structural analysis of carbohydrates by mass spectrometry

171
Purdue University Purdue e-Pubs Open Access Dissertations eses and Dissertations Fall 2013 Structural Analysis of Carbohydrates by Mass Spectrometry Chiharu Konda Purdue University Follow this and additional works at: hps://docs.lib.purdue.edu/open_access_dissertations Part of the Analytical Chemistry Commons is document has been made available through Purdue e-Pubs, a service of the Purdue University Libraries. Please contact [email protected] for additional information. Recommended Citation Konda, Chiharu, "Structural Analysis of Carbohydrates by Mass Spectrometry" (2013). Open Access Dissertations. 141. hps://docs.lib.purdue.edu/open_access_dissertations/141

Upload: others

Post on 04-Feb-2022

10 views

Category:

Documents


0 download

TRANSCRIPT

Page 1: Structural Analysis of Carbohydrates by Mass Spectrometry

Purdue UniversityPurdue e-Pubs

Open Access Dissertations Theses and Dissertations

Fall 2013

Structural Analysis of Carbohydrates by MassSpectrometryChiharu KondaPurdue University

Follow this and additional works at: https://docs.lib.purdue.edu/open_access_dissertations

Part of the Analytical Chemistry Commons

This document has been made available through Purdue e-Pubs, a service of the Purdue University Libraries. Please contact [email protected] foradditional information.

Recommended CitationKonda, Chiharu, "Structural Analysis of Carbohydrates by Mass Spectrometry" (2013). Open Access Dissertations. 141.https://docs.lib.purdue.edu/open_access_dissertations/141

Page 2: Structural Analysis of Carbohydrates by Mass Spectrometry

Graduate School ETD Form 9 (Revised 12/07)

PURDUE UNIVERSITY GRADUATE SCHOOL

Thesis/Dissertation Acceptance

This is to certify that the thesis/dissertation prepared

By

Entitled

For the degree of

Is approved by the final examining committee:

Chair

To the best of my knowledge and as understood by the student in the Research Integrity and Copyright Disclaimer (Graduate School Form 20), this thesis/dissertation adheres to the provisions of Purdue University’s “Policy on Integrity in Research” and the use of copyrighted material.

Approved by Major Professor(s): ____________________________________

____________________________________

Approved by: Head of the Graduate Program Date

Chiharu Konda

Structural Analysis of Carbohydrates by Mass Spectrometry

Doctor of Philosophy

Yu Xia

Peter T. Kissinger

Nikolai R. Skrynnikov

Hilkka I. Kenttamaa

Yu Xia

Robert E. Wild 10/25/2013

Page 3: Structural Analysis of Carbohydrates by Mass Spectrometry

i

STRUCTURAL ANALYSIS OF CARBOHYDRATES BY MASS SPECTROMETRY

A Dissertation

Submitted to the Faculty

of

Purdue University

by

Chiharu Konda

In Partial Fulfillment of the

Requirements for the Degree

of

Doctor of Philosophy

December 2013

Purdue University

West Lafayette, Indiana

Page 4: Structural Analysis of Carbohydrates by Mass Spectrometry
Page 5: Structural Analysis of Carbohydrates by Mass Spectrometry
Page 6: Structural Analysis of Carbohydrates by Mass Spectrometry
Page 7: Structural Analysis of Carbohydrates by Mass Spectrometry
Page 8: Structural Analysis of Carbohydrates by Mass Spectrometry
Page 9: Structural Analysis of Carbohydrates by Mass Spectrometry
Page 10: Structural Analysis of Carbohydrates by Mass Spectrometry

m/zm/z

Page 11: Structural Analysis of Carbohydrates by Mass Spectrometry
Page 12: Structural Analysis of Carbohydrates by Mass Spectrometry
Page 13: Structural Analysis of Carbohydrates by Mass Spectrometry

p m/z

m/z

m/zp p

m/zp

pp

p

Page 14: Structural Analysis of Carbohydrates by Mass Spectrometry

p

m/z

m/z pp m/z

m/zp m/z

m/zp

m/z m/z p

m/zm/z p

pp

pp

m/z p

m/z pp

m/z

Page 15: Structural Analysis of Carbohydrates by Mass Spectrometry

p

m/z

m/z

p

pp

ppp p

p

p

Page 16: Structural Analysis of Carbohydrates by Mass Spectrometry

m/z

m/z

m/z

p pp p p

p p p

m/z

m/z pp m/z

m/z pp p

p p pp

m/zm/z

p p pp

m/znative

m/z

Page 17: Structural Analysis of Carbohydrates by Mass Spectrometry

native p p p

p

p p

p p

p p

p

p

p p

Page 18: Structural Analysis of Carbohydrates by Mass Spectrometry

pp p

p p

pp p p

p

p pp p

pp pp

Page 19: Structural Analysis of Carbohydrates by Mass Spectrometry

m/z

Page 20: Structural Analysis of Carbohydrates by Mass Spectrometry
Page 21: Structural Analysis of Carbohydrates by Mass Spectrometry

p

Page 22: Structural Analysis of Carbohydrates by Mass Spectrometry
Page 23: Structural Analysis of Carbohydrates by Mass Spectrometry

m/z

Page 24: Structural Analysis of Carbohydrates by Mass Spectrometry

1

CHAPTER 1 INTRODUCTION

1.1 Carbohydrates

Carbohydrates (also called sugars, oligosaccharides, glycans) are defined as

polyhydroxyaldehydes, polyhydroxyketones and their simple derivatives, or larger

compounds that can be hydrolyzed into such units. A monosaccharide is the smallest unit

of carbohydrates which cannot be hydrolyzed into a simpler form. Free monosaccharides

can exist in ring or open-ring forms as shown in Figure 1.1. Changes in the orientation of

hydroxyl groups around specific carbon atoms generate new sugar molecules. For

example, mannose which has hydroxyl group on C2 is sticking up instead of down for

glucose and mannose is the C2 epimer of glucose. There are 3 chiral centers (C2-C4) and

8 (= 23) possible structures for D-monosaccharides and 16 possible structures (4 chiral

centers, C2-C5) if L-monosaccharides are included. The ring form of a monosaccharide

generates a chiral anomeric center at C1 for aldo sugars or at C2 for keto sugars. There

are two types of anomeric configurations, α and β, depending on the hydroxyl group is

sticking down or up, respectively.

Page 25: Structural Analysis of Carbohydrates by Mass Spectrometry

2

Figure 1.1 Ring and open-ring forms of D-glucose.

Any sugar that has aldehyde group or be able to form one in solution through

isomerism is called “reducing sugars” and can attach to another monosaccharide via the

hydroxyl group of anomeric center. A sugar which two monosaccharides connected

through the newly formed bond (“glycosidic bond”) is called a disaccharide. Due to the

enormous number of combinatory ways to construct an oligomer, even a small subunit of

carbohydrates, disaccharide, can possibly have more than 104 structural isomers

(estimated by Laine1 as shown in Figure 1.2). Unlike oligonucleotides and proteins

which are expressed in a linear fashion, carbohydrates can form branched structures. In

addition, the hydroxyl groups are subjected to various modifications (i.e. acetylamine,

sulfate, or sialic acid groups). This structural complexity imposes challenges to any

existing analytical methods for complete structural characterization. The heterogeneity

of many carbohydrate samples also requires separation tools such as high performance

liquid chromatography (HPLC) to be coupled with analysis methods including mass

spectrometry (MS) and nuclear magnetic resonance (NMR).

α-D-Glucose β-D-Glucose

1

2

3

45

6

12

3

45

61

2

3

4

5

6

Ring form Ring formOpen-ring form

Page 26: Structural Analysis of Carbohydrates by Mass Spectrometry

3

Figure 1.2 The number of possible structural isomers as increasing degree of polymerization (considering only aldohexoses and linear structure).

1.2 Glycosylation

Glycosylation is one of the most frequent post-translational modifications and

more than 50% of known proteins as well as 80% of membrane proteins are estimated to

be modified with glycans.2,3 Proteins and lipids modified with glycans are called

glycoproteins and glycolipids, respectively. Those glycan modifications are widely

involved in intermolecular and intercellular binding events from fertility to immunity.2 In

eukaryotic cells, there are two major classes of glycans according to the nature of the

linkage regions to the proteins: N- and O-linked glycans. An N-linked glycan (N-glycan)

is covalently linked to an asparagine residue of a peptide chain where the peptide

sequence matches with Asn-X-Ser or Asn-X-Thr (X could be any amino acid except

proline). N-glycans share a common core structure (Man3GlcNAc2) and can be divided

into three classes: high-mannose type, complex type, and hybrid type (Figure 1.3a).4 An

O-linked glycan (O-glycan) is typically linked to serine or threonine residue via N-

acetylgalactosamine (GalNAc). Unlike N-glycans, O-glycans have a variety of different

1.E+00

1.E+02

1.E+04

1.E+06

1.E+08

1.E+10

1.E+12

0 1 2 3 4 5#

of is

omer

sDegree of polymerization

Page 27: Structural Analysis of Carbohydrates by Mass Spectrometry

4

structural core classes. The most commonly occurring O-glycan structures are shown in

Figure 1.3b.4 Since the discovery of glycoproteins in bacteria and their pathogenicity

relates to the glycan structure, bacterial glycoproteins are attractive targets for therapeutic

intervention.5,6 However, prokaryotic glycans consist of unusual monosaccharide units

and their analysis is much more challenging than already complex eukaryotic glycans.5,7

The development of sensitive analytical method for structural characterization which

does not depend on the previous knowledge of biological glycan synthesis is highly

desirable.

Figure 1.3 Common core structures for (a) N- and (b) O-glycans.

Complex type Hybrid type High-mannose type Core 1 Core 2 Core 3 Core 4

Galactose (Gal)

Glucose (Glc)

Mannose (Man)

N-Acetylneuraminicacid (Neu5Ac)

N-Acetylgalactosamine(GalNAc)

N-Acetylglucosamine(GlcNAc)

(a) N-glycans

(b) O-glycans

Page 28: Structural Analysis of Carbohydrates by Mass Spectrometry

5

1.3 Overview of Glycan Structural Analysis

In order to understand the biological function of glycans, full-level of structural

information is necessary. Full-level of structural characterization of a glycan include the

following five levels: the identity (stereochemistry and modifications) of each

monosaccharide unit, the anomeric configuration of the glycosidic bonds, linkage

positions, branching location, and the sequence of the individual monosaccharides in the

oligomer (Figure 1.4).

Figure 1.4 Structural information necessary to fully characterize glycans.

NMR is the most powerful technique for structural analysis of glycans. The advantage of

NMR over other methods is that it offers the full-level of structural information in a non-

destructive way.8,9 The main drawback of this technique is that it typically requires large

amount of samples (μg to mg quantities) to characterize unknowns.9,10 In many cases,

such as the analysis of N- and O-linked glycans from glycoproteins, there is rarely

enough material available. The preferred techniques which are capable of smaller

quantity of samples (pg to ng) are enzymatic analysis or lectin affinity chromatography

Identity

AnomericConfiguration

Linkage

SequenceBranching location

Page 29: Structural Analysis of Carbohydrates by Mass Spectrometry

6

coupled with MS or HPLC. Enzymatic analysis of glycans uses exoglycosidases which

are known to cleave bonds that are specific to the type of linkage, anomeric

configuration, and stereochemistry of monosaccharide. Selection of exoglycosidase

largely relies on the knowledge of biological synthesis. Sialidase (cleave sialic acid

which has linkage of α2-3, -6, and -8), β-galactosidase (β1-4 galactose), α-fucosidase

(α1-2, -3, -4, -6 fucose), β-N-acetylhexosaminidase (β1-2, -3, -4, -6 GlcNAc), α-

mannosidase (α1-2, -3, -6 mannose) are some examples of exoglycosidases.11

Susceptibility to cleavage by different kinds of exoglycosidases is assessed by analysis of

the reaction products by MS or HPLC.12-14 This approach gives a lot of structural

information if prior knowledge about the glycan is available (e.g. the glycan is an N-

glycan).15,16 However, the variation of exoglycosidases which have been isolated so far is

very limited to N-glycans. Especially most of exoglycosidases cleave multiple linkages,

thus linkage information cannot be obtained for completely unknown molecule. Lectins

are carbohydrate-binding proteins which recognize specific sugar moieties.17 Based on

the relative affinity of glycans to different lectins, each glycan is eluted differently in

lectin affinity chromatography and their structures can be identified.18,19 A variety of

lectins has been discovered and a series of lectin columns as well as lectin array is

available for glycan analysis. These methods allow high-throughput analysis. However,

the detail glycan structure is difficult to obtain since lectins recognize whole sugar

moieties instead of a single sugar unit resolution.

Page 30: Structural Analysis of Carbohydrates by Mass Spectrometry

7

1.4 Structural Analysis of Glycans by Mass Spectrometry

Mass spectrometry was first applied to carbohydrate analysis in 195820 with

electron impact (EI) ionization.21,22 EI is said to be “hard” ionization method since it

induces extensive fragmentation and the molecular ions are typically not observed. Since

EI is only suitable for volatile organic molecules, hydroxyl groups in non-volatile

carbohydrates had to be protected by permethylation,23 or peracetylation24 to increase the

volatility prior to analysis.25 Since then, many researches for structural analysis of

glycans using mass spectrometry have been reported with the improvement of analytical

techniques, mass spectrometric instrumentation, as well as development of “soft”

ionization methods such as fast atom bombardment (FAB),26 electrospray ionization

(ESI)27,28 and matrix-assisted laser desorption ionization (MALDI)29,30 which enable

native carbohydrates to be ionized with molecular ion information to be obtained.

1.4.1 Ionization

FAB has been introduced in 198126 and was the first ionization method used in

MS experiment of carbohydrates. A mechanism of FAB is similar to EI, instead of

impacting the sample by electrons, FAB uses beam of atoms and the sample has to be in

the liquid matrix.26 The intact non-volatile compounds were ionized by FAB, especially

acidic oligosaccharides (containing Neu5Ac or sulfate) in negative ion mode worked

relatively well. However, the ion signal was not strong enough and the size of the

molecular weight applicable was typically low.31 Currently, FAB has been replaced by

later developed “soft” ionization methods: ESI and MALDI.

Page 31: Structural Analysis of Carbohydrates by Mass Spectrometry

8

ESI and MALDI are two common methods for the analysis of carbohydrates.

Both ionization techniques were introduced in the late 1980’s and enabled non-volatile

and thermally labile compounds to be ionized. In ESI, the sample solution is highly

charged and sprayed through a capillary into a strong electric field to form a fine mist of

charged droplets. The solvent evaporates and produces gas-phase ions or solvated ions. In

order to be efficiently ionized, analytes are necessary to stay on the surface of the

droplets. However, the hydrophilicity of carbohydrates limits their surface activity in ESI

droplets, and thus ionization efficiency of carbohydrates is significantly lower than those

of peptides and proteins.32,33 NanoESI,34 which uses much lower flow rates down to some

tens of nL/min as compared to 1 to 10 μL/min for regular ESI, produces smaller charged

droplet sizes and enhances the ionization efficiency for carbohydrates.35

MALDI uses a matrix which contains small organic molecules that have strong

absorption at the laser wavelength for ionization. Before the analysis, analytes are

dissolved in the matrix, which are subsequently dried to form crystals. Intense laser pulse

ablates the surface of the dried crystal in vacuum. This ablation excites and sublimates

the matrix molecule into gas phase.36 Although the exact mechanism of the MALDI

process is still under debate,37,38 the widely accepted theory for ion formation involves

proton transfer in the solid phase before desorption. Ions in the gas phase are then

accelerated by the electric field towards the mass analyzer.38 Both ESI and MALDI

ionization methods have advantages and disadvantages for carbohydrate analysis. The

advantage of MALDI is that the ionization efficiency for neutral carbohydrates is

relatively constant as the size of the molecule increases, in contrast to ESI, where the

ionization efficiency decreases with an increasing molecular weight.39 The limitation of

Page 32: Structural Analysis of Carbohydrates by Mass Spectrometry

9

MALDI comes from extensive fragmentation due to the higher internal energies

deposited to the ions by laser ablation as compared to ESI.

1.4.2 Derivatization Techniques for Glycan Analysis

Derivatization of sugars, such as permethylation of hydroxyl groups23,40 and

reducing end modification,41,42 reduces hydrophilicity of the sugar molecules and

increases the ionization efficiency significantly.43,44 Permethylation is the most widely

used modification for glycan analysis in mass spectrometry, especially after the

introduction of the solid phase permethylation method developed by Novotny

group.40,45,46 Solid phase permethylation allows simple and efficient permethylation

especially for small quantity of samples. The aldehyde group at the reducing end of

glycans can also react with alkylamines (typically aromatic amines, which also function

as chromophore for UV detection upon separation). Two approaches (amination and

reductive amination) of reducing end derivatization are summarized in Figure 1.5. Schiff

base formation by amination produces open-ring and closed-ring structures in

equilibrium. It has been confirmed by NMR and other techniques that closed-ring

(glycosylamine) structure is prevalent in solution with aromatic substitutions.47,48 The

Schiff base can be further reduced by sodium borohydride to form reduced Schiff base

which can only exist in the open-ring structure. While reduced Schiff base is more stable

than a glycosylamine, glycosylamine structure has been reported with analytical

advantages such as better HPLC separation and enhancement of cross-ring cleavages by

CID.47,49 A number of amine groups have been used as reducing end derivatives

containing cationic and anionic charges such as p-aminophenyl ammonium chloride

Page 33: Structural Analysis of Carbohydrates by Mass Spectrometry

10

(TMAPA)50 and 5-aminosalicylic acid.33 TMAPA was shown to increase sensitivity

5000-fold under ESI condition relative to the native form oligosaccharides.50 These

charge-containing derivatizations do not only offer improved ionization efficiency but

also show positive effects on the fragmentation pattern of CID.51 Reducing end

derivatization (including isotopic labeling, i.e. 18O) can also be used to simplify the

complex fragmentation spectrum from CID by introducing mass differences of product

ions derived from the reducing end vs. non-reducing end. 41,42

Figure 1.5 Reaction scheme for amination and reductive amination.

1.4.3 Nomenclature for the Gas-phase Fragmentation of Oligosaccharides

The nomenclature for oligosaccharide fragmentation commonly used in the mass

spectrometry field has been introduced by Domon and Costello as shown in Figure 1.6.52

Fragment ions that contain a non-reducing end are labeled with uppercase letters A, B,

H2N-R2(-H2O)

open-ring(Schiff base)

closed-ring(glycosylamine)

reduction

Non-Reduced (NR) formby amination

open-ring(reduced Schiff base)

Reduced (R) formby reductive amination

open-ring

closed-ring (hemiacetal)

Native Glycan Labeled Glycan

Page 34: Structural Analysis of Carbohydrates by Mass Spectrometry

11

and C, and those contain the reducing-end of the oligosaccharide or the aglycon are

labeled with X, Y, and Z; subscripts indicate the location of cleavage within the

oligosaccharide ion. B and Y ions are cleaved at the non-reducing side of a glycosidic

oxygen and C and Z ions are cleaved at the reducing side of a glycosidic oxygen. B, C, Y,

and Z ions are classified as “glycosidic bond cleavages.” A and X ions resulted from

cleavages across the glycosidic ring are termed as “cross-ring cleavages.” They are

labeled with superscript of two broken ring bond numbers that are assigned as shown in

Figure 1.6.

