Spectrom 31:1109–1118.
Moyer SC, Cotter RJ. 2002. Atmospheric pressure MALDI. Anal
Chem 74:468A–476A.
Moyer SC, Marzilli LA, Woods A, Laiko VV, Doroshenko VM,
Cotter RJ. 2002. Atmospheric pressure matrix-assisted laser
desorption/ionization (AP MALDI) on a quadrupole ion
trap mass spectrometer. Int J Mass Spectrom 226:133–
150.
Mueller DR, Domon BM, Blum W, Raschdorf F, Richter WJ.
1988. Direct stereochemical assignment of sugar subunits in
naturally occurring glycosides by low energy collision
induced dissociation. Application to papulacandin antibiotics. Biomed Environ Mass Spectrom 15:441–446.
Müller DR, Domon BM, Blum W, Raschdorf F, Richter WJ.
1988. Direct stereochemical assignment of sugar subunits in
naturally occurring glycosides by low energy collision
induced dissociation. Application to papulacandin antibiotics. Biomed Environ Mass Spectrom 15:441–446.
Nemeth JF, Hochgesang GP, Jr, Marnett LJ, Caprioli RM,
Hochensang GP, Jr. 2001. Characterization of the glycosylation sites in cyclooxygenase-2 using mass spectrometry.
Biochemistry 40:3109–3116.
Nemeth-Cawley JF, Karnik S, Rouse JC. 2001. Analysis of sulfated peptides using positive electrospray ionization tandem
mass spectrometry. J Mass Spectrom 36:1301–1311.
Nilsson B, Zopf D. 1982. Gas chromatography and mass
spectrometry of hexosamine-containing oligosaccharide
alditols as their permethylated, N-trifluoroacetyl derivatives.
Methods Enzymol 83:46–58.
Nilsson B, Zopf D. 1983. Oligosaccharides released from
glycolipids by trifluoroacetolysis can be analyzed by gas
chromatography-mass spectrometry. Arch Biochem Biophys 222:628–648.
O’Connor PB, Mirgorodskaya E, Costello CE. 2002. High
pressure matrix-assisted laser desorption/ionization Fourier
transform mass spectrometry for minimization of ganglioside fragmentation. J Am Soc Mass Spectrom 13:402–
407.
O’Connor PB, Costello CE. 2001. A high pressure matrixassisted laser desorption/ionization Fourier transform mass
spectrometry ion source for thermal stabilization of labile
biomolecules. Rapid Commun Mass Spectrom 15:1862–
1868.
Ohta M, Kawasaki N, Hyuga S, Hyuga M, Hayakawa T. 2001.
Selective glycopeptide mapping of erythropoietin by on-line
&
high-performance liquid chromatography-electrospray ionization mass spectrometry. J Chromatogr A 910:1–11.
Okafo G, Langridge J, North S, Organ A, West A, Morris M,
Camilleri P. 1997. High-performance liquid chromatographic analysis of complex N-linked glycans derivatized
with 2-aminoacridone. Anal Chem 69:4985–4993.
Okamoto M, Takahashi K, Doi T, Takimoto Y. 1997. Highsensitivity detection and postsource decay of 2-aminopyridine- derivatized oligosaccharides with matrix-assisted
laser desorption/ionization mass spectrometry. Anal Chem
69:2919–2926.
Orlando R, Bush CA, Fenselau C. 1990. Structural-analysis of
oligosaccharides by tandem mass-spectrometry—Collisional activation of sodium adduct ions. Biomed Environ
Mass Spectrom 19:747–754.
Overberg A, Karas M, Bahr U, Kaufmann R, Hillenkamp F. 1990.
Matrix-assisted infrared-laser (2.94 mm) desorption/ionization mass spectrometry of large biomolecules. Rapid Commun Mass Spectrom 4:293–296.
Papac DI, Wong A, Jones AJ. 1996. Analysis of acidic
oligosaccharides and glycopeptides by matrix-assisted laser
desorption/ionization time-of-flight mass spectrometry.
Anal Chem 68:3215–3223.
Papayannopoulos IA. 1995. The interpretation of collisioninduced dissociation tandem mass spectra of peptides. Mass
Spectrom Rev 14:49–73.