Figure 1.6. Nomenclature for glycoconjugate product ions generated by tandem MS.52

1.4.4 Structural Analysis of Oligosaccharides by Mass Spectrometry

Mass spectrometry is a widely applied method in structural analysis of

carbohydrates, due to its high sensitivity and capability of providing detailed molecular

information.53 However, obtaining full-level of structural information (sugar identity,

anomeric configuration, linkage position, sequence, and branching location) has not been

Y2 Z2 Y1 Z1 Y0 Z0

B1 C1 B2 C2 B3 C30,2A1

1,5X0

Reducing endNon-reducing end 0

12

34 5

0

12

34 5

Page 35: Structural Analysis of Carbohydrates by Mass Spectrometry

12

possible by using only MS. The molecular weight information of carbohydrates is readily

obtained from soft ionization methods such as ESI33,54 and MALDI.29,39,55 Tandem mass

spectrometry (MS/MS) based on collision-induced dissociation (CID) is heavily relied on

to obtain structural information for carbohydrates. Glycosidic bond cleavages and/or

cross-ring cleavages are typically observed from collisional activation. Glycosidic bonds

are labile and can be easily cleaved under low-energy collisions. These fragments are

useful for sequencing and identifying branching location. On the other hand, cross-ring

cleavages need relatively higher-energy to induce and these fragments are useful for

identifying linkage positions. Sequence and branching location can be readily obtained

from MS2 CID of permethylated33,56-59 or peracetylated59,60 oligosaccharides in the

positive ion mode and from native oligosaccharide in the negative ion mode.61

Assignment of linkage positions between disaccharides can be obtained based on the

product ions generated by MS2 CID in negative ion mode49,62-66 and positive ion mode

with metal cation adducts.60,67-73 Linkage determination by MS2 CID has been applied to

oligosaccharides. As the molecular gets larger, it is more difficult for linkage

determination due to possible ion suppression of the diagnostic product ions for structural

determination.63,74 Stereochemistry of monosaccharides and anomeric configuration are

the most difficult structural information to obtain by MS. Notable examples to obtain

stereo-structure information include tandem mass spectrometry (MS2) of transition metal

chelated75 and amino acid complexed monosaccharides.76

Page 36: Structural Analysis of Carbohydrates by Mass Spectrometry

13

1.4.5 MS2 vs MSn

Tandem mass spectrometry (MS/MS) is the key technique for structural analysis

by MS.77-81 MS2, the simplest MS/MS, has been routinely used for sequencing and

identifying branching locations by a series of product ions from glycosidic bond

cleavages. In order to identify linkage positions unambiguously, obtaining all of the

possible product ions from cross-ring cleavages is the key. Besides CID, different

activation/dissociation methods have been investigated, including infrared multiphoton

dissociation (IRMPD),82 electron capture dissociation (ECD),83,84 electron-induced

dissociation (EID),85 electron detachment dissociation (EDD),86 and electron excited

dissociation (EED).87 These methods have been shown to increase cross-ring cleavages

and are complementary to CID. However, the application of the above methods has been

limited to research groups which have these instrument capabilities. Overall, MS2 has

advantages of good sensitivity (pmol to fmol of sample quantiteis),88 high throughput,

and simplicity in performance and instrumentation. Depending on the nature of analytes

and the dissociation method used, key diagnostic ions (i.e. A type ions) may be missing

in MS2, leading to either miss-assigned or un-assigned linkage positions or other

structural information.63,74,89 Higher stages of tandem mass spectrometry (MSn, where

n>2), enables a sequential disassembly of the intact molecule, which can be back-tracked

based on the precursor-product relationships at each step. In general, MSn leads to higher

confidence in structural characterization, and is extremely useful in probing subtle

structure changes between isomers.57,90-92 Viseux et al. demonstrated the use of MS3 to

MS5 CID for structural characterization (sequence, linkage, and branching determination)

of linear and branched oligosaccharides. In their work, an unknown oligosaccharide was

Page 37: Structural Analysis of Carbohydrates by Mass Spectrometry

14

first fragmented into disaccharide and/or trisaccharide substructures via CID. These

substructures were further subjected to CID and the fragmentation spectra were compared

to that from known reference molecules for linkage determination. By using the

substructures generated in MSn CID as opposed to the whole oligosaccharide, it

significantly reduced the size of reference molecules needed for comparisons and it

overcame the low mass cut off issue for detecting key ions with the use of

electrodynamic ion trap mass spectrometer.58 Furthermore, determining the identity of

low mass ions through comparison to a much more limited number of possible structural

isomers provides greater confidence in their structural determination, particularly in

discriminating between their unique stereochemical or regiochemical isomers.

1.4.6 A New MSn Approach for High Level Structural Analysis of Oligosaccharides

One of the grand challenges of oligosaccharide analysis is to develop a complete

approach based on MS toward a full level structural analysis of linear oligosaccharides by

obtaining information of sugar unit identity, anomeric configuration, linkage types and

sequence. We propose a new MSn approach to achieve this goal by 4 steps: (1) find a

diagnostic ion which gives any of the structural information listed above in the product

ions from MS3 CID of disaccharide ions (shown in the blue square in Scheme 1.1), (2)

make a standard spectral library by collecting MSn (n = 2 or 3) CID of the diagnostic ions

from all of the possible isomeric structures which was either synthesized or fragmented in

gas-phase from disaccharide ions, (3) find the optimized pathway to fragment down

oligosaccharide ions into overlapping disaccharide substructures (shown in the red square

in Scheme 1.1), and (4) employ spectral matching algorithm for easy and fair comparison.

Page 38: Structural Analysis of Carbohydrates by Mass Spectrometry

15

Scheme 1.1 The MSn (n = 4 or 5) approach for detailed structural analysis of a linear oligosaccharide. Charges are not indicated on the structure. “M” stands for reducing end modification.

Figure 1.7 MS3 CID steps for sugar unit identity and anomeric configuration determination.93

HO 3 2 1O O M

3OOO4 2 1O O MMS1

MS3

MS3 MS4 MS3

MS4 MS4MS5

Diagnostic Ion

Disaccharide Substructure FormationHO

Disac3O OHOO4

cture Fcture F2O OHHO

SubstrO3 HO

n2 1O M

Y3 Y2

MS2

MS2MS2

C2

Anomeric Config.Identity Location Linkage

II. Disaccharide Ion Diagnostic Ion Structural Information

I. Oligosaccharide Ion Disaccharide Substructure Ions

-H -Sugar-GA (m/z 221)

Diagnostic Ion

MS3 CID

131

101113 22187 161

203

131

101113 22187 161

203

Sugar Identity Anomericity

Unknown Disaccharide-H -

MS2 CID

Page 39: Structural Analysis of Carbohydrates by Mass Spectrometry

16

Finding the diagnostic ion which is a substructure of disaccharide ion is the key

concept of this approach since the number of structural isomers can be minimized and

significantly reduce the cost of library construction. Previous study by our collaborator,

Bendiak group, has demonstrated that the CID patterns of m/z 221 product ions (C8H13O7,

non-reducing sugar glycosidically linked to a glycolaldehyde (sugar-GA)) from

disaccharide anions can be used to determine the non-reducing sugar identity and

anomeric configuration (Figure 1.7).93,94 Using this m/z 221 diagnostic ion successfully

reduces the size of synthetic standards from 11520 possible structures (disaccharides) to

24 (monosaccharides) as shown in Figure 1.2. By following this concept, further

discovery of diagnostic ion for linkage type is expected. Next step is the application of

this structural analysis method to oligosaccharides. Oligosaccharides need to be broken

down into overlapping disaccharide substructures. This can be done by MSn (n = 2 or 3)

via fragmenting the oligosaccharide ions into a ladder of Yn-1 to Y2 ions, then, cleaving

off disaccharide substructures (C2 ion) from non-reducing end of each of the Y ions. MSn

(n>2) is typically performed using ion trap (Paul trap, linear ion trap, and Fourier

transform ion cyclotron resonance (FTICR)) in a tandem-in-time fashion. In practice, the

number of stages (n) of MS/MS that can be executed on a certain instrument is limited by

the number of ions remained after each step of ion activation and isolation.95 The loss of

the low abundance fragment ion could be significant for ion traps due to the space charge

effect.96,97 Due to the above reason, a systematic study of fragmentation chemistry of

oligosaccharides and the way to enhance the formation of desired ions are important as

well to accomplish multiple stages of MSn (n = 4 or 5) experiment.

Page 40: Structural Analysis of Carbohydrates by Mass Spectrometry

17

Previously, Bendiak group successfully demonstrated the stereo-structure analysis

by the m/z 221 diagnostic ion for 1-2, 1-4, and 1-6 linked disaccharides. However, 1-3

linked disaccharides produced very few m/z 221 ions using ion trap CID, making it

difficult to perform MS3 analysis on the stand-alone ion trap instrument. We investigated

the unimolecular dissociation chemistry of 1-3 linked disaccharide anions under different

CID conditions to understand fundamental gas-phase ion chemistry for the formation of

m/z 221 diagnostic ion in Chapter 2. In Chapter 3, the MSn approach is discussed to

generate m/z 221 within linear oligosaccharides to obtain the stereo-structure information

of each sugar unit. Furthermore, we explored and found the new diagnostic ion for

linkage determination which is smaller than a monosaccharide, Z1 ion, and this Z1 ion

was used to obtain linkage information in linear and branched oligosaccharides in

Chapter 4. Finally, a variety of reducing end modification groups and their significance

on the fragmentation chemistry of oligosaccharides were studied in Chapter 5 in order to

enhance selective bond cleavages, which is critical in improving the sensitivity of the

overall MSn analysis.

Page 41: Structural Analysis of Carbohydrates by Mass Spectrometry

18

1.5 Conclusions

MS has been extensively used in glycan analysis both independently and coupled

with other analytical techniques. Due to its high sensitivity and simplicity, it is highly

expected that the full-level characterization can be obtained only using MS. The current

limitation of MS method is that linkage position cannot be obtained from

oligosaccharides unambiguously under most of circumstances, as well as obtaining

stereo-structure (stereochemistry and anomeric configuration) information. As opposed to

the trend in this field using MS2, we focused on the development of MSn approach which

overcomes the current weakness of MS method on glycan analysis, obtaining

unambiguous information of stereochemistry, anomeric configuration, and linkage

position from oligosaccharides.

Page 42: Structural Analysis of Carbohydrates by Mass Spectrometry

19

1.6 References (1) Laine, R. A., Glycobiology 1994, 4, 759-767. (2) Perkel, J. M., Science 2011, 331, 95-97. (3) Apweiler, R.; Hermjakob, H.; Sharon, N., Biochim. Biophys. Acta 1999, 1, 4-8. (4) Varki, A.; Cummings, R. D.; Esko, J. D.; Freeze, H. H.; Stanley, P.; Bertozzi, C. R.; Hart, G. W.; Etzler, M. E., Essentials of glycobiology. Cold Spring Harbor Laboratory Press: Cold Spring Harbor, NY, 2009. (5) Dube, D. H.; Champasa, K.; Wang, B., Chem. Commun. 2011, 47, 87-101. (6) Balonova, L.; Hernychova, L.; Bilkova, Z., Expert Rev. Proteomics 2009, 6, 75-85. (7) Maki, M.; Renkonen, R., Glycobiology 2004, 14, R1-R15. (8) Vliegenthart, J. F. G.; Dorland, L.; Halbeek, H. v., High-Resolution, 1H-Nuclear Magnetic Resonance Spectroscopy as a Tool in the Structural Analysis of Carbohydrates Related to Glycoproteins. In Advances in Carbohydrate Chemistry and Biochemistry, Tipson, R. S.; Derek, H., Eds. Academic Press: 1983; Vol. 41, pp 209-374. (9) Duus, J. Ø.; Gotfredsen, C. H.; Bock, K., Chem. Rev. 2000, 100, 4589-4614. (10) Allen, S.; Richardson, J. M.; Mehlert, A.; Ferguson, M. A. J., J. Biol. Chem. 2013, 288, 11093-11105. (11) Hoja-Łukowicz, D.; Link-Lenczowski, P.; Carpentieri, A.; Amoresano, A.; Pocheć, E.; Artemenko, K.; Bergquist, J.; Lityńska, A., Glycoconj J 2013, 30, 205-225. (12) Guile, G. R.; Rudd, P. M.; Wing, D. R.; Prime, S. B.; Dwek, R. A., Anal. Biochem. 1996, 233, 205-211. (13) Rudd, P. M.; Gulle, G. R.; Kuster, B.; Harvey, D. J.; Opdenakker, G.; Dwek, R. A., Nature 1997, 388, 205-207. (14) Edge, C. J.; Rademacher, T. W.; Wormald, M. R.; Parekh, R. B.; Butters, T. D.; Wing, D. R.; Dwek, R. A., Proc. Natl. Acad. Sci. U. S. A. 1992, 89, 6338-6342.

Page 43: Structural Analysis of Carbohydrates by Mass Spectrometry

20

(15) Mechref, Y.; Novotny, M. V., Anal. Chem. 1998, 70, 455-463. (16) Geyer, H.; Schmitt, S.; Wuhrer, M.; Geyer, R., Anal. Chem. 1998, 71, 476-482. (17) Sharon, N.; Lis, H., Science 1989, 246, 227-234. (18) Cummings, R. D., Methods Enzymol. 1994, 230, 66-86. (19) Endo, T., J. Chromatogr. A 1996, 720, 251-261. (20) Finan, P. A.; Reed, R. I.; Snedden, W., Chem. Ind. (London) 1958, 36, 1172. (21) Bleakney, W., Phys. Rev. 1929, 34, 157-160. (22) Nier, A. O., Rev. Sci. Instrum. 1947, 18, 398-411. (23) Ciucanu, I.; Kerek, F., Carbohydr. Res. 1984, 131, 209-217. (24) Bourne, E. J.; Stacey, M.; Tatlow, J. C.; Tedder, J. M., J. Chem. Soc. 1949, 2976-2979. (25) Finan, P. A.; Reed, R. I., Nature 1959, 184, 1866. (26) Barber, M.; Bordoli, R. S.; Sedgwick, R. D.; Tyler, A. N., Nature 1981, 293, 270-275. (27) Meng, C. K.; Mann, M.; Fenn, J. B., Z. Phys. D. 1988, 10, 361-368. (28) Fenn, J. B.; Mann, M.; Meng, C. K.; Wong, S. F., Mass Spectrom. Rev. 1990, 9, 37-70. (29) Karas, M.; Bachmann, D.; Bahr, U.; Hillenkamp, F., Int. J. Mass Spectrom. 1987, 78, 53-68. (30) Karas, M.; Hillenkamp, F., Anal. Chem. 1988, 60, 2229-2231. (31) Egge, H.; Peter-Katalinić, J., Mass Spectrom. Rev. 1987, 6, 331-393.

Page 44: Structural Analysis of Carbohydrates by Mass Spectrometry

21

(32) Burlingame, A. L.; Boyd, R. K.; Gaskell, S. J., Anal. Chem. 1994, 66, 634R-683R. (33) Reinhold, V. N.; Reinhold, B. B.; Costello, C. E., Anal. Chem. 1995, 67, 1772-1784. (34) Wilm, M. S.; Mann, M., Int. J. Mass Spectrom. Ion Processes 1994, 136, 167-180. (35) Bahr, U.; Pfenninger, A.; karas, M.; Stahl, B., Anal. Chem. 1997, 69, 4530-4535. (36) Dreisewerd, K., Chem. Rev. 2003, 103, 395-425. (37) Zenobi, R.; Knochenmuss, R., Mass Spectrom. Rev. 1998, 17, 337-366. (38) Knochenmuss, R.; Zenobi, R., Chem. Rev. 2003, 103, 441-452. (39) Harvey, D. J., Rapid Commun. Mass Spectrom. 1993, 7, 614-619. (40) Mechref, Y.; Kang, P.; Novotny, M. V., Methods in Mol. Biol. 2009, 534, 53-64. (41) Harvey, D. J., J. Am. Soc. Mass Spectrom. 2000, 11, 900-915. (42) Harvey, D. J., J. Chromatogr. B 2011, 879, 1196-1225. (43) Karas, M.; Bahr, U.; Dulcks, T., Fresenius J. Anal. Chem. 2000, 366, 669-676. (44) Dell, A., Methods Enzymol. 1990, 193, 647-660. (45) Kang, P.; Mechref, Y.; Klouckova, I.; Novotny, M. V., Rapid Commun. Mass Spectrom. 2005, 19, 3421-3428. (46) Kang, P.; Mechref, Y.; Novotny, M. V., Rapid Commun. Mass Spectrom. 2008, 22, 721-734. (47) Her, G. R.; Santikarn, S.; Reinhold, V. N.; Williams, J. C., 1987, 6, 129-139. (48) Nishikaze, T.; Kaneshiro, K.; Kawabata, S.; Tanaka, K., Anal. Chem. 2012, 84, 9453-9461. (49) Li, D. T.; Her, G. R., Anal. Biochem. 1993, 211, 250-257.

Page 45: Structural Analysis of Carbohydrates by Mass Spectrometry

22

(50) Hansson, G. C.; Karlsson, H., Methods Molec. Biol. Humana Press: Totowa, 1993. (51) Chen, S.-T.; Her, G.-R., J. Am. Soc. Mass Spectrom. 2012, 23, 1408-1418. (52) Domon, B.; Costello, C. E., Glycoconjugate J. 1988, 5, 397-409. (53) Zaia, J., Mass Spectrom. Rev. 2004, 23, 161-227. (54) Fenn, J.; Mann, M.; Meng, C.; Wong, S.; Whitehouse, C., Science 1989, 246, 64-71. (55) Tanaka, K.; Waki, H.; Ido, Y.; Akita, S.; Yoshida, Y.; Yoshida, T.; Matsuo, T., Rapid Commun. Mass Spectrom. 1988, 2, 151-153. (56) Reinhold, V. N.; Sheeley, D. M., Anal. Biochem. 1998, 259, 28-33. (57) Sheeley, D. M.; Reinhold, V. N., Anal. Chem. 1998, 70, 3053-3059. (58) Viseux, N.; de Hoffmann, E.; Domon, B., Anal. Chem. 1998, 70, 4951-4959. (59) Perreault, H.; Costello, C. E., J. Mass Spectrom. 1999, 34, 184-197. (60) Domon, B.; Müller, D. R.; Richter, W. J., Biol. Mass Spectrom. 1990, 19, 390-392. (61) Chai, W.; Lawson, A.; Piskarev, V., J. Am. Soc. Mass Spectrom. 2002, 13, 670-679. (62) Ballistreri, A.; Montaudo, G.; Garozzo, D.; Giuffrida, M.; Impallomeni, G.; Daolio, S., Rapid Commun. Mass Spectrom. 1989, 3, 302-304. (63) Garozzo, D.; Giuffrida, M.; Impallomeni, G.; Ballistreri, A.; Montaudo, G., Anal. Chem. 1990, 62, 279-286. (64) Lamb, D. J.; Wang, H. M.; Mallis, L. M.; Linhardt, R. J., J. Am. Soc. Mass Spectrom. 1992, 3, 797-803. (65) Carroll, J. A.; Willard, D.; Lebrilla, C. B., Anal. Chim. Acta 1995, 307, 431-447. (66) Mulroney, B.; Traeger, J. C.; Stone, B. A., J. Mass Spectrom. 1995, 30, 1277-1283.

Page 46: Structural Analysis of Carbohydrates by Mass Spectrometry

23

(67) Domon, B.; Müller, D. R.; Richter, W. J., Org. Mass Spectrom. 1989, 24, 357-359. (68) Domon, B.; Muller, D. R.; Richter, W. J., 1990, 100, 301-311. (69) Zhou, Z.; Ogden, S.; Leary, J. A., J. Org. Chem. 1990, 55, 5444-5446. (70) Hofmeister, G. E.; Zhou, Z.; Leary, J. A., J. Am. Chem. Soc. 1991, 113, 5964-5970. (71) Lemoine, J.; Fournet, B.; Despeyroux, D.; Jennings, K. R.; Rosenberg, R.; de Hoffmann, E., J. Am. Soc. Mass Spectrom. 1993, 4, 197-203. (72) König, S.; Leary, J., J. Am. Soc. Mass Spectrom. 1998, 9, 1125-1134. (73) Xue, J.; Song, L.; Khaja, S. D.; Locke, R. D.; West, C. M.; Laine, R. A.; Matta, K. L., Rapid Commun. Mass Spectrom. 2004, 18, 1947-1955. (74) Dallinga, J. W.; Heerma, W., Biol. Mass Spectrom. 1991, 20, 215-231. (75) Gaucher, S. P.; Leary, J. A., Anal. Chem. 1998, 70, 3009-3014. (76) Augusti, D. V.; Carazza, F.; Augusti, R.; Tao, W. A.; Cooks, R. G., Anal. Chem. 2002, 74, 3458-3462. (77) Cooks, R. G.; Beynon, J. H.; Caprioli, R. M., Metastable Ions. Elsevier: Amsterdam, 1973. (78) Burinsky, D. J.; Cooks, R. G.; Chess, E. K.; Gross, M. L., Anal. Chem. 1982, 54, 295-299. (79) McLafferty, F. W., Tandem Mass Spectrometry. John Wiley: New York, 1983. (80) Busch, K. L.; Glish, G. L.; McLuckey, S. A., Mass Spectrometry / Mass Spectrometry: Techniques and Applications in Tandem Mass Spectrometry. VCH: New York, 1988. (81) Burinsky, D. J.; Cooks, R. G.; Chess, E. K.; Gross, M. L., Anal. Chem. 1982, 54, 295-299. (82) Xie, Y.; Lebrilla, C. B., Anal. Chem. 2003, 75, 1590-1598.

Page 47: Structural Analysis of Carbohydrates by Mass Spectrometry

24

(83) Adamson, J. T.; Håkansson, K., Anal. Chem. 2007, 79, 2901-2910.

(84) Zhao, C.; Xie, B.; Chan, S.-Y.; Costello, C.; O’Connor, P., J. Am. Soc. Mass Spectrom. 2008, 19, 138-150.

(85) Wolff, J.; Laremore, T.; Aslam, H.; Linhardt, R.; Amster, I. J., J. Am. Soc. Mass Spectrom. 2008, 19, 1449-1458.

(86) Wolff, J.; Amster, I. J.; Chi, L.; Linhardt, R., J. Am. Soc. Mass Spectrom. 2007, 18, 234-244.

(87) Yu, X.; Huang, Y.; Lin, C.; Costello, C. E., Anal. Chem. 2012, 84, 7487-7494.

(88) Geyer, H.; Geyer, R., Biochim. Biophys. Acta 2006, 1764, 1853-1869.

(89) Dongre, A. R.; Wysocki, V. H., Org. Mass Spectrom. 1994, 29, 700-702.

(90) Li, B.; An, H.; Hedrick, J.; Lebrilla, C., Methods Mol. Biol. 2009, 534, 133-145.

(91) Ashline, D. J.; Lapadula, A. J.; Liu, Y.-H.; Lin, M.; Grace, M.; Pramanik, B.; Reinhold, V. N., Anal. Chem. 2007, 79, 3830-3842.

(92) Weiskopf, A. S.; Vouros, P.; Harvey, D. J., Anal. Chem. 1998, 70, 4441-4447.

(93) Fang, T. T.; Bendiak, B., J. Am. Chem. Soc. 2007, 129, 9721-9736.

(94) Fang, T. T.; Zirrolli, J.; Bendiak, B., Carbohydr. Res. 2007, 342, 217-235.

(95) McLuckey, S. A.; Glish, G. L.; Vanberkel, G. J., Int. J. Mass Spectrom., Ion Processes 1991, 106, 213-235.

(96) Cox, K. A.; Cleven, C. D.; Cooks, R. G., Int. J. Mass Spectrom., Ion Processes 1995, 144, 47-65.

(97) March, R. E., J. Mass Spectrom. 1997, 32, 351-369.

Page 48: Structural Analysis of Carbohydrates by Mass Spectrometry

25

CHAPTER 2 DIFFERENTIATION OF THE STEREOCHEMISTRY AND ANOMERIC CONFIGURATION FOR 1-3 LINKED DISACCHARIDES

2.1 Introduction

A tandem mass spectrometry approach has been developed to differentiate the

stereochemistry and anomeric configuration for the non-reducing unit of hexose-

containing disaccharides having any of the 16 possible stereochemical variants.1,2 In this

method, diagnostic ions at m/z 221 were formed from CID of deprotonated disaccharide

ions (m/z 341). It was established that the m/z 221 ions consisted of the intact non-

reducing sugar glycosidically linked to glycolaldehyde, as indicated in Scheme 2.1

(where GA abbreviates glycolaldehyde). Note that an open-chain form for the reducing

sugar is indicated in Scheme 2.1 and also for other disaccharides discussed later. This is

based on the observation of absorbance in the carbonyl stretch region in variable

wavelength infrared radiation photo-dissociation of deprotonated monosaccharide anions

in the gas phase.3 When m/z 221 ions were further dissociated by collisional activation,

disaccharides having different non-reducing sugar units and anomeric configurations

showed distinct fragmentation patterns that matched synthetic glycosyl-GAs. This

method was shown to be useful for assigning the stereochemistry as well as the anomeric

configuration of the glycosidic bond for the non-reducing sugar in disaccharides having

1-2, 1-4, and 1-6 linkages.1,2 However, due to the low abundance of m/z 221 ions

Page 49: Structural Analysis of Carbohydrates by Mass Spectrometry

26

produced from 1-3 linked disaccharides, MS3 CID of m/z 221 ions could not be

performed, and it was unclear whether the fragmentation patterns could be used for

assigning either their stereochemistry or anomeric configuration.

Scheme 2.1 Formation of glycosyl-GA anions at m/z 221 from CID of deprotonated 1-4 linked disaccharides.

In this chapter, a series of 1-3 linked disaccharides were studied on a hybrid triple

quadrupole-linear ion trap mass spectrometer (QTRAP 4000). MS3 CID data of m/z 221

ions from the 1-3 linked disaccharides were obtained for the first time. The formation of

m/z 221 ions was examined using different collisional activation methods, i.e. beam-type

CID and on-resonance ion trap CID of the deprotonated disaccharides. 18O-labeling of the

reducing carbonyl oxygen in 1-3 linked disaccharides was used to enable mass-

discrimination of structural isomers of the (usually) m/z 221 ions. By choosing the proper

CID conditions, the diagnostic m/z 221 ions (the glycosyl-GAs) could be formed as the

dominant isomer. Their CID fragmentation patterns could be used to establish the

stereochemistry and anomeric configuration of the non-reducing sugar unit from 1-3

linked disaccharides.

-H --H -

α-D-Glcp-(1-4)-Glc, m/z 341 α-D-Glcp-GA, m/z 221

CID CID

Page 50: Structural Analysis of Carbohydrates by Mass Spectrometry

27

2.2 Experimental

2.2.1 Materials

A list of 7 disaccharides (6 reducing sugars and 1 non-reducing sugars) is shown

in Table 2.1. All samples were purchased from commercial sources (indicated by the

superscripts in Table 2.1) and used without further purification. H218O was purchased

from Sigma-Aldrich, Inc. (St. Louis, MO). α- and β-D-monosaccharide-glycolaldehyde

standards, glucopyranosyl-glycolaldehydes (Glcp-GA), galactopyranosyl-

glycolaldehydes (Galp-GA), and mannopyranosyl-glycolaldehydes (Manp-GA) were

synthesized as previously described.1 Disaccharides and synthetic standards were

dissolved in methanol to a final concentration of 0.01 mg/mL and NH4OH was added to a

final concentration of 1 % immediately before use.

Table 2.1 List of disaccharides being studied.

Analytes were purchased from: aSigma-Aldrich, Inc. (St. Louis, MO, USA) and bCarbosynth, Ltd. (Berkshire, UK).

Type Linkage

Disaccharides

1-2 α-D-Glcp-D-Glca β-D-Glcp-D-Glca

1-3α-D-Glcp-D-Glca β-D-Glcp-D-Glcb

α-D-Glcp-D-Frua α-D-Galp-D-Galb

α-D-Manp-D-Manb

Page 51: Structural Analysis of Carbohydrates by Mass Spectrometry

28

2.2.2 18O-labeling of Reducing Disaccharides

Carbonyl oxygen of the reducing sugar was 18O-labeled by dissolving 0.5 mg of a

disaccharide or oligosaccharide in 100 μL of H218O and heated up to 60 °C for 6 - 24 h.

The progress of the reaction was monitored by MS. Once the reaction completed; the

above solution was further diluted to 0.01 mg/mL (30 μM) with methanol. 1% (volume)

of triethylamine was added to the diluted solution right before MSn analysis.

2.2.3 Mass Spectrometry

All samples were analyzed in the negative-ion mode on a QTRAP 4000 mass

spectrometer (Applied Biosystems/Sciex, Toronto, Canada) equipped with a home-built

nanoelectrospray ionization (nanoESI) source. A schematic diagram of the instrument

ion optics is shown in the Scheme 2.2. Two types of low energy collisional activation

methods were accessible on this instrument, i.e. beam-type CID and ion trap CID. In

beam-type CID, the precursor ions (m/z 341 or 343) were isolated in Q1, accelerated in

the Q2 collision cell for collisional activation, and all products were analyzed in the Q3

linear ion trap. Collision energy (CE) was defined by the potential difference (absolute

value) between Q0 and Q2. In ion trap CID, the precursor ions were isolated in the Q3

linear ion trap via the RF/DC mode and a dipolar excitation was used for collisional

activation. In order to perform ion trap CID at different Mathieu q-parameters, an AC

(alternating current) generated from an external waveform generator (Agilent

Technologies, Santa Clara, CA, USA) was used for resonance excitation. Frequency and

the low mass cut-off were calculated by SxStability (Pan Galactic Scientific, Omemee,

Ontario, Canada). MS3 CID experiments were carried out by first performing beam-type

Page 52: Structural Analysis of Carbohydrates by Mass Spectrometry

29

CID of precursor ions in Q2. The fragment ions of interest were isolated in Q3 and then

subjected to ion trap CID. Analyst 1.5 software was used for instrument control, data

acquisition, and processing. The typical parameters of the mass spectrometer used in this

study were set as follows: spray voltage, -1.1 to -1.5 kV; curtain gas, 10; declustering

potential, 50 V; beam-type CID collision energy (CE), 5 to 30 V; ion trap CID activation

energy (AF2), 5 to 60 (arbitrary units); scan rate, 1000 m/z/s; pressure in Q2, 6.9x10-3

Torr, and in Q3, 3.6x10-5 Torr. Ion injection time was controlled to keep a similar parent

ion intensity: typically 3x106 counts per second (cps) for MS2 CID experiments and

1x106 cps for MS3 CID experiments. Activation time was kept constant at 200 ms for all

ion trap CID experiments. Seven spectra were collected for CID of m/z 221 ions from

synthesized monosaccharide-GA standards (deprotonated molecules) and disaccharides

over a one year period. Standard deviations of peak heights were calculated for major

fragments such as m/z 87, 99, 101, 113, 129, 131, 159, 161, 203, and 221, which were

observed from all the standards and disaccharides studied here except β-D-Glcp-GA and

β-D-Glcp-(1-2)-D-Glc, which showed no peaks at m/z 99.

Scheme 2.2 A schematic diagram of the QTRAP 4000, a hybrid triple-quadrupole/linear ion trap mass spectrometer.

Q1Q0 Q2 Q3nanoESI

Dipolar AC

Beam-type CID

Ion trap CID

5 mTorr 2.5x10-5 Torr

Page 53: Structural Analysis of Carbohydrates by Mass Spectrometry

30

2.3 Results and Discussion

2.3.1 CID of Deprotonated Disaccharide Ions from 1-3 Linked Disaccharides

Ion trap CID of deprotonated 1-3 linked disaccharides (m/z 341) typically

generates ions at m/z 221 in trace abundance on a Paul trap instrument, and isolation or

further CID of m/z 221 ions have not been achieved before.2,4-6 The 4000QTRAP mass

spectrometer used in this study has a unique triple quadrupole-linear ion trap

configuration, offering high sensitivity due to the large capacity of the linear ion trap, and

allowing either beam-type or ion trap collisional activation. In beam-type CID, the

precursor ions were isolated in Q1 and accelerated in Q2 for collisional activation, while

ion trap CID was conducted in Q3 with a dipolar excitation for collisional activation.

Since CID fragmentation patterns can be sensitive to the means of activation, the

formation of m/z 221 ions from five 1-3 linked disaccharides was investigated via both

beam-type and ion trap CID. Figure 2.1 compares the MS2 beam-type and ion trap CID

of deprotonated β-D-Glcp-(1-3)-D-Glc (m/z 341) using low energy CID conditions. A

relatively low CE (6 V) was used for beam-type CID; in ion trap CID, the AF2 for an AC

dipolar excitation was set to 25 (arbitrary units) for 200 ms. Under either activation

condition, the absolute intensities of m/z 221 ions (indicated by an arrow in Figure 2.1)

were very low and their relative intensities were less than 1% (normalized to the base

peak in the spectrum). This phenomenon was generally observed for all 1-3 linked

disaccharides studied herein. The insets in Figure 2.1 demonstrate the isolated m/z 221

ions (with a 2 m/z isolation window) from each set of dissociation conditions. For beam-

type CID, 1x106 cps of m/z 221 ions could be accumulated with an injection time of 1 s,

Page 54: Structural Analysis of Carbohydrates by Mass Spectrometry

31

which was sufficient for performing the next stage of tandem mass spectrometry (MS3 in

this case) with reasonable ion statistics and sensitivity. Far lower abundance of the m/z

221 ions (4.6x104 cps) could be isolated from ion trap CID of m/z 341, even after

doubling the injection time to 2 s. As a result, it was not feasible to obtain MS3 CID for

m/z 221 ions generated from m/z 341 precursor ions initially isolated within the trap. In

experiments described below, beam-type CID was used to dissociate disaccharide

precursor anions in the Q2 collision cell thereby generating m/z 221 product ions in high

enough abundance to acquire their spectra in the linear trap reproducibly.

Figure 2.1 MS2 CID spectra in the negative ion mode obtained from deprotonated β-D-Glcp-(1-3)-D-Glc (m/z 341) under low energy dissociation conditions: (a) beam-type CID (CE = 6 V), and (b) ion trap CID (AF2 = 25). Insets in (a) and (b) show the isolation of m/z 221 ions generated from beam-type CID (injection time = 1 s) and ion trap CID (injection time = 2 s), respectively.

100 140 180 220 260 300 340m/z

0

100

Rel

ativ

e In

tens

ity, %

341

161179

113143

100 140 180 220 260 300 340m/z

0

100179

161

341

113143

(a)

(b)

1.05e6221

222

Rel

ativ

e In

tens

ity,%

4.6e4 221

60

60

Inte

nsity

, cps

Inte

nsity

, cps

Page 55: Structural Analysis of Carbohydrates by Mass Spectrometry

32

2.3.2 CID of m/z 221 Ions Generated from 1-3 Linked Disaccharides

Previous studies have demonstrated that m/z 221 product ions formed from

collisional activation of disaccharide anions typically consist of an intact non-reducing

sugar with a 2-carbon aglycon derived from the reducing sugar.2 Three dominant

fragment peaks are commonly observed from CID of m/z 221: m/z 101, 131 and 161.

The relative intensities of these peaks together with some other fragment ions can be used

to establish the fragmentation patterns and to distinguish the stereochemistry and

anomeric configuration of the non-reducing sugar. Given that the CID patterns of m/z

221 ions will be used for structural identification, spectral reproducibility is an important

issue. Similar to the findings from a Paul trap instrument,1,2 we noticed that the number

of ions (m/z 221) in the linear ion trap and the energy input into an ion were among the

most important parameters affecting spectral reproducibility. To ensure reasonable ion

statistics and avoid adverse space charge effects, the intensity of the m/z 221 ions was

kept at 1x106 cps before MS3 CID. Based on previous studies, the CID energies were

tuned so that the ratio of remaining precursor ion to the most abundant product ion was

kept around 18 ± 3%.1 Figures 2.2, a, b, e, and f were the averaged spectra from seven

repetitions collected over a one year period and they were further used to make spectral

comparisons in later discussion. Error bars in the spectra indicate the standard deviation

of the peak intensity for 10 major fragment ions which were frequently observed for all

the disaccharides studied herein (m/z 87, 99, 101, 113, 129, 131, 159, 161, 203, and 221).

The standard deviations for these peaks were less than 5% in most cases, indicating high

reproducibility of the spectra from day to day by controlling the ion counts in the trap

before CID and the energy input to the ions. Since the CID patterns upon dissociation of

Page 56: Structural Analysis of Carbohydrates by Mass Spectrometry

33

m/z 221 ions can differ to some extent from instrument to instrument7, CID spectra of the

synthetic monosaccharide-GA were collected as standards for comparisons. Figures 2.2,

a and b show the CID data of - and β-D-Glcp-GA, respectively. The abundant peaks at

m/z 101 and 131 in Figure 2.2a are a signature of a non-reducing glucose with an

anomeric configuration. Note that a distinct fragmentation pattern is observed for the β

configuration (Figure 2.2b), where m/z 131 and 161 ions are dominant. The same

collisional activation conditions were applied to m/z 221 ions derived from α-D-Glcp-(1-

3)-D-Glc and β-D-Glcp-(1-3)-D-Glc, anomeric isomers containing a non-reducing

glucose. It is obvious that the spectra from the two anomeric isomers (Figures 2.2, c and

d) were drastically different from their corresponding D-Glcp-GA standards, however,

were similar to each other. This indicates that the m/z 221 ions generated using low

collision energies from m/z 341 precursors have different structures from the D-Glcp-GA

standards, and that their CID patterns cannot be used to assign either the stereochemistry

or anomeric configurations of the ions. Note that beam-type CID was used to generate

the m/z 221 ions from disaccharides shown in Figure 2.2, a condition differing from

previous studies where ion trap CID had been used.1 This difference in activation could

have contributed to the formation of structural isomers observed for the m/z 221 product

ions. In order to test this hypothesis, m/z 221 ions of 1-2 linked disaccharides, α-D-Glcp-

(1-2)-D-Glc and β-D-Glcp-(1-2)-D-Glc, were formed using similar beam-type CID

conditions and further subjected to MS3 CID (Figures 2.2, e and f). Except for a larger

fluctuation in peak intensity for m/z 203, almost identical fragmentation patterns to the

standards were observed (compare Figure 2.2, a to e and b to f), strongly indicating that

the expected D-Glcp-GA structures were formed. We further investigated a wide variety

Page 57: Structural Analysis of Carbohydrates by Mass Spectrometry

34

of disaccharides and found that the CID patterns of m/z 221 ions matched with their

corresponding monosaccharide-GA standards with the exception of 1-3 linked

disaccharides when low collision energy beam-type CID conditions were used to

dissociate the disaccharides.