Penn SG, Cancilla MT, Lebrilla CB. 2000. Fragmentation behavior of multiple-metal-coordinated acidic oligosaccharides studied by matrix-assisted laser desorption ionization
Fourier transform mass spectrometry. Int J Mass Spectrom
196:259–269.
Perrault H, Costello CE. 1994. Liquid secondary ionization,
tandem and matrix-assisted laser desorption/ionization
time-of-flight mass spectrometric characterization of glycosphingolipid derivatives. Org Mass Spectrom 29:720–735.
Perreault H, Costello CE. 1996. Differentiation of cerebroside
isomers and study of fragmentation by liquid secondary ion
mass spectrometry and mass spectrometry mass spectrometry of selected derivatives. Canadian J Chem-Revue
Canadienne De Chimie 74:1682–1695.
Perreault H, Hronowski XL, Koul O, Street J, McCluer RH,
Costello CE. 1997. High sensitivity mass spectral characterization of glycosphingolipids from bovine erythrocytes,
mouse kidney and fetal calf brain. Int J Mass Spectrom
169:351–370.
Peter-Katalinic J. 1994. Analysis of glycoconjugates by fast atom
bombardment mass spectrometry and related MS techniques. Mass Spectrom Rev 13:77–98.
Pitt JJ, Gorman JJ. 1996. Matrix-assisted laser desorption/
ionization time-of-flight mass spectrometry of sialylated
glycopeptides and proteins using 2,6-dihydroxyacetophenone as a matrix. Rapid Commun Mass Spectrom 10:1786–
1788.
Pope RM, Raska CS, Thorp SC, Liu J. 2001. Analysis of heparan
sulfate oligosaccharides by nano-electrospray ionization
mass spectrometry. Glycobiology 11:505–513.
223
&
ZAIA
Poulter L, Burlingame AL. 1990. Desorption mass spectrometry
of oligosaccharides coupled with hydrophobic chromophores. Methods Enzymol 193:661–689.
Powell AK, Harvey DJ. 1996. Stabilization of sialic acids in Nlinked oligosaccharides and gangliosides for analysis by
positive ion matrix-assisted laser desorption/ionization mass
spectrometry. Rapid Commun Mass Spectrom 10:1027–
1032.
Prakobphol A, Thomsson KA, Hansson GC, Rosen SD, Singer
MS, Phillips NJ, Medzihradszky KF, Burlingame AL,
Leffler H, Fisher SJ. 1998. Human low-molecular-weight
salivary mucin expresses the sialyl lewisx determinant and
has L-selectin ligand activity. Biochemistry 37:4916–4927.
Rademaker GJ, Pergantis SA, Blok-Tip L, Langridge JI, Kleen A,
Thomas-Oates JE. 1998. Mass spectrometric determination
of the sites of O-glycan attachment with low picomolar
sensitivity. Anal Biochem 257:149–160.
Reinhold BB, Chan S-Y, Reinhold VN. 1994. Profiling glycosphingolipid structural detail—Periodate-oxidation, electrospray, collision-induced dissociation and tandem massspectrometry. Org Mass Spectrom 29:736–7469.
Reinhold VN, Reinhold BB, Costello CE. 1995. Carbohydrate
molecular weight profiling, sequence, linkage, and branching data: ES-MS and CID. Anal Chem 67:1772–1784.
Reinhold VN, Sheeley DM. 1998. Detailed characterization of
carbohydrate linkage and sequence in an ion trap mass
spectrometer: Glycosphingolipids. Anal Biochem 259:
28–33.
Reinhold VN, Carr SA, Green BN, Petitou M, Choay J, Sinay P.
1987. Structural characterization of sulfated glycosaminoglycans by fast atom bombardment mass spectrometry:
application to heparin fragments prepared by chemical
synthesis. Carbohydr Res 161:305–313.
Rhomberg AJ, Ernst S, Sasisekharan R, Biemann K. 1998. Mass
spectrometric and capillary electrophoretic investigation of
the enzymatic degradation of heparin-like glycosaminoglycans. Proc Natl Acad Sci USA 95:4176–4181.