Figure 2.2 MS2 ion trap CID spectra of m/z 221 ions derived from synthetic standards (a) α-D-Glcp-GA, AF2 = 25 and (b) β-D-Glcp-GA, AF2 = 18. MS3 CID spectra of m/z 221 ions derived from low energy beam-type CID of glucose-containing disaccharides: (c) α-D-Glcp-(1-3)-Glc, CE = 6 V for MS2 and AF2 = 25 for MS3, (d) β-D-Glcp-(1-3)-Glc, CE = 6 V for MS2 and AF2 = 25 for MS3, (e) α-D-Glcp-(1-2)-Glc, CE = 5 V for MS2 and AF2 = 27 for MS3, and (f) β-D-Glcp-(1-2)-Glc, CE = 5 V for MS2 and AF2 = 25 for MS3. The error bars in the spectra show the standard deviation of the peak intensity based on seven spectra collected over a 1 y period.

-H -α-D-Glcp-GA, m/z 221

(b)

-H -β-D-Glcp-GA, m/z 221

-

60 100 140 180 220m/z

161

203131113

221

-H

α-D-Glcp-(1-3)-Glc, m/z 341

(c)

β-D-Glcp-(1-3)-Glc, m/z 341-H -

m/z 221

-

-H -α-D-Glcp-(1-2)-Glc, m/z 341

-H

β-D-Glcp-(1-2)-Glc, m/z 341

m/z 221 m/z 221

m/z 221

0

100

Rel

ativ

e In

tens

ity, %

0 60 100 140 180 220m/z

100

Rel

ativ

e In

tens

ity, %

161

203131113

221

(d)

60 100 140 180 220m/z

0

100

Rel

ativ

e In

tens

ity, %

60 100 140 180 220m/z

0

100

Rel

ativ

e In

tens

ity, %

87

99101

113

129

131

159161203

221

(a)

87101 113

129

131

159

161

203221

60 100 140 180 220m/z

0

100

Rel

ativ

e In

tens

ity, %

87101113

129159

161

221

131

(f)

60 100 140 180 220m/z

0

100

Rel

ativ

e In

tens

ity, %

87

101

113

129

131

159161203

221

99

(e)

Page 58: Structural Analysis of Carbohydrates by Mass Spectrometry

35

18O-labeling at the carbonyl position of the reducing sugar was used to mass-

discriminate the “sidedness” of dissociation events to either side of the glycosidic linkage

and thus the origins of the m/z 221 and/or potential 223 product ions. Figure 2.3

compares relatively low energy beam-type and ion trap CID of 18O-labeled deprotonated

α-D-Glcp-(1-3)-D-Glc, m/z 343. Similar fragments were observed for both conditions;

however, the ion abundance for m/z 283 (loss of 60 Da, C2H4O2) was much higher in

beam-type CID relative to ion trap CID. It is possible that this fragmentation channel

requires higher activation energy and is promoted, even in lower-energy beam-type CID,

since higher collision energies (several eV) may have been obtained as compared to ion

trap CID (hundreds of meV). The inset in Figure 2.3a shows data collected using a wide

isolation window (6 m/z units) around m/z 221 after CID of m/z 343 precursor ions. A

peak at m/z 223, due to incorporation of 18O, appeared with much higher abundance than

m/z 221 ions for both low-energy beam-type and ion trap CID. Note that if the expected

D-Glcp-GA structure were formed, it should consist of the intact reducing sugar unit with

a 2-carbon aglycon derived from the reducing sugar (C-2 and C-3 or C-3 and C-4). In

this case, 18O should not be incorporated into the product ion and it should still appear at

m/z 221. Therefore, the observation of abundant m/z 223 ions indicated that under

relatively low energy CID conditions, most of the m/z 221 ions formed from α-D-Glcp-

(1-3)-D-Glc do not have the D-Glcp-GA structure which is the structural isomer needed

to distinguish the stereochemistry and anomeric configuration of the non-reducing sugar.

Ions at m/z 221 and m/z 223 were further subjected to ion trap CID. Figure 2.4

compares the CID spectra of the isolated m/z 221 ions and the m/z 223 ions generated

from 18O-labeled α-D-Glcp-(1-3)-D-Glc and β-D-Glcp-(1-3)-D-Glc. The CID spectra of

Page 59: Structural Analysis of Carbohydrates by Mass Spectrometry

36

the m/z 221 ions (Figures 2.4, a and c) from the two anomeric isomers are distinct from

each other and almost identical to those of the corresponding α and β-D-Glcp-GA

standards (Figures 2.4, a and b). The fragmentation patterns of m/z 223 (Figures 2.4, b

and d), however, were similar to each other yet were very different from the synthetic

glucosyl-GA standards.

Figure 2.3 MS2 spectra of 18O-labeled α-D-Glcp-(1-3)-D-Glc under (a) relatively low-energy beam-type CID (CE = 6 V), and (b) ion trap CID (AF2 = 25). Insets in (a) and (b) show isolation of m/z 221 and 223 ions generated from beam-type CID (injection time = 1 s) and ion trap CID (injection time = 2 s), respectively.

100 140 180 220 260 300 340m/z

0

100R

elat

ive

Inte

nsity

, %343

283163179

100 140 180 220 260 300 340m/z

0

100

Rel

ativ

e In

tens

ity, %

343

179163

(a)

(b)

0

2.9e5

Inte

nsity

, cps

223

221

0

4.0e4

Inte

nsity

, cps

223

221

60

60

Page 60: Structural Analysis of Carbohydrates by Mass Spectrometry

37

The major fragments that resulted from CID of the m/z 223 ions included product ions at

m/z 205, 163, 131, and 113. These ions are likely due to losses of water (-18 Da), a 2-

carbon piece, C2H4O2 (-60 Da), a 3-carbon piece including 18O, C3H6O2 + 18O (-92 Da),

and sequential or concerted losses of a 3-carbon piece plus water including 18O (-110

Da), respectively. Interestingly, neither the loss of water nor the loss of 60 Da

significantly involves loss of the 18O oxygen. The m/z 223 ions are hypothesized to have

a structure in which the reducing sugar is connected to a 2-carbon piece from the non-

reducing sugar as shown in the scheme above Figures 2.4b and 2.4d. Note that C-1 is no

longer chiral on the piece from the (former) non-reducing sugar, which also explains the

similarity in the CID data of the m/z 223 ion derived from the two anomeric isomers

(Figure 2.4, b and d). We also noticed some subtle differences between Figure 2.4b and

2.4d. For example, the relative intensities of m/z 205 and m/z 159 are higher (more than

10%) in Figure 2.4b than in 2.4d. These differences may be due to the existence of a

small fraction of structural isomers other than that hypothesized for the m/z 223 ions.

Page 61: Structural Analysis of Carbohydrates by Mass Spectrometry

38

Figure 2.4 MS3 spectra of m/z 221 and 223 ions derived from 18O-labeled α-D-Glcp-(1-3)-D-Glc and β-D-Glcp-(1-3)-D-Glc. (a) CID of m/z 221 ions, CE = 15 V (MS2), AF2 = 15 (MS3), (b) CID of m/z 223 ions, CE = 5 V (MS2), AF2 = 14 (MS3) from 18O-labeled α-D-Glcp-(1-3)-D-Glc, and (c) CID of m/z 221 ions, CE = 15V (MS2) , AF2 = 28 (MS3), (d) CID of m/z 223 ions, CE = 5 V (MS2), AF2 = 24 (MS3) from 18O-labeled β-D-Glcp-(1-3)-D-Glc.

2.3.3 The Effect of CID Conditions on the Formation of m/z 221 Diagnostic Ions

As demonstrated in Figure 2.4, abundant structural isomers of m/z 221 product

ions, (m/z 223 ions from the 18O-labeled disaccharides) were observed under relatively

low-energy dissociation conditions of 1-3 linked disaccharides, either using beam-type or

ion trap CID. This prevents the assignment of the stereochemistry or anomeric

configuration of the non-reducing sugar in a typical scenario where either a disaccharide

(d)

-H - m/z 223 -H -

β-D-Glcp-(1-3)-Glc, m/z 343m/z 221 -H

-H --

β-D-Glcp-(1-3)-Glc, m/z 343

(c)

60 100 140 180 220m/z

0

100

Rel

ativ

e In

tens

ity, %

131

10199

22111387159

203

(b)(a)

-H - m/z 223 -H -

α-D-Glcp-(1-3)-Glc, m/z 343 m/z 221 -H

-H -

-α-D-Glcp-(1-3)-Glc, m/z 343

60 100 140 180 220m/z

163

205

113 131

223

0

100

Rel

ativ

e In

tens

ity, %

0

100

Rel

ativ

e In

tens

ity, %

159161

60 100 140 180 220m/z0

100

Rel

ativ

e In

tens

ity, %

163

113 205131

223159

60 100 140 180 220m/z

131

161

22187

203159101

1818

1818

Page 62: Structural Analysis of Carbohydrates by Mass Spectrometry

39

is unlabeled or when it is isolated (unlabeled) from a larger oligosaccharide structure. It

would be highly desirable to optimize CID conditions or, for that matter, to find any

dissociation conditions whereby the relatively pure, structurally informative glycosyl-GA

(m/z 221) ions could be formed predominantly. Figure 2.5 shows the effect of collision

energies on the formation of m/z 221 and m/z 223 ions under beam-type and ion-trap

CID, using 18O-labeled β-D-Glcp-(1-3)-D-Glc as an example. The data were collected

using a wide isolation window around m/z 221 to observe both m/z 221 and m/z 223 ions.

It is clear from Figures 2.5a to c that the collision energy in beam-type CID

affects the absolute and relative intensities of m/z 221 ions. When the CE was relatively

low (CE = 5 V), m/z 223 ions were predominantly formed, with four times higher

intensity than that of m/z 221. At a higher CE (CE = 10 V), m/z 221 and m/z 223 ions

were seen at nearly equal intensities. Once the CE was increased to 15 V, m/z 221 ions

became the dominant peak, accounting for 80% of the total intensities from m/z 221 and

223. Further increasing CE, however, resulted in a huge loss of ion abundance possibly

due to competitive ion ejection thus the ratio was not improved.

Parameters that might affect the formation of m/z 221 ions versus m/z 223 ions

were also examined for ion trap CID. When ion trap CID of m/z 343 was performed

under the instrument default Mathieu q-parameter (q = 0.235), m/z 223 ions were formed

exclusively independent of activation energies (data not shown). By changing the

activation Mathieu q-parameter to a higher value, precursor ions are placed under a

higher potential well depth, and higher activation energies can be applied. An AC

generated from an external waveform generator was used for resonance excitation at q =

0.4. As shown in Figures 2.5d to f, the ratio of m/z 221 to m/z 223 ions was increased

Page 63: Structural Analysis of Carbohydrates by Mass Spectrometry

40

from almost zero to about 1 as the activation amplitude was increased from 100 mVpp to

400 mVpp (activation time: 50 ms for all cases). Further increasing the activation

amplitude resulted in a decrease in m/z 221 to 223 ratio as well as a huge ion loss. The

data in Figure 2.5 suggest that m/z 221 ions, which have the desired monosaccharide-GA

structures, are generated more favorably under relatively high collision energy conditions

in both beam-type and ion trap CID. As compared to ion trap CID, beam-type CID

provided more abundant and higher relative intensities of the m/z 221 ions that were

wanted for discrimination of the stereochemistry and anomeric configuration of the non-

reducing sugar. Evidently, a higher activation energy is needed for the formation of these

m/z 221 product ions, and the pathway to generate the glycosyl-GAs is favored when the

internal energies of the molecular ions increase. In beam-type CID, much higher

collision energies can be applied (typically more than 10 V) as compared to ion trap CID

(hundreds of mV), which leads to a shift in the internal energy distribution of the

molecular ions to the high energy direction.8 It is interesting to point out that the

glycosyl-glycolaldehyde product ions are virtually the only isomeric species generated

under ion trap or low-energy beam-type CID of the 1-2 linked disaccharide anions.2

Since much higher relative intensities of the glycosyl-glycolaldehyde product ions (10 –

40%, normalized to the most abundant peak) can be formed from 1-2 linkages as

compared to that of 1-3 linkages (typically < 1% relative intensity) under ion trap CID

conditions, it is reasonable to conclude that the formation of these ions from 1-2 linkages

needs less energy than required for their generation from 1-3 linkages. Therefore, the

formation of the glycosyl-glycolaldehyde product ions is a much lower energy

dissociation channel for 1-2 linked disaccharides but a fragmentation pathway for this

Page 64: Structural Analysis of Carbohydrates by Mass Spectrometry

41

isomeric species can only be promoted for 1-3 linked disaccharides when the collision

energy is higher.

Figure 2.5 Isolation of m/z 221 and m/z 223 ions derived from 18O-labeled β-D-Glcp-(1-3)-D-Glc under different collisional activation conditions. Beam-type CID: (a) CE = 5 V, (b) CE = 10 V, (c) CE = 15 V. Ion trap CID at q = 0.4, f = 119.248 kHz, excitation time = 50 ms: (d) 100 mVpp, (e) 250 mVpp, (f) 400 mVpp.

Given the high pressure in the collision cell (~ 5 mTorr), multiple collisions

happen in beam-type CID and the first-generation product ions may also be subjected to

collisional activation once they are formed within the collision cell especially under

higher CE conditions. In this sense, beam-type CID is less selective than ion trap CID,

where fragment ions are not typically further activated. Indeed, MS3 CID studies in the

ion trap showed that many fragment ions, including m/z 325, 323, 283, 281, 253, and 251

generated m/z 221 ions, which might contribute to the observation of higher intensity m/z

221 ions under beam-type CID due to secondary dissociation.

221 223m/z

3.9e5

Inte

nsity

, cps

223

221

221 223m/z

2.5e5221

223

221 223m/z

3.9e5 221

223

(a) (b) (c)

221 223m/z

9.1e4223

221 223m/z

2.0e5223

221

221 223m/z

1.0e5223

221(d) (e) (f)In

tens

ity, c

ps

Inte

nsity

, cps

Inte

nsity

, cps

Inte

nsity

, cps

Inte

nsity

, cps

0 0 0

0 0 0

Page 65: Structural Analysis of Carbohydrates by Mass Spectrometry

42

2.3.4 Stereochemistry Determination of the Non-Reducing Sugar within 1-3 Linked Disaccharides

Since relatively pure m/z 221 ions containing the intact non-reducing sugars could

be formed using beam-type CID with high collision energies, it was possible to

differentiate the stereochemistry and anomeric configuration of the non-reducing sugar in

disaccharides without 18O-labeling. Figure 2.6 shows the MS3 CID spectra of m/z 221

ions generated by beam-type CID with relatively high CE (13 V to 22 V) from five 1-3

linked disaccharides. Each spectrum was an average of seven spectra and the error bars

indicate standard deviations of the peak intensities. The standard deviations were found

to be higher (0 - 12 %) for the disaccharide samples than those from standards (0 - 4 %).

This larger degree of spectral variation is likely contributed by the fluctuation in ion

intensity of the low abundance m/z 221 isomers under slightly different instrument

conditions, and these isomers fragment differently from the diagnostic and more

abundant m/z 221 ions that have the monosaccharide-GA structures. Note that α-D-Glcp-

(1-3)-D-Glc, α-D-Glcp-(1-3)-D-Fru, and β-D-Glcp-(1-3)-Glc are disaccharide isomers

containing a glucose as the non-reducing sugar; however, each has either a different

anomeric configuration or reducing sugar. The characteristic fragmentation profile for

disaccharides having glucose as the non-reducing sugar and an α-anomeric configuration

can be clearly identified for Figure 2.6a (α-D-Glcp-(1-3)-D-Glc) and Figure 2.6c (α-D-

Glcp-(1-3)-D-Fru), which is distinct from the β-anomeric isomer as shown in Figure 2.6b

(β-D-Glcp-(1-3)-D-Glc, compare all three to the synthetic standards shown in Figure 2.2a

and 2.2b). Figure 2.6d shows the characteristic fragmentation profile for the m/z 221 ion

Page 66: Structural Analysis of Carbohydrates by Mass Spectrometry

43

of disaccharides having galactose as the non-reducing sugar and having an α-anomeric

configuration (compare to Figure 2.8a, CID of m/z 221 from -D-Galp-GA standard).

Figure 2.6 MS3 CID of m/z 221 ions generated via using high CE (CE = 13 to 22 V) for the dissociation of deprotonated disaccharide ions (m/z 341). (a) α-D-Glcp-(1-3)-D-Glc, CE = 15 V (MS2), AF2 = 26 (MS3), (b) β-D-Glcp-(1-3)-D-Glc, CE = 13 V (MS2), AF2 = 30 (MS3), (c) α-D-Glcp-(1-3)-D-Fru, CE = 18 V (MS2), AF2 = 25 (MS3), (d) α-D-Galp-(1-3)-D-Gal, CE = 22 V (MS2), AF2 = 35 (MS3), and (e) α-D-Manp-(1-3)-D-Man, CE = 20 V (MS2), AF2 = 36 (MS3). The error bars in the spectra show the standard deviation of peak intensities based on seven spectra collected over a 1 y period. The MS3 CID of α-D-Manp-(1-3)-Man (Figure 2.6e) was similar to that of the -D-

Manp-GA (Figure 2.8b). It is also important to note that under low-energy dissociation

conditions, the spectra of the m/z 221 product ions derived from the disaccharides α-D-

(b) (c)

-H -α-D-Glcp-(1-3)-Glc, m/z 341

m/z 221

β-D-Glcp-(1-3)-Glc, m/z 341-H - -H -

α-D-Glcp-(1-3)-Fru, m/z 341

-H -α-D-Galp-(1-3)-Gal, m/z 341-H -α-D-Manp-(1-3)-Man, m/z 341

m/z 221 m/z 221

m/z 221 m/z 221

(d)

60 100 140 180 220m/z

0

100

Rel

ativ

e In

tens

ity, %

87101113

129

131

159

161

203

221

60 100 140 180 220m/z

0

100

Rel

ativ

e In

tens

ity, %

87

99101

113

129

131

159161

203

221

60 100 140 180 220m/z

0

100

Rel

ativ

e In

tens

ity, %

8799

101

113129131

159

161

203221

60 100 140 180 220m/z

0

100

Rel

ativ

e In

tens

ity, %

8799101

113

129

131

159

161

203 221

(e)

60 100 140 180 220m/z

0

100

Rel

ativ

e In

tens

ity, %

8799

101

113

131

129 159

161

203221

(a)

Page 67: Structural Analysis of Carbohydrates by Mass Spectrometry

44

Glcp-(1-3)-D-Fru, α-D-Galp-(1-3)-D-Gal and α-D-Manp-(1-3)-Man did not match those

of their respective glycosyl-glycolaldehydes (Figures 2.7 and 2.8c and 2.8d). We

conclude this is due to the presence of alternate isomers, possibly related in their origins

to the hypothetical structures shown in Figures 2.4b and 2.4d but having different

reducing monosaccharides.

Figure 2.7 CID spectra of m/z 221 ions derived from α-D-Glcp-(1-3)-D-Fru, CE = 5 V (MS2), AF2 = 30 (MS3).

60 100 140 180 220m/z

0

100

Rel

ativ

eIn

tens

ity, %

161

131

113

22120387

-H

α-D-Glcp-(1-3)-Fru, m/z 341

m/z 221

-

Page 68: Structural Analysis of Carbohydrates by Mass Spectrometry

45

Figure 2.8 CID spectra of m/z 221 ions derived from (a) α-D-Galp-GA, AF2 = 17 (MS2), (b) α-D-Manp-GA, AF2 = 13 (MS2), (c) α-D-Gal-(1-3)-D-Gal, CE = 5 V (MS2), AF2 = 20 (MS3), (d) α-D-Man-(1-3)-D-Man, CE = 5 V (MS2), AF2 = 30 (MS3). The error bars in the spectra (a) and (b) show the standard deviation of the relative intensity calculated based on seven spectra collected over a 1 y period.

2.3.5 Spectral Matching by Similarity Scores

The methodology for assigning the stereochemistry and anomeric configuration

for the non-reducing sugar unit within a disaccharide is based on the comparison of the

CID patterns of m/z 221 ions to those of the synthetic monosaccharide-GA standards 1.

A high similarity between the compared spectra indicates a large likelihood of them

sharing the same structure. Spectral similarity scores, which have been widely used in

-H -α-D-Manp-GA, m/z 221-H -α-D-Galp-GA, m/z 221

-H - -H-

m/z 221

α-D-Galp-(1-3)-Gal, m/z 341 α-D-Manp-(1-3)-Man, m/z 341

m/z 221

(a)

100

60 100 140 180 220m/z

113131

161203

221101

(c)

0

Rel

ativ

eIn

tens

ity, %

60 100 140 180 220m/z

189161

131

203113

221

101

(d)

0

100

Rel

ativ

eIn

tens

ity, %

163

60 100 140 180 220m/z

0

100

Rel

ativ

e In

tens

ity, %

(b)

60 100 140 180 220m/z

0

100

Rel

ativ

e In

tens

ity, %

87 99 113129

131

159

161

203

221

8799

101

113131129

159

161

203 221

-

Page 69: Structural Analysis of Carbohydrates by Mass Spectrometry

46

mass spectral library search for both small molecules,9 peptides10-12 and

oligosaccharides13 were chosen to facilitate these comparisons. The spectral similarity

scores were calculated between each of the averaged spectra in Figure 6 and the

averaged spectra from the monosaccharide-GA standards based on the following

equation,

Spectral similarity score , and k (Eq. 2.1)

where Im1 and Im

2 are the normalized intensities of an ion at m/z = m for the two spectra.

Note that the spectral similarity score always has a value between 0 and 1. If two spectra

are exactly the same, then, the spectral similarity score becomes 1. In general, a large

similarity score indicates close similarity between the two spectra and a large degree of

structural similarity. As shown in Table 2.2, the spectral similarity scores between a

standard and a disaccharide having the same stereochemistry and anomeric configuration

for the non-reducing side were the highest scores, ranging between 0.9838 and 0.9977.

When a disaccharide’s stereochemistry and anomeric configuration on the non-reducing

side did not match with the standard, the spectral similarity score was significantly lower,

between 0.6942 and 0.9178. Clearly, by comparing the spectral similarity scores,

assigning the stereochemistry and anomeric configuration for the non-reducing side of

the 1-3 linked disaccharides could be achieved with high confidence. Note that this was

only possible under high-energy beam-type CID conditions where the m/z 221 product

anions containing the intact non-reducing sugars were optimally generated from

precursor disaccharides.

Page 70: Structural Analysis of Carbohydrates by Mass Spectrometry

47

Table 2.2 Spectral similarity scores for 1-3 linked disaccharides vs. monosaccharide-GA standards.

2.4 Conclusions

Collisional activation of deprotonated 1-3 linked hexose-containing disaccharides

(m/z 341) generated a low-abundance m/z 221 product ion. By 18O-labeling the reducing

sugar carbonyl oxygen of these disaccharides, at least two structural isomers of the m/z

221 ion with the main portion derived from either side of the glycosidic linkage could be

mass-discriminated (m/z 221 vs. m/z 223), which enabled the isomers to be isolated and

independently studied. The m/z 221 isomer containing the intact non-reducing sugar

attached in glycosidic linkage to a glycolaldehyde aglycon was found to be analytically

useful, since CID of this species provided the structural information that identified the

stereochemistry and anomeric configuration of the non-reducing sugar. No structural

information could be obtained from m/z 223 isomer(s) to determine the stereochemistry

of the non-reducing sugar or its anomeric configuration. The formation of the diagnostic

m/z 221 isomer was found to be affected by CID conditions and was favored under higher

energy beam-type CID. It was demonstrated that under optimized CID conditions, this

DisaccharidesSynthesized Standards

α-D-Glcp-GA β-D-Glcp-GA α-D-Galp-GA α-D-Manp-GAα-D-Glcp-(1-3)-Glc 0.9977 0.8845 0.8823 0.7968β-D-Glcp-(1-3)-Glc 0.8572 0.9840 0.7608 0.8568α-D-Glcp-(1-3)-Fru 0.9930 0.8899 0.9027 0.8045α-D-Galp-(1-3)-Gal 0.9178 0.7530 0.9838 0.6942α-D-Manp-(1-3)-Man 0.8461 0.8702 0.7527 0.9891

Page 71: Structural Analysis of Carbohydrates by Mass Spectrometry

48

structural isomer could be generated predominantly from five different 1-3 linked

disaccharides without requiring 18O-labeling of the reducing sugar. Identification of the

nonreducing sugar and the anomeric configuration therefore were achieved at a high

confidence level by statistically comparing the CID data of m/z 221 ions generated from

the disaccharide samples to those of the synthetic standards via spectral similarity scores.

This study also demonstrated that beam-type CID was a more desirable activation method

as compared to ion trap CID to characterize disaccharides using the methodology based

on the CID patterns of m/z 221 ions. This method now enables the anomeric

configuration and stereochemistry of the m/z 221 ions derived from 2-, 3-, 4-, or 6-linked

disaccharides to be assigned in the negative ion mode. This capability was afforded due

to the specific arrangement of the triple quadrupole-linear ion trap instrument. It

combined (1) selection of the precursor (m/z 341) in the first quadrupole with (2) higher

energy dissociation in the second quadrupole collision cell followed by (3) buildup of the

desired low abundance m/z 221 product ion in the linear trap, all three of which were

necessary to obtain these structural details for 3-linked disaccharides.

Page 72: Structural Analysis of Carbohydrates by Mass Spectrometry

49

2.5 References

(1) Fang, T. T.; Bendiak, B., J. Am. Chem. Soc. 2007, 129, 9721-9736.

(2) Fang, T. T.; Zirrolli, J.; Bendiak, B., Carbohydr. Res. 2007, 342, 217-235.

(3) Brown, D. J.; Stefan, S. E.; Berden, G.; Steill, J. D.; Oomens, J.; Eyler, J. R.; Bendiak, B., Carbohydr. Res. 2011, 346, 2469-2481.

(4) Dallinga, J. W.; Heerma, W., Biol. Mass Spectrom. 1991, 20, 215-231.

(5) Carroll, J. A.; Ngoka, L.; Beggs, C. G.; Lebrilla, C. B., Anal. Chem. 1993, 65, 1582-1587.

(6) Garozzo, D.; Giuffrida, M.; Impallomeni, G.; Ballistreri, A.; Montaudo, G., Anal.Chem. 1990, 62, 279-286.

(7) Bendiak, B.; Fang, T. T., Carbohydr. Res. 2010, 345, 2390-2400.

(8) Wells, J. M.; McLuckey, S. A., Collision-Induced Dissociation (CID) of Peptides and Proteins. In Biol. Mass Spectrom., Volume 402 ed.; Burlingame, A. L., Ed. Academic Press: 2005; pp 148-185.

(9) Stein, S.; Scott, D., J. Am. Soc. Mass Spectrom. 1994, 5, 859-866.

(10) Frewen, B. E.; Merrihew, G. E.; Wu, C. C.; Noble, W. S.; MacCoss, M. J., Anal. Chem 2006, 78, 5678-5684.

(11) Lam, H.; Deutsch, E. W.; Eddes, J. S.; Eng, J. K.; King, N.; Stein, S. E.; Aebersold, R., Proteomics 2007, 7, 655-667.

(12) Lam, H.; Deutsch, E. W.; Eddes, J. S.; Eng, J. K.; Stein, S. E.; Aebersold, R., Nat. Methods 2008, 5, 873-875.

(13) Kameyama, A.; Kikuchi, N.; Nakaya, S.; Ito, H.; Sato, T.; Shikanai, T.; Takahashi, Y.; Takahashi, K.; Narimatsu, H., Anal. Chem. 2005, 77, 4719-4725.

(14) Zhang, Z., Anal. Chem. 2004, 76, 3908-3922.

Page 73: Structural Analysis of Carbohydrates by Mass Spectrometry

50

CHAPTER 3 OBTAINING STEREO-STRUCTURAL INFORMATION WITH SINGLE-SUGAR RESOLUTION FROM LINEAR OLIGOSACCHARIDES

3.1 Introduction

MS3 CID of the m/z 221 diagnostic ions was demonstrated to assign the

stereochemistry and anomeric configuration of the non-reducing sugar unit of D-

aldohexose (D-aldoHex) containing disaccharides for 1-2, -4, and -6 linked

disaccharides1 and 1-3 linked disaccharides2 in Chapter 2. This method was based on the

unique fragmentation patterns of diagnostic product ions formed by MS2 CID of

deprotonated disaccharide ions, having a non-reducing sugar glycosidically linked to a

glycolaldehyde (D-aldoHex-GA, m/z 221, C8H13O7; and D-aldoHexNAc-GA, m/z 262,

C10H17NO7 from acetylamine modified disaccharides). By comparing the MS3 CID

spectrum of thus obtained diagnostic ions from disaccharides to MS2 CID spectrum of the

synthetic standards using spectral matching algorithm, highly confident structural

assignment was achieved.2 Note that the diagnostic ion contains a substructure of a

disaccharide with significantly reduced number of stereo-centers, and thus the number of

standards for the diagnostic ion could be largely reduced (e.g. 16 for D-aldoHex-GA).

This is a significant reduction as compared to the fact that there are theoretically more

than 104 possible structural isomers for disaccharides (only considering D-hexoses).3

Twenty different sugar-GA standards have been synthesized so far (8 D-hexoses, D-

Page 74: Structural Analysis of Carbohydrates by Mass Spectrometry

51

GlcNAc, and D-GalNAc, each with 2 anomeric configurations).1 These standards form

abundant diagnostic ions as the deprontated molecular ions ([M-H]-1) upon ESI in the

negative ion mode. MS2 CID of the synthetic standards provides a straightforward way

of generating standard spectra library on a given instrument. This approach has been

successfully applied for the determination of non-reducing sugars within 13 disaccharides

with a combination of 5 sugar units (D-Glc, D-Gal, D-Man, D-GlcpNAc, and D-

GalpNAc), 2 anomeric configurations, and 4 different linkage positions (1-2, -3, -4, and -

6).1,2

Based on the establishment of m/z 221 diagnostic ions for non-reducing sugar

identity and anomeric configuration from disaccharides, we attempted to develop an MSn

approach which can pinpoint individual sugar identity and anomeric configuration from

an oligosaccharide. The key characteristics of this MSn (n = 4, 5) approach involves

optimizing cleavages at glycosidic linkages during early stages of MSn to form

overlapping disaccharide substructures, then enhancing cross-ring cleavages for the

formation of the m/z 221 diagnostic ions, and finally obtaining the fragmentation pattern

of m/z 221 ions which can lead to the assignment of sugar unit identities and anomeric

configurations. A modified hybrid triple quadrupole/linear ion trap mass spectrometer

was employed in this study to enable high efficiency bidirectional ion transfer between

quadrupole arrays4 so that two types of CID, i.e., beam-type or on-resonance ion trap CID

can be executed at any stage of MSn up to 5 stages (MS5). This capability was shown to

be critical for optimizing the formation of the ladder of product ions involved in the MSn

analysis. This MSn approach using highly modified instrument successfully lead us to

garner detailed stereo-structural information of individual sugar units, i.e. sugar identity

Page 75: Structural Analysis of Carbohydrates by Mass Spectrometry

52

and anomeric configuration, as well as their locations within two pentasaccharide

isomers: [α-D-Glcp-(1-4)]4-D-Glc and [β-D-Glcp-(1-4)]4-D-Glc.

3.2 Experimental

3.2.1 Materials

[α-D-Glcp-(1-4)]4-D-Glc, [β-D-Glcp-(1-4)]4-D-Glc, 4-aminobenzoic acid (ABA),

chloroform, sodium cyanoborohydride, acetic acid, and dimethyl sulfoxide (DMSO) were

purchased from Sigma-Aldrich, Inc. (St. Louis, MO, USA). Sugar-GA standards were

synthesized by Fischer glycosidation to generate allyl glycosides. They were converted to

the sugar-GAs by ozonolysis.1

3.2.2 Reducing End Modification (Reductive Amination)

Reducing end modification of oligosaccharides with ABA was conducted via

reductive amination.5 Briefly, oligosaccharides (0.5 mg) were dissolved in 10 μL of a

solution containing 0.15 M ABA in acetic acid: water (3:17, v/v), and 10 μL of 1.0 M

sodium cyanoborohydride dissolved in DMSO was added to achieve reductive amination.

After the vial was heated at 37 °C for 16 h, the solvent was removed under vacuum

(Speed-Vac, LABCONCO, Kansas City, MO) and the residue was dissolved in 300 μL

water followed by an addition of 300 μL chloroform. The aqueous phase containing the

derivatized oligosaccharide was collected after vortex mixing. This step was repeated

with another 300 μL of water for extraction and the aqueous phases were combined and

Page 76: Structural Analysis of Carbohydrates by Mass Spectrometry

53

vacuum dried. The dried sample was dissolved in 100 μL DMSO and 900 μL methanol.

This solution was further diluted 100-fold in methanol prior to MS analysis.

3.2.3 Mass Spectrometry

A modified hybrid triple quadrupole/linear ion trap mass spectrometer (QTRAP

4000, AB SCIEX, Concord, Ontario, Canada) was used for performing experiments with

high efficiency bidirectional ion transfer between Q1, Q2, and Q3 quadrupole arrays.4

Ion selection in Q1 utilized RF/DC isolation, while ion isolation in Q2 and Q3 was

typically performed by broadband waveform at q = 0.45 for the center of the frequency

notch. Isolation of the m/z 221 diagnostic ion was performed with unit resolution before

the last stage of CID, while the isolation windows for other precursors were varied from

unit to 3 Th to achieve optimized sensitivity. Ion trap CID was performed in Q2 or Q3

through on-resonance excitation by applying a dipolar ac signal. Beam-type CID was

performed with energetic ion transfer either from the Q1 Q2 or from Q3 Q2.

Derivatized samples were diluted to 0.01 mg/mL in methanol with 1% ammonium

hydroxide for negative mode ESI, which was conducted at flow rates of 1-2 μL/min.

3.2.4 MS2 CID Database of 16 D-AldoHex-GA Synthetic Standards

MS2 ion trap CID of 16 deprotonated D-aldoHex-GA standards (m/z 221) was

collected in the Q3 linear ion trap on a QTRAP 4000 instrument. To avoid space charge

effects, the parent ion intensity before CID was kept around 1 x 106 counts per second

(cps). The remaining parent ion intensity after CID was kept around 18 ± 5% relative to

Page 77: Structural Analysis of Carbohydrates by Mass Spectrometry

54

its base product ion (100%). Seven spectra from individual standards were collected over

a one year period. Standard deviations of peak heights were calculated for 21 major peaks

listed below, which were commonly observed from the standards. The averaged data

(from seven spectra) for the D-aldoHex-GA standards are shown in Figure 3.1, with the

standard deviations for each peak indicated. It is worth noting that depending on the

instrument and type of dissociation6 the specific product ion abundances for a single

compound can vary, but are reproducible within measured error bars for a specific

instrument, when energy input is defined by controlling the ratio of the precursor/base

product ion.

Page 78: Structural Analysis of Carbohydrates by Mass Spectrometry

55

Figure 3.1 Averaged MS2 trap CID data of 14 deprotonated D-aldoHex-GA standards (m/z 221). The 21 major peaks were labelled with standard deviation based on 7 averaged spectra obtained over a 1 y period.

β-D-Galp-GA, m/z 221α-D-Galp-GA, m/z 221

β-D-Manp-GA, m/z 221α-D-Manp-GA, m/z 221

β-D-Altp-GA, m/z 221α-D-Altp-GA, m/z 221

8799

101

113 131129 159161

203 221

60 100 140 180 220m/z0

100

%87

131

129 203221

60 100 140 180 220m/z0

100

%

16122187

159

20310160 100 140 180 220m/z

0

100

%131

129 87 99

101

113129131 161 203 221

60 100 140 180 220m/z0

100

%

87101 129

131

161203 221

15960 100 140 180 220m/z

0

100

%

101

99131 159 221113 16185

60 100 140 180 220m/z0

100

%129

97

8799101

113129

131

159 161 203 221 87101113129

131

159

161

221

60 100 140 180 220m/z0

100

%

60 100 140 180 220m/z0

100

%

β-D-Glcp-GA, m/z 221α-D-Glcp-GA, m/z 221

α-D-Allp-GA, m/z 221

α-D-Talp-GA, m/z 221

β-D-Idop-GA, m/z 221α-D-Idop-GA, m/z 221

β-D-Allp-GA, m/z 221

β-D-Talp-GA, m/z 221

131

16122115911387 20397 141

60 100 140 180 220m/z0

100

%101

131203101

11387

161 221

60 100 140 180 220m/z0

100

%189

9785 129 141

101

113

2218985 20399 16112513160 100 140 180 220m/z

0

100

%

97131

201113

141129101161 221

203125

87 189159

60 100 140 180 220m/z0

100

%85

159

161

22112999 131113 20360 100 140 180 220m/z

0

100

%85

189

131113 22110187 20316160 100 140 180 220m/z

0

100

%85

β-D-Gulp-GA, m/z 221α-D-Gulp-GA, m/z 221161101

113203 221

143131 189

97177

60 100 140 180 220m/z0

100

%8985

131

159 221203161129 189

60 100 140 180 220m/z0

100

%8797101

Page 79: Structural Analysis of Carbohydrates by Mass Spectrometry

56

3.2.5 Spectral Matching by Similarity Score

Spectral similarity scores were calculated between the CID spectrum of m/z 221

ions obtained from MSn (n =4 or 5) of oligosaccharides and each of the averaged spectra

in Figure 3.1 (CID spectra of m/z 221 ions of the D-aldoHex-GA standards). The

following equation was used for similarity score:7

Spectral similarity score , and k

where Im and Im* are the normalized intensities (to the highest fragment peak) of an ion at

m/z = m for the two spectra. 21 major peaks (m/z 85, 87, 89, 97, 99, 101, 111, 113, 125,

129, 131, 141, 143, 159, 161, 175, 177, 189, 201, 203, and 221) were chosen for this

calculation. Note that the spectral similarity score always has a value between 0 and 1. If

two spectra are exactly the same, then the spectral similarity score is equal to 1. In

general, a large similarity score indicates close similarity between the two spectra and a

large degree of structural similarity.