Rhomberg AJ, Shriver Z, Biemann K, Sasisekharan R. 1998.
Mass spectrometric evidence for the enzymatic mechanism
of the depolymerization of heparin-like glycosaminoglycans by heparinase II. Proc Natl Acad Sci USA 95:12232–
12237.
Ritchie MA, Gill AC, Deery MJ, Lilley K. 2002. Precursor ion
scanning for detection and structural characterization of
heterogeneous glycopeptide mixtures. J Am Soc Mass
Spectrom 13:1065–1077.
Roboz J. 1968. Introduction to mass spectrometry: Instrumentation and Techniques. New York, NY: Interscience.
Ruiz-Calero V, Moyano E, Puignou L, Galceran MT. 2001.
Pressure-assisted capillary electrophoresis-electrospray ion
trap mass spectrometry for the analysis of heparin
depolymerised disaccharides. J Chromatogr A 914:277–
291.
Saba JA, Shen XD, Jamieson JC, Perreault H. 1999. Effect of 1phenyl-3-methyl-5-pyrazolone labeling on the fragmentation behavior of asialo and sialylated N-linked glycans under
224
electrospray ionization conditions. Rapid Commun Mass
Spectrom 13:704–711.
Sandhoff R, Hepbildikler ST, Jennemann R, Geyer R, Gieselmann V, Proia RL, Wiegandt H, Grone HJ. 2002. Kidney
sulfatides in mouse models of inherited glycosphingolipid
disorders: determination by nano-electrospray ionization
tandem mass spectrometry. J Biol Chem 277:20386–20398.
Sato Y, Suzuki M, Nirasawa T, Suzuki A, Endo T. 2000.
Microsequencing of glycans using 2-aminobenzamide and
MALDI-TOF mass spectrometry: Occurrence of unique
linkage-dependent fragmentation. Anal Chem 72:1207–
1216.
Schiller J, Arnhold J, Benard S, Reichl S, Arnold K. 1999.
Cartilage degradation by hyaluronate lyase and chondroitin
ABC lyase: a MALDI-TOF mass spectrometric study.
Carbohydr Res 318:116–122.
Senko MW, Speir JP, McLafferty FW. 1994. Collisional
activation of large multiply-charged ions using Fourier
transform mass spectrometry. Anal Chem 66:2801–2808.
Severs JC, Carnine M, Eguizabal H, Mock KK. 1999.
Characterization of tyrosine sulfate residues in antihemophilic recombinant factor VIII by liquid chromatography
electrospray ionization tandem mass spectrometry and
amino acid analysis. Rapid Commun Mass Spectrom 13:
1016–1023.
Sheeley DM, Reinhold VN. 1998. Structural characterization of
carbohydrate sequence, linkage, and branching in a quadrupole ion trap mass spectrometer: Neutral oligosaccharides
and N-linked glycans. Anal Chem 70:3053–3059.
Shevchenko A, Chernushevich I, Ens W, Standing KG, Thomson
B, Wilm M, Mann M. 1997. Rapid ‘de novo’ peptide
sequencing by a combination of nanoelectrospray, isotopic
labeling and a quadrupole/time-of-flight mass spectrometer.
Rapid Commun Mass Spectrom 11:1015–1024.
Shi SD, Hemling ME, Carr SA, Horn DM, Lindh I, McLafferty
FW. 2001. Phosphopeptide/phosphoprotein mapping by
electron capture dissociation mass spectrometry. Anal Chem
73:19–22.
Shriver Z, Raman R, Venkataraman G, Drummond K, Turnbull J,
Toida T, Linhardt R, Biemann K, Sasisekharan R. 2000.
Sequencing of 3-O-sulfate containing heparin decasaccharides with a partial antithrombin III binding site. Proc Natl
Acad Sci USA 97:10359–10364.
Shriver Z, Sundaram M, Venkataraman G, Fareed J, Linhardt R,
Biemann K, Sasisekharan R. 2000. Cleavage of the
antithrombin III binding site in heparin by heparinases and
its implication in the generation of low molecular weight
heparin. Proc Natl Acad Sci USA 97:10365–10370.