Page 80: Structural Analysis of Carbohydrates by Mass Spectrometry

57

3.3 Results and Discussion

3.3.1 MSn Approach

The general strategy of the MSn approach is shown in Scheme 3.1. It involves

optimizing cleavages at glycosidic linkages during early stages of MSn, then enhancing

cross-ring cleavages only with disaccharides isolated as an ordered set of overlapping

substructures (illustrated with a tetrasaccharide in Scheme 3.1, where sugar unit number

was assigned from reducing end (i.e., unit 1 4). Key steps include: 1) dissociation of a

precursor derivatized at the reducing end (M, Scheme 3.1) to a ladder of smaller

oligosaccharides, each successively one sugar unit shorter, containing the reducing end

tag (Ym ions);8 2) generation of disaccharide fragments (C2 ions) from each Ym ion,

derived from the opposite end; 3) dissociation of the disaccharides to their m/z 221

diagnostic product ions (the sugar-GA); 4) fragmentation of these diagnostic ions; 5)

spectral matching of the diagnostic ion CID data to that of a database of synthetic

standards for assigning stereochemistry.

Page 81: Structural Analysis of Carbohydrates by Mass Spectrometry

58

Scheme 3.1 The MSn (n = 4 or 5) approach for stereo-structural analysis of oligosaccharides. “M” stands for reducing end modification. Sugar unit number is assigned from the reducing end.

The above described MSn approach has the advantage of knowing the exact origin

of each fragment with respect to its initial position within the oligomer. Also, the

dissociation patterns of disaccharides in negative-ion mode provide information that can

assign the linkage between them.9 Given the need to preserve the structural integrity of

intact sugars and their linkage, disaccharides are key substructures to isolate during

earlier stages of MSn. The reducing end modification (M) introduces a mass distinction

between Y and C ions and also bears a negative charge that enables selection of the

ladder of singly charged Ym ions after MS2. Yet, through charge-transfer, it permits

Page 82: Structural Analysis of Carbohydrates by Mass Spectrometry

59

disaccharide (C2) ions to be isolated via a neutral loss of the derivatized end of the

molecule (MS3). Note that this approach does not define the structure of the reducing end

unit.

The success of this method highly depends on producing reasonable yields of the

ladder of ions employed in MSn, i.e., Ym, C2, and m/z 221 ions, from a given

oligosaccharide. Formation of Ym and C2 ions requires glycosidic bond cleavages. These

bonds can be cleaved preferentially using low-energy CID under most conditions.10 Our

previous studies showed that the higher-energy beam-type CID condition was critical to

produce m/z 221 diagnostic ions with both high structural purity and reasonable yield for

many 3-linked disaccharides.2 On the other hand, ion trap CID of the m/z 221 ions was

necessary to obtain reproducibly distinct fingerprint patterns for structural identification.

In order to have the capacity of optimizing the product ions involved in the tree of MSn

analysis, it is desirable to perform experiments on an instrument platform which provides

both beam-type and ion trap CID at any MS/MS stage.

3.3.2 A Hybrid Triple Quadrupole/Linear Ion Trap for MSn

We choose to work on a hybrid triple quardrupole/linear ion trap instrument since

this type of instrument already has the capability of both beam-type and ion trap CID and

therefore, efforts of instrument modification can be significantly reduced based on this

platform. A schematic representation of the instrument is shown in Figure 3.2.

Conventionally, beam-type CID is performed by accelerating ions from Q0 to Q2 (Figure

3.2a), while the precursor ions are mass isolated in Q1 (RF/DC) during fly. In this mode,

beam-type CID can only be performed in the MS2 stage. However, if the flow of ions

Page 83: Structural Analysis of Carbohydrates by Mass Spectrometry

60

can be reversed, the mass analysis quadrupole arrays (Q1 and Q3) and reaction sections

(Q0 and Q2) can be re-accessed and the tandem-in-space experiments such as beam-type

CID can be achieved as shown in Figure 3.2b and 3.2c. Indeed, Thompson et al.11 and

Xia et al.4 have shown that with the use of high efficiency bidirectional ion transfer

between quadrupole arrays, beam-type CID) can be performed in MSn (n>2) without

significant hardware modification of a Q-q-TOF instrument. Given the above

considerations, we employed a research grade triple quadrupole/linear ion trap mass

spectrometer to implement the proposed MSn approach. In order to facilitate high

efficiency bi-directional ion transfer, an axial electric field (LINAC)12,13 was superposed

to the center axis of Q2 and Q3 quadrupole arrays. The direction and amplitude of the

electric field were controlled from instrument software and were optimized during

experiments. Results collected from this instrument showed ~80% transfer efficiency

after three cycles of transfers (Q2 Q3, Q3 Q2, and Q2 Q3). Compared to beam-type

CID, it is relatively simple to perform ion trap CID at a given stage of MS/MS. The only

requirement is to transfer ions to the quadrupole array which is capable of ion trap CID.

On this modified instrument, on-resonance ion trap CID can be either performed in Q2 or

Q3 quadrupole arrays via dipolar ac excitation. Ion trap CID in Q2 improved CID

efficiency due to the higher pressure in Q2 (6 to 8 mTorr) as compared to that in Q3 and

Q1 (3 x 10-5 Torr), as previously described.14,15 With all above capabilities (i.e.,

reversing ion beams and ion trap CID in Q2 and Q3), MSn (n = 4 or 5) employing either

of the two CID methods is possible to conduct at any stage of MS/MS.

Page 84: Structural Analysis of Carbohydrates by Mass Spectrometry

61

Figure 3.2 A schematic of a modified triple quadrupole/linear ion trap mass spectrometer (QTRAP 4000). Beam-type CID can be performed through energetic axial ion transfer between quadrupole arrays (Q0 to Q3): a) conventional method: Q1 to Q2 with mass isolation in Q1; b) reversed ion transfer from Q3 to Q2 with mass isolation in Q3; and c) reversed ion transfer from Q1 to Q0 with mass isolation in Q1.

3.3.3 Effect of CID Conditions on the Formation of Key Product Ions

A pentasaccharide modified with 4-aminobenzoic acid ([α-D-Glcp-(1-4)]4-D-Glc-

ABA, structure shown in the inset of Figure 3.3) was used as a model compound for

method and instrumental parameter optimization. Based on the MSn method described in

Scheme 3.1, an MS5 CID with following sequence is necessary to obtain sugar unit 3

structural information: [474 ([M-2H]2-) 624 (Ym) 341 (C2) 221 (diagnostic

ion) fragments]. Figure 3.3a shows the ion transfer scheme for conducting MS3 CID

using the modified triple quadrupole/linear ion trap. In short, the doubly charged parent

anions ([M-2H]2-, m/z 474) formed by ESI in negative ion mode, were mass selected in

the Q1 quadrupole array (RF/DC mode) and accelerated to the Q2 collision cell to effect

beam-type CID. Product ions were collected in Q3, followed by a broadband isolation of

Y3 (m/z 624) ions. Either beam-type or ion trap CID of m/z 624 ions could be performed

Q1Q0 Q2 Q3-’ve ESI

Auxiliary AC

3e-5 Torr 6-8 mTorr 3e-5 Torr10 mTorr

Beam CIDa)

Beam CID b)

Beam CIDc)

SK IQ1 IQ2 IQ3 EXST1 ST3

Page 85: Structural Analysis of Carbohydrates by Mass Spectrometry

62

for MS3 by reversing the ion beam. Beam-type CID involved accelerating ions from Q3

to Q2 in the opposite direction of the traditional beam-type CID. For ion trap CID, m/z

624 ions were transferred from Q3 to Q2 with minimum kinetic energies, followed by on-

resonance dipolar ac excitation in Q2.

Note that the goal of the MS3 is to maximize the formation of C2 (m/z 341) from

Y3 (m/z 624), which requires a glycosidic bond cleavage between the first and the second

sugar units (refer to Figure 3.3 for the oligosaccharide nomenclature). A range of

collision energies (CEs) (10 – 60 V) were tested using beam-type CID from Q3 and Q2,

but m/z 341 ions were barely detectable under those conditions. At higher CEs, small

fragments in the range of m/z 200-300 dominated due to sequential fragmentation of Y3

ions (Figure 3.3b, CE = 42 V). The ion trap CID spectrum of Y3 ions (Figure 3.3c, q =

0.45, 1.4 V, and 300 ms activation) consisted of fewer fragments, most of which derived

from glycosidic bond cleavages including C2 ions (m/z 341) as well as m/z 462, 444, 300,

282. Ion trap CID consistently yielded more C2 ions from the set of singly charged Ym

ions as compared to beam-type CID. This observation is consistent with glycosidic bonds

being preferentially cleaved under slow heating conditions.16

Figure 3.3d shows the ion transfer scheme for conducting either beam-type CID

or ion trap CID in the 4th stage of MS/MS. The goal of MS4 is to maximize the formation

of m/z 221 diagnostic ions from C2 (m/z 341) ions. The MS1 and MS2 steps was the same

as Figure 3.3a and ion trap CID was employed in Q2 for the MS3 step. After that, the

MS3 product ions were sent to Q3 and C2 (m/z 341) ions were isolated in Q3 via

broadband. Using reversed ion transfer (Q3 Q2), C2 ions were subjected to either

beam-type or ion trap CID in Q2 for MS4. Our previous studies on disaccharides showed

Page 86: Structural Analysis of Carbohydrates by Mass Spectrometry

63

that diagnostic product ions (m/z 221) were formed by cross-ring cleavages and usually

required higher activation energies, especially for 3- and 4-linked structures. Similarly,

C2 ions isolated from oligosaccharides required higher-energy CID to generate m/z 221

ions. As shown in Figure 3.3e, the m/z 221 diagnostic ion is the second most abundant

using beam-type CID (CE = 22 V), yet it is not detectable using ion trap CID (Figure

3.3f, q = 0.45, 1.4 V, and 300 ms activation). The data in Figure 3.3 demonstrate that the

capability of employing different types of CID is extremely important to optimize the

yields of key product ions, which is also critical to the success of the planned MSn

experiments.

Page 87: Structural Analysis of Carbohydrates by Mass Spectrometry

64

Figure 3.3 MS3 and MS4 experiments of a pentasaccharide, [α-D-Glcp-(1-4)]4-D-Glc-ABA. (a) Ion transfer scheme for conducting either beam-type CID or ion trap CID in the 3rd stage MS/MS for the formation of C2 ions (m/z 341) via MS3 [474(2-) 624 fragments]). (b) MS3 beam-type CID from Q3 to Q2, CE = 42 V; (c) MS3 ion trap CID in Q2, q = 0.45, 1.4 V, and 300 ms of activation. (d) Ion transfer scheme for conducting beam-type or ion trap CID in the 4th stage MS/MS for the formation of m/z 221 diagnostic ions via MS4 [474(2-) 624 341 fragments] (e) MS4 beam-type CID from Q3 to Q2, CE = 22 V; (f) MS4 ion trap CID in Q2, q = 0.45, 1.4 V, and 300 ms of activation. The structure of [α-D-Glcp-(1-4)]4-D-Glc-ABA, is shown in the inset.

300

282222

624264246204 462365 444

0

100

%

m/z200 300 400 500 600 700

624

462444300282

341263 606179

c)

Beam CID

Trap CID

100 200 300 400

161

179341281

323143 263

m/z0

100 f)

Precursor ion Ion for further CID

Trap CID

221

Y3, 624C2, 341

342

1

5Unit #

b)

200 300 400 500 600 7000

100

%

m/z

341

%

MS3

MS3

MS4

161

221179143

263131 341100 200 300 400m/z

0

100e) Beam CID

%

MS4

[α-D-Glcp-(1-4)]4-D-Glc-ABA

Q1 Q2 Q3

Iso 474 Iso 624Beam CID

MSAE

c)Beam CIDd)Trap

CIDion transfer

MMM

))))))))c)BBBBBBBBeam) CCCCCICICICIDDDDdd)Trap)Trap

CICICICICICICICICICICCIC DDDDDDDDDDDDCIDCICICICICICICICIDDDDDDDD

Q1 Q2 Q3

Iso 474 Beam CID Iso 624

Iso 341TrapCID

MSAE

e)Beam CIDf)Trap

CIDion transfer

MMMMM

))))))))e)BBBBBBBBeam) CCCCCICICICIDDDDf)Trapf)T

CICICICICCICICICICICICICICIDDDDDDDDDDDDDDDCIDCICICICICCICICIDDDDDDDDp

a) MS3: [474 624 fragments]

d) MS4: [474 624 341 fragments]

Page 88: Structural Analysis of Carbohydrates by Mass Spectrometry

65

3.3.4 Determination of Individual Sugar Identity and Anomeric Configuration from Pentasaccharides

The feasibility of determining individual sugar unit identity and anomeric

configuration within an oligosaccharide was tested with two model pentasaccharides: [α-

D-Glcp-(1-4)]4-D-Glc-ABA and [β-D-Glcp-(1-4)]4-D-Glc-ABA, isomers having - and

β-anomeric configurations, respectively. To determine the structural information for

sugar units from the non-reducing end (i.e. unit 5 2), the following ions need to be

produced at MS2: m/z 341 (C2), 786 (Y4), 624 (Y3), and 462 (Y2), respectively. Both

singly (1-) and doubly (2-) deprotonated parent ions were formed from ESI in the

negative mode. Beam-type CID (CE = 21 – 24 V) of 2- charge state parents yielded

higher intensities for most product ions thus beam-type CID was chosen for MS2 prior to

higher stages of MSn. Using sugar unit 3, for example, the set of MS5 experiments

required to obtain its stereo-structural information were: [474(2-

) 624 341 221 fragments]. Based on previous study, ion trap CID was used for

MS3 of m/z 624 ions and beam-type CID was used for MS4 CID of m/z 341. Ion transfer

scheme specific to this MS5 experiment is shown in Figure 3.4a and CID conditions of

each step were summarized in Table 3.1. MS5 CID of m/z 221 was obtained for both

pentasaccharide isomers differing only in their anomeric configurations. For [α-D-Glcp-

(1-4)]4-D-Glc-ABA, the m/z 221 fragment derived from sugar unit 3 (Figure 3.4b)

yielded a fragmentation pattern characteristic of α-D-Glcp-GA (Figure 3.4d, error bars

from 7 experiments acquired over a 1 yr period). Spectral matching to the D-aldoHex-GA

standard database (16 structures: 8 sugar identities × 2 anomeric configurations) was

based on similarity scores calculated as described in the experimental section. α-D-Glc

Page 89: Structural Analysis of Carbohydrates by Mass Spectrometry

66

was identified as the top match for sugar unit 3 with a high similarity score (0.9771) as

compared to the second match (β-D-Ido, 0.9171, Table 3.1). Figure 3.4c shows the MS5

CID spectrum of [β-D-Glcp-(1-4)]4-D-Glc-ABA acquired from the same MSn sequence.

This spectrum is markedly different from the corresponding α-anomer (Figure 3.4b) and

almost identical to the ion trap CID data of synthetic β-D-Glcp-GA (Figure 3.4e). Here,

β-D-Glc was the top match (similarity score: 0.9654) in comparing to the database of

standards. The alternate anomers matched poorly (scores of 0.75-0.86, Table 3.1). The

data in Figure 3.4 demonstrate that even for a subtle structural difference, i.e. anomeric

configuration, sugar unit 3 could be confidently determined.

Figure 3.4 MS5 experiments of pentasaccharides, [α-D-Glcp-(1-4)]4-D-Glc-ABA and [β-D-Glcp-(1-4)]4-D-Glc-ABA: (a) Ion transfer scheme for conducting MS5 CID for both pentasaccharides. MS5 ion trap CID of (b) [α-D-Glcp-(1-4)]4-D-Glc-ABA and (c) [β-D-Glcp-(1-4)]4-D-Glc-ABA. MS2 ion trap CID of synthetic standards: (d) α-D-Glcp-GA and (e) β-D-Glcp-GA.

[α-D-Glcp-(1-4)]4-D-Glc-ABA [ -D-Glcp-(1-4)]4-D-Glc-ABA

0

100

%

100 150 200m/z

131

101

16111387 221159

(b) MS5

100 150 200m/z

131

161

87 159 221

(c)

0

100

%

MS5

(d)

100 150 200m/z

0

100

%87

99101

113129

131

159161203

221

MS2(e)

87101113129

131

159

161

221

100 150 200m/z

0

100

%

MS2

a) MS5:[474 624 341 221 frag]

Q1 Q2 Q3

Iso 474 Beam CID Iso 624

Iso 341TrapCID

MSAE

Beam CID

TrapCID

Iso 221

Page 90: Structural Analysis of Carbohydrates by Mass Spectrometry

67

Determining sugar units 5 and 4 involved MS4 [474(2-) 341 221 fragments]

and MS5 [474(2-) 786 341 221 fragments] experiments. The data are shown in

Figures 3.5 and 3.6 for [α-D-Glcp-(1-4)]4-D-Glc-ABA and [β-D-Glcp-(1-4)]4-D-Glc-

ABA, respectively. Conditions for each MSn stage are summarized in Table 3.2. Sugar 2

could not be identified since m/z 221 ions were not formed by CID of the Y2 ion. Table

3.1 summarizes the top 3 similarity scores of diagnostic ions derived from the two

pentasaccharides against the D-aldoHex-GA standard database. A full list is reported in

Table 3.3. The highest scores were assigned to the correct identity and anomeric

configuration for each sugar unit. The stereo-structures of 3 individual sugar units were

successfully determined for each pentasaccharide based on MSn (n = 4 or 5) experiments

and spectral matching to standards.

Page 91: Structural Analysis of Carbohydrates by Mass Spectrometry

68

Figure 3.5 MSn (n = 4 or 5) for the determination of stereo-structural information of sugar units 3-5 from [α-D-Glcp-(1-4)]4-D-Glc-ABA. Spectra (a) to (c) show the sequential MS2 to MS4 data based on [474(2-) 341 221 fragments] to determine sugar unit 5 structural information. Spectra (d) to (f) show the sequential MS3 to MS5 data based on [474(2-) 786 341 221 fragments] to determine sugar unit 4. Spectra (g) to (i) show the sequential MS3 to MS5 data based on [474(2-) 624 341 221 fragments] to determine sugar unit 3. The experimental conditions for each step can be found in Table 3.2.

221

161143 179

263

383

425 786444341

282 606 7680

100

%

200 300 400 500 600 700m/z

800 0

100

150 200 250 300 350m/z

131

101

113 159 2211611299987 141 2030

100

100 150 200m/z

0

100

150 200 250 300 350m/z

900

545

587474768503444341 7860

100

%

300 400 500 600 700 800m/z

161

179

221 263281341

338

131

159101 221113 16199

12987 141 2030

100

100 150 200m/z

[M-2H]2-

a) b) c)

d) e) f)

624

Precursor ionIon for further CID

[α-D-Glcp-(1-4)]4-D-Glc-ABA34 2 15Unit#

0

100

I%

m/z200 300 400 500 600 700

624

462444300282

341263 606179

161

221179143

263131 341100 200 300

m/z

0

100

0

100

100 150 200m/z

131

101

16111387 221159

g) h) i)

MS2 MS3 MS4

MS3 MS4 MS5

MS3 MS4 MS5

Page 92: Structural Analysis of Carbohydrates by Mass Spectrometry

69

Figure 3.6 MSn (n = 4 or 5) for the determination of stereo-structural information of sugar units 3-5 from [β-D-Glcp-(1-4)]4-D-Glc-ABA. Spectra (a) to (c) show the sequential MS2 to MS4 data based on [474(2-) 341 221 fragments] to determine sugar unit 5 structural information. Spectra (d) to (f) show the sequential MS3 to MS5 data based on [474(2-) 393(2-) 341 221 fragments] to determine sugar unit 4. Spectra (g) to (i) show the sequential MS3 to MS5 data based on [474(2-) 624 341 221 fragments] to determine sugar unit 3. The experimental conditions for each step can be found in Table 3.2.

131

161

221

m/z

87 101100 150 200

m/z

282

474

503 786393341

[M-2H]2-

222545

587

200 500 800

161

221179

143341

m/z100 200 300 400

393

282

624341

m/z200 400 800600

161

221

179143

263 341

m/z100 200 300

131

161

221

87 101

m/z100 150 200

624

a) b) c)

d) e) f)

Precursor ionIon for further CID

[ -D-Glcp-(1-4)]4-D-Glc-ABA34 2 15Unit#

0

100

%

0

100

%

0

100

0

100

0

100

0

100

400

200 300 400 500 600m/z

462

444300

282

179263

624341222 606

100 200 300 400m/z

161

143 179 221113 263

100 150 200m/z

131

161

87 159 221730

100

%

0

100

0

100g) h) i)

MS2 MS3 MS4

MS3 MS4 MS5

MS3 MS4 MS5

Page 93: Structural Analysis of Carbohydrates by Mass Spectrometry

70

Table 3.1 The top 3 ranked D-aldoHex candidates based on spectral similarity scores (indicated in parenthesis).

Like any MSn experiment, each stage of MS/MS enhances the selectivity of the

experiment at the cost of reduced ion signal. The signal loss arises from partitioning of

parents among fragments and ion loss during isolation/dissociation. Six orders of

magnitude reduction of signal was observed in MS5 experiments. This led to long signal

averaging periods to achieve good signal/noise (S/N) ratios. For instance, Figure 3.4b

accumulated 31160 scans (~11 hr) to reach a S/N of 50, using 6-7 μg of sample. The

number of scans, however, can be lowered using the spectral matching approach. For

example, fewer scans (3422, ~1.2 hr) than the spectrum shown in Figure 3.4b with a

lower S/N (7) still provides the first-ranked spectral similarity score (0.9475) against α-

D-Glcp-GA (second match: β-D-Idop-GA, 0.8928). Note that these spectra could not be

obtained at all using only one type of dissociation methods, for example; there was

simply not enough m/z 221 product ion to afford a spectrum without beam-type CID.

sample Unit # Spectral Similarity Score1st 2nd 3rd

[α-D-Glcp-(1-4)]4-D-Glc-ABA

3 α-D-Glc(0.9771)

β-D-Ido(0.9171)

β-D-Alt(0.9084)

4 α-D-Glc(0.9939)

β-D-Alt(0.9082)

β-D-Ido(0.9070)

5 α-D-Glc(0.9862)

β-D-Alt(0.8913)

β-D-Ido(0.8902)

[β-D-Glcp-(1-4)]4-D-Glc-ABA

3 β-D-Glc(0.9654)

α-D-Ido(0.9407)

α-D-Alt(0.9253)

4 β-D-Glc(0.9802)

α-D-Ido(0.9649)

α-D-Alt(0.9498)

5 β-D-Glc(0.9748)

α-D-Ido(0.9639)

α-D-Alt(0.9499)

Page 94: Structural Analysis of Carbohydrates by Mass Spectrometry

1

Tabl

e 3.

2 E

xper

imen

tal c

ondi

tions

for c

ondu

ctin

g M

Sn exp

erim

ents

.

Supe

rscr

ipts

indi

cate

the

figur

es, a

bove

, fro

m w

hich

cor

resp

ondi

ng M

S da

ta w

as d

eriv

ed.

Sam

ple

Uni

t #M

S2M

S3M

S4M

S5

[α-D

-Glc

p-(1

-4)] 4

-D

-Glc

-AB

A

347

4(2-

)[S

2a]

Q1

Q2

Beam

(24

V)47

4(2-

)62

4[S

2g]

Q2

Trap

(q =

0.4

5, 1

.4 V

, 300

ms)

474(

2-)

624

341

[S2h

]

Q3

Q2

Beam

(24

V)47

4(2-

)62

434

122

1[S

2i, 4

a]

Q3

Trap

(q =

0.2

3, 1

5 m

V, 1

00 m

s)

447

4(2-

)Q

1Q

2 B

eam

(23

V)47

4(2-

)78

6[S

2d]

Q2

Trap

(q =

0.4

5, 1

.8 V

, 100

ms)

47

4(2-

)78

634

1[S

2e]

Q3

Q2

Beam

(24

V)47

4(2-

)78

634

122

1[S

2f]

Q3

Trap

(q =

0.2

3, 1

8 m

V, 1

00 m

s)

547

4(2-

)Q

1Q

2 B

eam

(23

V)47

4(2-

)34

1[S

2b]

Q3

Q2

Beam

(23

V)47

4(2-

)34

122

1[S

2c]

Q3

Trap

(q =

0.2

3, 1

8 m

V,1

00 m

s)

------

-----

------

-----

------

----

[β-D

-Glc

p-(1

-4)] 4

-D

-Glc

-AB

A

347

4(2-

)Q

1Q

2 Be

am (2

3 V)

474(

2-)

624

[S3g

]

Q2

Trap

(q =

0.4

5, 1

.4 V

, 300

ms)

474(

2-)

624

341

[S3h

]

Q3

Q2

Beam

(22

V)47

4(2-

)62

434

122

1[S

3i,4

b]

Q3

Trap

(q =

0.2

3, 1

4 m

V, 1

00 m

s)

447

4(2-

)Q

1Q

2 B

eam

(21

V)47

4(2-

)39

3(2-

)[S

3d]

Q3

Q2

Beam

(15

V)47

4(2-

)39

3(2-

)34

1[S

3e]

Q3

Q2

Beam

(16

V)47

4(2-

)39

3(2-

)34

122

1[S

3f]

Q3

Trap

(q =

0.2

3, 1

2 m

V, 1

00 m

s)

547

4(2-

)[S

3a]

Q1

Q2

Bea

m (2

2 V)

474(

2-)

341

[S3b

]

Q3

Q2

Beam

(23

V)47

4(2-

)34

122

1[S

3c]

Q3

Trap

(q =

0.2

3, 1

5 m

V, 1

00 m

s)---

-----

------

------

-----

------

-

71

Page 95: Structural Analysis of Carbohydrates by Mass Spectrometry

2

Tabl

e 3.

3 S

pect

ral s

imila

rity

scor

es fo

r eac

h un

it in

olig

osac

char

ides

aga

inst

16

D-a

ldoH

ex-G

A st

anda

rds.

Th

e nu

mbe

rs in

red

indi

cate

the

high

est s

core

s (f

irst m

atch

) w

hile

the

num

bers

in b

lack

bol

d in

dica

te th

e se

cond

hig

hest

sco

res

(sec

ond

mat

ch).

Sam

ple

Uni

t #

Aldo

hexo

se-G

A Sy

nthe

ticSt

anda

rds

α-D

-All

β-D

-All

α-D

-Alt

β-D

-Alt

α-D

-Gal

β-D

-Gal

α-D

-Glc

β-D

-Glc

α-D

-Glu

β-D

-Glu

α-D

-Idoβ-

D-Id

oα-

D-M

anβ-

D-M

anα-

D-T

alβ-

D-T

al

[α-D

-Glc

p-(1

-4)] 4

-D

-Glc

-AB

A

30.

8094

0.76

730.

8656

0.90

840.

8994

0.78

570.

9771

0.85

560.

8543

0.85

720.

8599

0.91

710.

8140

0.86

310.

9008

0.78

38

40.

7940

0.75

440.

8863

0.90

820.

8844

0.81

930.

9939

0.86

290.

8006

0.86

510.

8723

0.90

700.

8137

0.85

210.

8567

0.74

11

50.

7701

0.76

850.

8862

0.89

130.

8517

0.83

120.

9862

0.85

970.

7785

0.88

360.

8867

0.89

020.

8361

0.82

070.

8406

0.71

83

[β-D

-Glc

p-(1

-4)] 4

-D

-Glc

-AB

A

30.

4633

0.73

890.

9253

0.59

290.

6621

0.81

450.

7962

0.96

540.

6712

0.82

730.

9407

0.69

910.

8466

0.53

120.

7047

0.59

96

40.

4708

0.72

860.

9498

0.60

320.

6647

0.86

360.

7910

0.98

020.

6544

0.83

910.

9649

0.67

100.

8520

0.56

270.

6690

0.60

40

50.

4099

0.70

750.

9499

0.54

270.

6094

0.87

210.

7559

0.97

480.

6092

0.85

270.

9639

0.64

760.

8507

0.51

320.

6443

0.57

28 72

Page 96: Structural Analysis of Carbohydrates by Mass Spectrometry

73

3.4 Conclusions

An MSn approach to pinpoint the stereo-structures (sugar identity, anomeric

configuration, and location) of individual sugar units within linear oligosaccharides was

described. The approach involved first optimizing the isolation of disaccharide units as an

ordered set of overlapping substructures via glycosidic bond cleavages. Subsequently,

cross-ring cleavages were optimized for individual disaccharides to yield key diagnostic

product ions (m/z 221). High confidence stereo-structural determination was achieved by

matching MSn CID of the diagnostic ions to synthetic standards via a spectral similarity

scores. By using this approach, structural information (identity and amoneric

configuration) of 3 individual sugar units (out of 5) was successfully determined for two

isomeric pentasaccharides. To our knowledge, this is the first report showing that specific

sugar identities, their anomeric configurations, and their positions within

oligosaccharides can be determined by MS. The above described approach of higher

stages of MSn could be used as a complementary means to NMR for oligosaccharide

structural analysis while offering enhanced sensitivity. A larger pool of oligosaccharides

will be tested for the applicability in future studies. In the case of branched

oligosaccharides (N-glycans), exoglycosidases will be used to convert the branched

structure into linear structures before subjected mass spectrometric analysis. We will also

explore different types of reducing end modifications which could enhance fragmentation

at selected bonds within the oligosaccharides. Several instrumentation methods are being

evaluated for enriching minor fragment ions at any MSn stage for improved sensitivity.

Page 97: Structural Analysis of Carbohydrates by Mass Spectrometry

74

3.5 References

(1) Fang, T. T.; Bendiak, B., J. Am. Chem. Soc. 2007, 129, 9721-9736.

(2) Konda, C.; Bendiak, B.; Xia, Y., J. Am. Soc. Mass Spectrom. 2012, 23, 347-358.

(3) Laine, R. A., Glycobiology 1994, 4, 759-767.

(4) Xia, Y.; Thomson, B. A.; McLuckey, S. A., Anal. Chem. 2007, 79, 8199-8206.

(5) Harvey, D. J., J. Am. Soc. Mass Spectrom. 2000, 11, 900-915.

(6) Bendiak, B.; Fang, T. T., Carbohydr. Res. 2010, 345, 2390-2400.

(7) Zhang, Z., Anal. Chem. 2004, 76, 3908-3922.

(8) Domon, B.; Costello, C. E., Glycoconjugate J. 1988, 5, 397-409.

(9) Dallinga, J. W.; Heerma, W., Biol. Mass Spectrom. 1991, 20, 215-231.

(10) Zaia, J., Mass Spectrom. Rev. 2004, 23, 161-227.

(11) Thomson, B. A. Apparatus and method for MSnth in a tandem mass spectrometer system. U.S. Patent 7,145,133 B2, December 5, 2006.

(12) Thomson, B. A.; Jolliffe, C. L. Spectrometer with axial field. U.S. Patent 5,847,386, December 8, 1998.

(13) Loboda, A.; Krutchinsky, A.; Loboda, O.; McNabb, J.; Spicer, V.; Ens, W.; Standing, K., Eur. J. Mass Spectrom. 2000, 6, 531-536.

(14) Collings, B. A., J. Am. Soc. Mass Spectrom. 2007, 18, 1459-1466.

(15) Morris, M.; Pierre, T.; B., R. K., J. Am. Soc. Mass Spectrom. 1994, 5, 1042-1063.

(16) Carroll, J. A.; Willard, D.; Lebrilla, C. B., Anal. Chim. Acta 1995, 307, 431-447.

Page 98: Structural Analysis of Carbohydrates by Mass Spectrometry

75

CHAPTER 4 Z1 IONS AS DIAGNOSTIC IONS FOR LINKAGE DETERMINATION

4.1 Introduction

In terms of linkage determination, disaccharides have been extensively studied as

model systems in the positive ion mode as metal ion adducts1-6 and in the negative ion

mode as deprotonated molecular ions7-11 or anion adducts of the molecules.12 Based on

the patterns of A-type product ions, such as m/z 221, 251, 263, and 281, formed by cross-

ring cleavages within the reducing sugar via various dissociation methods, the linkage

positions can be determined.1-15 Furthermore, the relative intensities of A ions as well as

the cleavage products from either side of the glycosidic oxygen provided additional

evidence for linkage assignment in the negative ion mode.11 Linkage determination for

oligosaccharides takes the same approach by detecting a certain combination of A ions.

Data analysis, however, can be much more challenging for oligosaccharides due to the

co-existence of isomeric and isobaric peaks from MS/MS, potentially derived from either

end of the molecule. Permethylation16 or reducing-end modification17 have therefore been

utilized to simplify the situation by providing mass discrimination of the product ions.

Currently, MS2 employing collision-induced dissociation (CID) is still the most widely

used strategy for linkage determination. However, depending on the nature of analytes

(trisaccharides or larger) and CID conditions, key diagnostic ions (i.e. A type ions) may

be missing, leading to either miss-assigned or un-assigned linkage positions.5,8,9 In order

Page 99: Structural Analysis of Carbohydrates by Mass Spectrometry

76

to maximize the structural information inherent in fragmentation patterns using MS2 for

oligosaccharide analysis, different activation/dissociation methods have been

investigated, including infrared multiphoton dissociation (IRMPD),18 electron capture

dissociation (ECD),19,20 electron-induced dissociation (EID),21 electron detachment

dissociation (EDD),22 and electron excited dissociation (EED).23 The application of the

above methods has been limited to research groups owing these instrumental capabilities.

Overall, MS2, the simplest MS/MS, has advantages such as good sensitivity, high

throughput, and simplicity in performance and instrumentation. However, structural

information to firmly establish anomeric configurations and stereochemistries of

individual monosaccharides and some of their linkages within oligosaccharides often

cannot be obtained using MS2 alone.

In this chapter, we describe an MSn approach to extract linkage information from

linear oligosaccharides together with their locations within the molecules with greater

confidence than is currently possible. We discovered that CID of Z1 ions derived from the

reducing sugar of deprotonated disaccharides showed distinct fragmentation patterns

virtually solely based on their linkage types and not significantly affected by the sugar

unit identities or their anomeric configurations. An MSn CID (n = 3-5) strategy was

developed that permitted linkages to be determined more confidently within a linear

oligosaccharide by decomposing the oligomer in the gas-phase into a set of overlapping

disaccharide structures with known origins within the oligosaccharide. Z1 ions were then

optimized for each disaccharide subunit, and repeatable CID spectra were obtained for Z1

ions derived from independent disaccharides. Spectral similarity scores were employed

for linkage determination via spectral matching of the thus obtained Z1 CID data to the

Page 100: Structural Analysis of Carbohydrates by Mass Spectrometry

77

standard disaccharide database. At a stage of proof-of-principle, this approach was

successfully demonstrated with two trisaccharides and a pentasaccharide.

4.2 Experimental

4.2.1 Materials

A list of 17 disaccharides and 3 oligosaccharides is shown in Table 4.1. All

samples were purchased from commercial sources (indicated by the superscripts in Table

4.1) and used without further purification. 4-Aminobenzoic acid (ABA), ethyl 4-

aminobenzoate (ABEE), H218O (97 atom %), chloroform, sodium cyanoborohydride,

acetic acid, and dimethyl sulfoxide (DMSO) were purchased from Sigma-Aldrich, Inc.

(St. Louis, MO, USA). Detailed procedures of 18O-labeling24 and reductive amination25,26

of reducing saccharides with ABA and ABEE were described in the previous chapters

(18O-labeling in chapter 2 and reductive amination in chapter 3). α1-2,3 Mannosidase

with G6 reaction buffer and BSA was purcharsed from New England Biolabs, Inc.

(Ipswich, MA, USA). Digestion condition was followed the procedure from

manufacture.

Page 101: Structural Analysis of Carbohydrates by Mass Spectrometry

78

Table 4.1 List of disaccharides and oligosaccharides being studied.

Analytes were purchased from: a Carbosynth, Ltd. (Berkshire, UK) and b Sigma-Aldrich, Inc. (St. Louis, MO, USA).

4.2.2 Reducing End Modification (Amination without Reduction)

Similar to the reductive amination except for using sodium cyanoborohydride was

used. Briefly, disaccharides or oligosaccharides (0.5 mg) were dissolved in 10 μL of a

solution containing 0.15 M ABA, ABEE, or APA in acetic acid: DMSO (3:7, v/v), and

the solution was incubated at 60 °C for 10-12 h. The solvent was removed under vacuum

(Speed-Vac, LABCONCO, Kansas City, MO) and the residue was dissolved in 300 μL

water followed by an addition of 300 μL chloroform. The aqueous phase containing the

derivatized sugar was collected after vortex mixing. This step was repeated with another

300 μL of water for extraction and the aqueous phases were combined and vacuum dried.

Page 102: Structural Analysis of Carbohydrates by Mass Spectrometry

79

The dried sample was dissolved in 100 μL DMSO and 900 μL methanol. This solution

was further diluted 100-fold in methanol prior to MS analysis.

4.2.3 Mass Spectrometry

All samples were analyzed in the negative ion mode on a hybrid triple

quadrupole/linear ion trap mass spectrometer (QTRAP4000 Applied Biosystems,

Toronto, Canada) equipped with a home-built nanoelectrospray ionization (nanoESI)

source. A schematic diagram of the instrument is shown in Scheme 2.2. Two types of low

energy collisional activation were accessible on this instrument, i.e. beam-type CID and

ion trap CID (via on-resonance single frequency excitation). In beam-type CID, the

precursor ions were isolated in Q1, accelerated to the Q2 collision cell for collisional

activation, and all products were analyzed in the Q3 linear ion trap. The collision energy

(CE, in the lab frame) was defined by the difference of the dc rod offsets between Q0 and

Q2. For ion trap CID, the precursor ions were isolated in the Q3 linear ion trap via

RF/DC mode and a dipolar excitation was applied at a Mathieu q value of 0.23. MS2 CID

experiments were carried out by either beam-type or ion trap CID. Higher stages of CID,

i.e., MS3 to MS5, were carried out by ion trap CID in Q3. MSn (n≥3) experiments were

performed through access to the original method table. Analyst 1.5 software was used for

instrument control, data acquisition, and processing. The typical parameters of the mass

spectrometer used in this study are listed in Supporting Information. The standard

spectra of Z1 ions (either m/z 163 from 18O-labeled glucose homodimers or m/z 161 from

intact glucose homodimers) were generated based on the average of 6 spectra over a 6-

month period. Standard deviations of peak heights were calculated for major 14

Page 103: Structural Analysis of Carbohydrates by Mass Spectrometry

80

fragments including m/z 73, 83, 89, 97, 101, 103, 113, 115, 131, 133, 135, 143, and 145

for CID spectra of m/z 163 and m/z 73, 83, 87, 89, 97, 101, 103, 113, 115, 125, 131, 133,

143, and 161 for CID spectra of m/z 161. These peaks were also used to calculate spectral

similarity scores.