Sible EM, Brimmer SP, Leary JA. 1997. Interaction of first row
transition metals with alpha 1-3, alpha 1-6 mannotriose and
conserved trimannosyl core oligosaccharides: A comparative electrospray ionization study of doubly and singly
charged complexes. J Am Soc Mass Spectrom 8:32–42.
Siegel MM, Tabei K, Kagan MZ, Vlahov IR, Hileman RE,
Linhardt R. 1997. Polysulfated carbohydrates analyzed as
ion-paired complexes with basic peptides and proteins using
MASS SPECTROMETRY OF OLIGOSACCHARIDES
electrospray negative ionization mass spectrometry. J Mass
Spectrom 32:760–772.
Solouki T, Pasa-Tolic L, Jackson GS, Guan S, Marshall AG.
1996. High-resolution multistage MS, MS2, and MS3
matrix-assisted laser desorption/ionization FT-ICR mass
spectra of peptides from a single laser shot. Anal Chem
68:3718–3725.
Solouki T, Reinhold BB, Costello CE, O’Malley M, Guan S,
Marshall AG. 1998. Electrospray ionization and matrixassisted laser desorption/ionization Fourier transform ion
cyclotron resonance mass spectrometry of permethylated
oligosaccharides. Anal Chem 70:857–864.
Spengler B, Dolce JW, Cotter RJ. 1990. Infrared-laser desorption mass-spectrometry of oligosaccharides—Fragmentation mechanisms and isomer analysis. Anal Chem 62:1731–
1737.
Spengler B, Kirsch D, Kaufmann R. 1991. Metastable decay of
peptides and proteins in matrix-assisted laser-desorption
mass-spectrometry. Rapid Commun Mass Spectrom 5:198–
202.
Spengler B, Kirsch D, Kaufmann R, Lemoine J. 1994. Structureanalysis of branched oligosaccharides using post-source
decay in matrix-assisted laser-desorption ionization massspectrometry. Org Mass Spectrom 29:782–787.
Steen H, Küster B, Mann M. 2001. Quadrupole time-of-flight
versus triple-quadrupole mass spectrometry for the determination of phosphopeptides by precursor ion scanning.
J Mass Spectrom 36:782–790.
Stensballe A, Jensen ON, Olsen JV, Haselmann KF, Zubarev RA.
2000. Electron capture dissociation of singly and multiply
phosphorylated peptides. Rapid Commun Mass Spectrom
14:1793–1800.
Strecker G, Wieruszeski J-M, Martel C, Montreuil J. 1987.
Determination of the structure of sulfated tetra- and
pentaoligosaccharides obtained by alkaline borohydride
degradation of hen ovomucin. A fast atom bombardmentmass spectrometric study. Glycoconj J 4:329–337.
Sturiale L, Naggi A, Torri G. 2001. MALDI mass spectrometry as
a tool for characterizing glycosaminoglycan oligosaccharides and their interaction with proteins. Semin Thromb
Hemost 27:465–472.
Sugahara K, Takemura Y, Sugiura M, Kohno Y, Yoshida K,
Takeda K, Khoo KH, Morris HR, Dell A. 1994. Chondroitinase ABC-resistant sulfated trisaccharides isolated from
digests of chondroitin/dermatan sulfate chains. Carbohydr
Res 255:165–182.
Sugahara K, Tanaka Y, Yamada S, Seno N, Kitagawa H,
Haslam SM, Morris HR, Dell A. 1996. Novel sulfated
oligosaccharides containing 3-O-sulfated glucuronic acid
from king crab cartilage chondroitin sulfate K. Unexpected
degradation by chondroitinase ABC. J Biol Chem 271:
26745–26754.
Tadano-Aritomi K, Kubo H, Ireland P, Okuda M, Kasama T,
Handa S, Ishizuka I. 1995. Structural analysis of monoand bis-sulfated glycosphingolipids by negative liquid
secondary ion mass spectrometry with high- and low-
&
energy collision-induced dissociation. Carbohydr Res 273:
41–52.
Taguchi T, Iwasaki M, Muto Y, Kitajima K, Inoue S, Khoo
KH, Morris HR, Dell A, Inoue Y. 1996. Occurrence and
structural analysis of highly sulfated multiantennary Nlinked glycan chains derived from a fertilization-associated
carbohydrate-rich glycoprotein in unfertilized eggs of
Tribolodon hakonensis. Eur J Biochem 238:357–367.
Takagaki K, Kojima K, Majima M, Nakamura T, Kato I, Endo M.
1992. Ion-spray mass spectrometric analysis of glycosaminoglycan oligosaccharides. Glycoconj J 9:174–179.
Takao T, Tambara Y, Nakamura A, Yoshino K, Fukuda H, Fukuda
M, Shimonishi Y. 1996. Sensitive analysis of oligosaccharides derivatized with 4-aminobenzoic acid 2-(diethylamino)ethyl ester by matrix-assisted laser desorption/ionization
mass spectrometry. Rapid Commun Mass Spectrom 10:
637–640.
Talbo G, Roepstorff P. 1993. Determination of sulfated peptides
via prompt fragmentation by UV matrix-assisted laser
desorption/ionization mass spectrometry. Rapid Commun
Mass Spectrom 7:201–204.
Tang XJ, Ens W, Standing KG, Westmore JB. 1988. Daughter ion
mass spectra from cationized molecules of small oligopeptides in a reflecting time-of-flight mass spectrometer. Anal
Chem 60:1791–1799.
Tarelli E, Byers HL, Wilson M, Roberts G, Homer KA, Beighton
D. 2000. Detecting mannosidase activities using ribonuclease B and matrix-assisted laser desorption/ionizationtime of flight mass spectrometry. Anal Biochem 282:165–
172.
Tetaert D, Briand G, Soudan B, Richet C, Demeyer D, Boersma
A, Degand P. 1994. Analysis by electrospray mass spectrometry of glycopeptides from the in vitro O-glycosylation
reaction using human mucin motif peptide. Anal Biochem
222:409–416.
Todd JFJ. 1991. Ion trap mass spectrometer–Past, present, and
future(?). Mass Spectrom Rev 10:3–52.
Tseng K, Hedrick JL, Lebrilla CB. 1999. Catalog-library
approach for the rapid and sensitive structural elucidation
of oligosaccharides. Anal Chem 71:3747–3754.
Tseng K, Lindsay LL, Penn S, Hedrick JL, Lebrilla CB. 1997.
Characterization of neutral oligosaccharide-alditols from
Xenopus laevis egg jelly coats by matrix-assisted laser
desorption Fourier transform mass spectrometry. Anal
Biochem 250:18–28.
Venkataraman G, Shriver Z, Davis JC, Sasisekharan R. 1999a.
Fibroblast growth factors 1 and 2 are distinct in oligomerization in the presence of heparin-like glycosaminoglycans.
Proc Natl Acad Sci USA 96:1892–1897.
Venkataraman G, Shriver Z, Raman R, Sasisekharan R. 1999b.
Sequencing complex polysaccharides. Science 286:537–
542.
Viseux N, Costello CE, Domon B. 1999. Post-source decay mass
spectrometry: optimized calibration procedure and structural characterization of permethylated oligosaccharides.
J Mass Spectrom 34:364–376.
225
&
ZAIA
Viseux N, de Hoffmann E, Domon B. 1997. Structural analysis of
permethylated oligosaccharides by electrospray tandem
mass spectrometry. Anal Chem 69:3193–3198.
Viseux N, de Hoffmann E, Domon B. 1998. Structural assignment of permethylated oligosaccharide subunits using sequential tandem mass spectrometry. Anal Chem 70:4951–
4959.
Viseux N, Hronowski X, Delaney J, Domon B. 2001. Qualitative
and quantitative analysis of the glycosylation pattern of
recombinant proteins. Anal Chem 73:4755–4762.
Wang WT, Matsuura F, Sweeley CC. 1983. Chromatography of
permethylated oligosaccharide alditols. Anal Biochem 134:
398–405.
Wang WT, LeDonne NC, Jr, Ackerman B, Sweeley CC. 1984.
Structural characterization of oligosaccharides by highperformance liquid chromatography, fast-atom bombardment-mass spectrometry, and exoglycosidase digestion.
Anal Biochem 141:366–381.
Weiskopf AS, Vouros P, Harvey DJ. 1997. Characterization of
oligosaccharide composition and structure by quadrupole
ion trap mass spectrometry. Rapid Commun Mass Spectrom
11:1493–1504.
Weiskopf AS, Vouros P, Harvey DJ. 1998. Electrospray
ionization-ion trap mass spectrometry for structural analysis
of complex N-linked glycoprotein oligosaccharides. Anal
Chem 70:4441–4447.
Wheeler SF, Harvey DJ. 2000. Negative ion mass spectrometry of
sialylated carbohydrates: discrimination of N-acetylneuraminic acid linkages by MALDI-TOF and ESI-TOF mass
spectrometry. Anal Chem 72:5027–5039.
Wheeler SF, Harvey DJ. 2001. Extension of the in-gel release
method for structural analysis of neutral and sialylated Nlinked glycans to the analysis of sulfated glycans: Application to the glycans from bovine thyroid-stimulating hormone. Anal Biochem 296:96–100.
Wilm M, Neubauer G, Mann M. 1996. Parent ion scans of
unseparated peptide mixtures. Anal Chem 68:527–533.
Wilm MS, Mann M. 1994. Electrospray and Taylor-cone theory,
Dole’s beam of macromolecules at last? Int J Mass Spectrom
Ion Processes 136:167–180.
Wolfender JL, Chu F, Ball H, Wolfender F, Fainzilber M, Baldwin
MA, Burlingame AL. 1999. Identification of tyrosine
sulfation in Conus pennaceus conotoxins alpha-PnIA and
alpha-PnIB: Further investigation of labile sulfo- and phosphopeptides by electrospray, matrix-assisted laser desorption/ionization (MALDI) and atmospheric pressure
MALDI mass spectrometry. J Mass Spectrom 34:447–454.
Wong AW, Wang H, Lebrilla CB. 2000. Selection of anionic
dopant for quantifying desialylation reactions with MALDIFTMS. Anal Chem 72:1419–1425.
Wong AW, Cancilla MT, Voss LR, Lebrilla CB. 1999. Anion
dopant for oligosaccharides in matrix-assisted laser desorption/ionization mass spectrometry. Anal Chem 71:205–
211.
Wu QY, Vanorden S, Cheng XH, Bakhtiar R, Smith RD. 1995.
Characterization of cytochrome-C variants with high-
226
resolution FTICR mass-spectrometry—Correlation of fragmentation and structure. Anal Chem 67:2498–2509.
Yagami T, Kitagawa K, Futaki S. 1995. Liquid secondary-ion
mass spectrometry of peptides containing multiple tyrosineO-sulfates. Rapid Commun Mass Spectrom 9:1335–1341.
Yagami T, Kitagawa K, Aida C, Fujiwara H, Futaki S. 2000.
Stabilization of a tyrosine O-sulfate residue by a cationic
functional group: Formation of a conjugate acid-base pair.
J Pept Res 56:239–249.
Yamada S, Yoshida K, Sugiura M, Sugahara K, Khoo KH,
Morris HR, Dell A. 1993. Structural studies on the bacterial
lyase-resistant tetrasaccharides derived from the antithrombin III-binding site of porcine intestinal heparin. J Biol
Chem 268:4780–4787.
Yamagaki T, Nakanishi H. 1999. A new technique distinguishing
alpha2-3 sialyl linkage from alpha2-6 linkage in sialyllactoses and sialyl-N-acetyllactosamines by post-source decay
fragmentation method of MALDI-TOF mass spectrometry.
Glycoconj J 16:385–389.
Yamagaki T, Nakanishi H. 2000. Post-source decay fragmentation analyses of linkage isomers of Lewis-type oligosaccharides in curved-field reflection matrix-assisted laser
desorption/ionization time-of-flight mass spectrometry:
Combined in-source decay/post-source decay experiments
and relative ion abundance analysis. J Mass Spectrom 35:
1300–1307.
Yang Y, Bergmann C, Benen J, Orlando R. 1997. Identification of
the glycosylation site and glycan structures of recombinant
endopolygalacturonase II by mass spectrometry. Rapid
Commun Mass Spectrom 11:1257–1262.
Yang HO, Gunay NS, Toida T, Kuberan B, Yu G, Kim YS,
Linhardt RJ. 2000. Preparation and structural determination
of dermatan sulfate-derived oligosaccharides. Glycobiology
10:1033–1039.
Yergey AL, Coorssen JR, Backlund PS, Jr, Blank PS, Humphrey
GA, Zimmerberg J, Campbell JM, Vestal ML. 2002. De novo
sequencing of peptides using MALDI/TOF-TOF. J Am Soc
Mass Spectrom 13:784–791.
Yu G, LeBrun L, Gunay NS, Hoppensteadt D, Walenga JM,
Fareed J, Linhardt RJ. 2000. Heparinase I acts on a synthetic
heparin pentasaccharide corresponding to the antithrombin
III binding site. Thromb Res 100:549–556.
Zaia J, McClellan JE, Costello CE. 2001. Tandem mass spectrometric determination of the 4S/6S sulfation sequence in
chondroitin sulfate oligosaccharides. Anal Chem 73:6030–
6039.
Zaia J, Costello CE. 2001. Compositional analysis of glycosaminoglycans by electrospray mass spectrometry. Anal
Chem 73:233–239.
Zaia J, Costello CE. 2003. Tandem mass spectrometry of sulfated
heparin-like glycosaminoglycan oligosaccharides. Anal
Chem 75:2445–2455.
Zaia J, Boynton RE, McIntosh A, Marshak DR, Olsson H,
Heinegård D, Barry FP. 1997. Post-translational modifications in cartilage oligomeric matrix protein. Characterization of the N-linked oligosaccharides by matrix-assisted
MASS SPECTROMETRY OF OLIGOSACCHARIDES
laser desorption ionization time-of-flight mass spectrometry. J Biol Chem 272:14120–14126.
Zaia J, Liu B, Boynton R, Barry F. 2000. Structural analysis of
cartilage proteoglycans and glycoproteins using matrixassisted laser desorption/ionization time-of-flight mass
spectrometry. Anal Biochem 277:94–103.
Zaia J, Boynton R, Heinegård D, Barry F. 2001. Posttranslational
modifications to human bone sialoprotein determined by
mass spectrometry. Biochemistry 40:12983–12991.
Zamfir A, Seidler DG, Kresse H, Peter-Katalinc J. 2002.
Structural characterization of chondroitin/dermatan sulfate
oligosaccharides from bovine aorta by capillary electrophoresis and electrospray ionization quadrupole time-offlight tandem mass spectrometry. Rapid Commun Mass
Spectrom 16:2015–2024.
Zamze S, Wing DR, Wormald MR, Hunter AP, Dwek RA, Harvey
DJ. 2001. A family of novel, acidic N-glycans in Bowes
melanoma tissue plasminogen activator have L2/HNK-1bearing antennae, many with sulfation of the fucosylated
chitobiose core. Eur J Biochem 268:4063–4078.
&
Zeng R, Xu Q, Shao XX, Wang KY, Xia QC. 1999.
Characterization and analysis of a novel glycoprotein from
snake venom using liquid chromatography-electrospray
mass spectrometry and Edman degradation. Eur J Biochem
266:352–358.
Zhou ZR, Ogden S, Leary JA. 1990. Linkage position determination in oligosaccharides—MS/MS study of lithiumcationized carbohydrates. J Org Chem 55:5444–5446.
Zhu X, Borchers C, Bienstock RJ, Tomer KB. 2000. Mass
spectrometric characterization of the glycosylation pattern
of HIV-gp120 expressed in CHO cells. Biochemistry
39:11194–11204.
Zubarev RA, Kelleher NL, McLafferty FW. 1998. Electron
capture dissociation of multiply charged protein cations. A
nonergodic process. J Am Chem Soc 120:3265–3266.
Zubarev RA, Kruger NA, Fridriksson EK, Lewis MA, Horn DM,
Carpenter BK, McLafferty FW. 1999. Electron capture
dissociation of gaseous multiply-charged proteins is favored
at disulfide bonds and other sites of high hydrogen atom
affinity. J Am Chem Soc 121:2857–2862.
227