4.2.4 HPLC

ABEE modified branched mannose 5 and exoglycosidase digested mannose 3

were separated using Agilent 1200 series HPLC system (Agilent Technologies, Santa

Clara, CA, USA). Separation was carried out on a HILIC (PolyLC, Inc., Columbia, MD,

USA) at a flow rate of 1 mL/min with a linear gradient of 20-35% solvent A in 30 min.

Solvent A was a mixture of 5mM ammonium acetate in water and solvent B contained 5

mM ammonium acetate in 100% acetonitrile. The eluent was detected at a wavelength of

305 nm.

4.3 Results and Discussion

4.3.1 Disaccharide Linkage Analysis Based on MS3 CID of Z1 Ions

Disaccharides, the smallest substructures of oligosaccharides containing linkage

information between two complete sugar molecules were used as model systems for

evaluating mass spectra of product ions that may yield detailed information about the

linkage positions. Initially, we used 18O-labeling at the carbonyl oxygen of 15

disaccharides to introduce a mass discrimination between X, Y, and Z vs. A, B and C

Page 104: Structural Analysis of Carbohydrates by Mass Spectrometry

81

product ions,27 recognizing that some of the ions might be isomeric and need to be

resolved as isotopomers. MS2 beam-type CID spectra of m/z 343 precursor ions of

glucose homodimers ( -anomeric configurations) with four different linkage positions

are shown in Figure 4.1. In these spectra, m/z 223, 253, 265, and 283 ions have

incorporated 18O (compared to the non-labeled data), suggesting that these fragments

contained a labeled component derived from the reducing sugar unit. Depending on the

disaccharide, some of the other product ions were observed as pairs varying in mass by 2

m/z units, the higher m/z isomer not having been observed in unlabeled disaccharides. Of

pertinence to this paper, peaks at m/z 161 and 163 were observed, indicative of two

isomers which would have been indistinguishable without 18O-labelling. A zoomed-in

region showing m/z 161 and 163 product ions is presented in insets for each linkage

position. Ions at m/z 163 should be Z1 ions, formed by cleavage at the reducing side of

the glycosidic oxygen. This ion species is unlikely to be formed as a sequential water loss

from Y1 ions (m/z 181), which we have not observed for MS2 CID of 18O-labeled

disaccharides. The assignment is also supported by the fact that MS3 CID of m/z 163

product ions showed totally different fragmentation patterns as compared to those

generated from H2O loss of 18O-labeled monosaccharides (data not shown). Furthermore,

MS3 CID of m/z 161 product ions (Figure 4.2) from the 18O-labeled disaccharides yielded

mass spectra that were very different from those derived from the m/z 163 ions,

suggesting that they have different origins and structures. These above results indicated

that from disaccharides, m/z 163 product ions were generated by dissociation pathways

unique to the substitution position of the reducing sugar.

Page 105: Structural Analysis of Carbohydrates by Mass Spectrometry

82

Figure 4.1 MS2 beam-type CID of 18O-labeled α-D-glucose homodimers, m/z 343 ([M-H]-), with different linkage positions: (a) 1-2, CE: 8 V; (b)1-3, CE: 12 V; (c) 1-4, CE: 10 V; and (d) 1-6, CE: 10 V. Insets in each of the spectra show the expanded region covering m/z 161 and 163.

161 163

265

179 343325223

161143119

101

161 163

179

343

221

323143101

161 325

281

113

251

113

163119

0

50

100

Rel

. Int

. (%

)

60 100 140 180 220 260 300 340m/z

(d) α-D-Glcp-(1-6)-D-Glc-18O

(a) α-D-Glcp-(1-2)-D-Glc-18O

(c) α-D-Glcp-(1-4)-D-Glc-18O

0

50

100

Rel

. Int

. (%

)

60 100 140 180 220 260 300 340m/z

0

50

100

Rel

. Int

. (%

)

60 100 140 180 220 260 300 340m/z

343

115

163343

179 283145113

22714389

161

253281

325

161 163

(b) α-D-Glcp-(1-3)-D-Glc-18O

0

50

100R

el. I

nt. (

%)

60 100 140 180 220 260 300 340m/z

343

343

343

163

179263 283103

161 161 163

281145143

343

Page 106: Structural Analysis of Carbohydrates by Mass Spectrometry

83

Figure 4.2 MS3 ion trap CID of m/z 161 product ions derived from 18O-labeled disaccharides ( -anomeric configuration) with different sugar identities and linkage positions: (a) α-D-Glcp-(1-2)-D-Glc, (b) α-D-Galp-(1-3)-D-Gal, (c) α-D-Glcp-(1-3)-D-Glc, (d) α-D-Manp-(1-4)-D-Man, (e) α-D-Glcp-(1-4)-D-Glc, (f) β-D-Galp-(1-4)-D-Man, (g) α-D-Glcp-(1-6)-D-Glc, and (h) α-D-Galp-(1-6)-D-Glc.

113

143131

161125101

(a) α-D-Glcp-(1-2)-D-Glc-18O

60 80 100 120 140 160m/z0

100R

el. I

nt. (

%)

50

343

161

(e) α-D-Glcp-(1-4)-D-Glc-18O113

131143

161125101838760 80 100 120 140 160m/z

0

100

Rel

. Int

. (%

)

50

343

161

113

131143

10183 16112587

(c) α-D-Glcp-(1-3)-D-Glc-18O

60 80 100 120 140 160m/z0

100

Rel

. Int

. (%

)

50

343

161

113

143131

16187 12510183

(g) α-D-Glcp-(1-6)-D-Glc-18O

60 80 100 120 140 160m/z0

100

Rel

. Int

. (%

)

50

343

161

60 80 100 120 140 160m/z0

100

Rel

. Int

. (%

)

50

113

143161131

60 80 100 120 140 160m/z0

100

Rel

. Int

. (%

)

50

113

131143

101 161

(b) α-D-Galp-(1-3)-D-Gal-18O343

161

(d) α-D-Manp-(1-4)-D-Man-18O343

161

73 8387

83 87

(f) β-D-Galp-(1-4)-D-Man-18O

(h) α-D-Galp-(1-6)-D-Glc-18O 60 80 100 120 140 160m/z

0

100R

el. I

nt. (

%)

50

113

131

14316110183

60 80 100 120 140 160m/z0

100

Rel

. Int

. (%

)

50

113

143131

161101

343

161

343

161

Page 107: Structural Analysis of Carbohydrates by Mass Spectrometry

84

MS3 ion trap CID spectra of m/z 163 ions (Z1 ions) were collected for 15 18O-

labele disaccharides. Figure 4.3 shows the data sets obtained from glucose homodimers

with 4 different linkage positions. Note that each panel is an average of 6 spectra (3

spectra each from α- and β-anomeric configurations) collected over 6 months. The error

bars on the major peaks represent the standard deviation of the six spectra. In the standard

spectra, CE was controlled to reduce the parent ion intensity to 18% of the base peak,

which was found to provide good reproducibility of the fragmentation patterns. Clearly,

CID of m/z 163 produces distinct fragmentation patterns characteristic of the linkage

positions. There are 13 common peaks (>3% relative intensity) observed from these

spectra, however with quite different relative intensities for different linkage positions.

These include ions at m/z 73, 83, 89, 97, 101, 103, 113, 115, 131, 133, 135, 143, and 145.

For the 1-2 linkage (Figure 4.3a), the base fragment peak is at m/z 103 and other

signature peaks include m/z 101, 115, 133, 135, and 145. We did notice that the α- and β-

anomers showed a small difference in the CID fingerprints, which was reflected by the

larger standard deviations of peak intensities at m/z 115 and 145 (22% and 9%),

respectively. These two peaks are of higher relative intensities in the β-anomer as

compared to the α-anomer. The comparisons of MS3 CID of m/z 163 from the α- and β-

anomers can be found in Figure 4.4. For the 1-3 linkage (Figure 4.3b), the base peak

shifts to m/z 115 and peaks at m/z 115 and 145 are relatively abundant. All other fragment

ions are present with relatively low intensities (<5%). The signatures of the 1-4 linkage

(Figure 4.3c) include dominant fragments at m/z 83 and 103 (base peak), while other

major fragments (m/z 73, 89, 97, 115, 133, and 145) have similar intensities around 20-

30%. The 1-6 linkage (Figure 4.3d) produces a relatively simple fragmentation spectrum

Page 108: Structural Analysis of Carbohydrates by Mass Spectrometry

85

with m/z 101 as the base peak, while other fragments (m/z 103, 131, and 135) show

relative intensities lower than 10%. Note that these fragmentation patterns are very

distinctive according to the linkage positions and repeatable as reflected by the relatively

small standard deviations of all major peaks. These characteristics of Z1 fragmentation

make it suitable as a diagnostic ion for linkage determination. With 18O-labeling on the

reducing sugar carbonyl group, dissociation profiles of the m/z 163 ion enable the linkage

position to be determined with high confidence.

Another interesting aspect that we discovered from CID of Z1 ions from 15

disaccharides was that neither the stereochemistry (identity) of the sugar units nor the

anomeric configuration (except for the 1-2 linkage discussed above) had any noticeable

effects on the fragmentation patterns. CID of Z1 ions from 7 disaccharides other than

glucose homodimers were compiled and shown in Figure 4.5. For example, the five

different 1-4 linked disaccharides (Figure 4.5a to e) showed almost identical

fragmentation patterns to that of the 1-4 linked glucose homodimer (Figure 4.3c). This

was consistently observed for other disaccharides containing 1-3 and 1-6 linkages (for

instance, compare Figure 4.5f to Figure 4.3b and Figure 4.5g to Figure 4.3d). Although

this phenomenon precludes obtaining stereo-structural information from the m/z 163 (Z1)

product ion (i.e., identity and anomeric configuration) for the reducing sugar unit, it is

advantageous for linkage determination due to the simplification of producing data from

standards and in making comparisons to unknowns for spectral analysis.

Page 109: Structural Analysis of Carbohydrates by Mass Spectrometry

86

Figure 4.3 Averaged MS3 ion trap CID data of Z1 ions (m/z 163) derived from 18O-labeled glucose homodimers with different linkage positions: (a) 1-2, (b) 1-3, (c) 1-4, and (d) 1-6. The error bars represent standard deviations of peak intensities based on 6 averaged spectra (3 spectra each from - and -anomers) obtained over a 6-month period.

(a) 1-2 linkage343

163

60 80 100 120 140 160 1800

100

Rel

. Int

., %

m/z

103

133101

163135

115

145

(b) 1-3 linkage115

145

163

343

163

60 80 100 120 140 160 1800

100

Rel

. Int

., %

m/z(c) 1-4 linkage

343

163

60 80 100 120 140 160 1800

100

Rel

. Int

., %

m/z

10383

73 133 16397 11589 145

(d) 1-6 linkage101

163131103 135

343

163

60 80 100 120 140 160 1800

100

Rel

. Int

., %

m/z

Page 110: Structural Analysis of Carbohydrates by Mass Spectrometry

87

Figure 4.4 MS3 CID spectra of m/z 163 product ions from (a) α-D-Glcp-(1-2)-D-Glc-18O and (b) β-D-Glcp-(1-2)-D-Glc-18O. Additional peaks (m/z 115 and 145) were found in the β configuration.

-H -β-D-Glcp-(1-2)-D-Glc-18O

103

133115

101 135145 163113 131 14360 80 100 120 140 160

m/z0

100

Rel

. Int

. (%

)

50

-H -α-D-Glcp-(1-2)-D-Glc-18O

(a) 343

163

60 80 100 120 140 160m/z

0

100

Rel

. Int

. (%

)

50

103

133101

163135

(b)

113 143

343

163

Page 111: Structural Analysis of Carbohydrates by Mass Spectrometry

88

Figure 4.5 MS3 ion trap CID of m/z 163 product ions from (a) α-D-Manp-(1-4)-D-Man-18O, (b) β-D-Galp-(1-4)-D-Glc-18O, (c) α-D-Galp-(1-4)-D-Gal-18O, (d) β-D-Galp-(1-4)-D-Man-18O, (e) α-D-Manp-(1-4)-D-Glc-18O, (f) α-D-Galp-(1-3)-D-Gal-18O, and (g) α-D-Galp-(1-6)-D-Glc-18O.

60 80 100 120 140 160m/z

0

100

Rel

. Int

. (%

)

50

60 80 100 120 140 160m/z

0

100

Rel

. Int

. (%

)

50

60 80 100 120 140 160m/z

0

100

Rel

. Int

. (%

)

50

60 80 100 120 140 160m/z

0

100

Rel

. Int

. (%

)

50

60 80 100 120 140 160m/z

0

100

Rel

. Int

. (%

)

50

115

14510383 163113

101

163131103 135

(f) α-D-Galp-(1-3)-D-Gal-18O

(a) α-D-Manp-(1-4)-D-Man-18O103

83

73 163115 13397 14589

(b) β-D-Galp-(1-4)-D-Glc-18O10383

73133 145 1631159789

(e) α-D-Manp-(1-4)-D-Glc-18O10383

73115 133 145 1639789

60 80 100 120 140 160m/z

0

100

Rel

. Int

. (%

)

50

10383

73133 145 1631159789

343

163

(c) α-D-Galp-(1-4)-D-Gal-18O

60 80 100 120 140 160m/z

0

100

Rel

. Int

. (%

)

50

10383

73133 1631451159789

343

163

(d) β-D-Galp-(1-4)-D-Man-18O

(g) α-D-Galp-(1-6)-D-Glc-18O

343

163

343

163

343

163

343

163

343

163

Page 112: Structural Analysis of Carbohydrates by Mass Spectrometry

89

4.3.2 The Effect of CID Conditions on the Formation of Z1 Ions

The data in Figure 4.3 demonstrate that CID of Z1 ions (m/z 163) can be used for

linkage determination. Note that the formation of Z1 ions was typically accompanied by

its structural isomer, ions at m/z 161 (insets of Figure 4.1). Collisional activation of the

m/z 161 ions from 18O-labeled disaccharides did not provide any distinguishable

fragmentation patterns according to their sugar unit identities, anomeric configurations,

or linkage types within the disaccharides. Although reducing-end labeling with 18O can

be used to clearly distinguish Z1 ions vs. m/z 161 ions, it was also highly desirable to

optimize the formation of Z1 ions while minimizing the abundance of their structural

isomer(s) (m/z 161) when studying native disaccharides or disaccharide units formed by

gas-phase dissociation of oligosaccharides. We investigated the effect of CID (beam-type

and ion trap) conditions on the formation of Z1 ions vs. their structural isomers using 18O-

labeled glucose homodimers (1-2, 1-3, 1-4, and 1-6 linkages and both α and β anomers).

The results are summarized in Figure 4.6 showing one anomer type as an example for

each linkage because the same trend of fragmentation behavior was observed for both

anomeric configurations. These spectra were collected using a wide isolation window

(around 5 Da) to observe both m/z 161 and 163 ions. Data from three different CEs (low,

medium, and high) under beam-type CID (left column) and ion trap CID (right column)

are compared side-by-side.

The best condition for forming Z1 ions from 1-2 linked disaccharides was the

lower energy ion trap CID (Figure 4.6a, 5 mVpp), from which m/z 163 accounted for 80%

of the summed intensities of m/z 161 and 163. Higher CE ion trap CID (7.5 and 10 mVpp)

significantly increased the relative intensity of m/z 161. Note that when beam-type CID

Page 113: Structural Analysis of Carbohydrates by Mass Spectrometry

90

was employed, m/z 161 was the dominant peak, while Z1 ions (m/z 163) were not

detected above noise level. On the other hand, 1-3 and 1-4 linked disaccharides showed

dominant formation of Z1 ions relative to m/z 161 under both beam-type and ion trap CID

conditions (Figure 4.6b and c, respectively). For 1-6 linked disaccharides, the best

conditions for the formation of Z1 ions were found using lower CEs for both beam-type

and ion trap CID as shown in Figure 4.6d. The data in Figure 4.6 clearly demonstrate that

the formation of m/z 161 and 163 is strongly affected by collisional activation conditions

and the best common condition to obtain relatively pure Z1 ion for all the linkage types

was to use ion trap CID with relatively low energy (around 5 mVpp). It should be noted

that even using these optimized conditions, the purity of Z1 ions from 1-2 linked

disaccharides was about 80%. After determining the CID conditions for preferential

formation of Z1 ions, the spectra in Figure 4.3 were re-collected for native unlabeled

glucose homodimers with relatively low energy ion trap CID. The data are shown in

Figure 4.7. Characteristic and highly reproducible fragmentation patterns were obtained,

which were almost identical to the CID of Z1 ions derived from 18O-labeled samples. The

data indicated that under the optimized CID conditions, the existence of small impurities

within the Z1 ions did not significantly affect their distinct fragmentation patterns. These

m/z 161 CID spectra were later used as standard spectra for Z1 ion fragmentation derived

from native disaccharides or disaccharide substructures of oligosaccharides.

We also performed MS3 CID of Z1 ions from two fructose containing

disaccharides (α-D-Glcp-(1-3)-D-Fru and β-D-Galp-(1-4)-D-Fru) with and without 18O-

labeling. Fructose has a ketone group at C2 position and 18O is incorporated in hydroxyl

group at C2 position as compared to the C1 position for aldohexoses. Collisional

Page 114: Structural Analysis of Carbohydrates by Mass Spectrometry

91

activation of Z1 ions from the two 18O-labeled fructose containing disaccharides showed

characteristic fragmentation patterns of 1-3 and 1-4 linkages (Figure 4.8a and c),

respectively. The fragment fingerprint patterns are identical to the data obtained from the

1-3 and 1-4 linked glucose homodimers (Figure 4.3b and c). The 1-4 linked fructose

containing disaccharide data (Figure 4.8c), however, showed 2 m/z shifts for fragments

such as m/z 75, 87, 99, 117, and 135, as compared to the aldohexose containing

disaccharide data (Figure 4.3c). The mass shifts are likely contributed by the difference

of 18O-labeling position between an aldohexose and a ketohexose. Using the optimized

conditions for Z1 ion formation (ion trap CID, 5 mVpp), MS3 CID of Z1 ions of the two

unlabeled fructose containing disaccharides also showed characteristic fragmentation

patterns of 1-3 and 1-4 linkages (Figure 4.8b and d). These results confirmed that MS3

CID of Z1 ions can be applied to disaccharide units containing fructose for determination

of linkage position.

Page 115: Structural Analysis of Carbohydrates by Mass Spectrometry

92

Figure 4.6 The effect of CID conditions on the formation of Z1 (m/z 163) vs. its structural isomers (m/z 161). Data were obtained from 18O-labeled glucose homodimers: (a) α-D-Glcp-(1-2)-D-Glc, (b) β-D-Glcp-(1-3)-D-Glc, (c) β-D-Glcp-(1-4)-D-Glc, and (d) α-D-Glcp-(1-6)-D-Glc.

161 163 161 163

510

15

5

2030

m/z m/z

(d) 1-6 linkage

Beam-type CID Ion trap CID

(c) 1-4 linkage

161 163 161 163

525

35

5

12.520

m/z m/z

161 163 161 163

57.5

10

m/z m/z

161 163 161 163

5

1015

5

7.5

10

m/z m/z

(b) 1-3 linkage

(a) 1-2 linkage

5

1015

Page 116: Structural Analysis of Carbohydrates by Mass Spectrometry

93

Figure 4.7 Averaged MS3 ion trap CID data of Z1 product ions (m/z 161) derived from native unlabeled glucose homodimers with different linkage positions: (a) 1-2, (b) 1-3, (c) 1-4, and (d) 1-6. The error bars show standard deviation of peak intensities based on 6 averaged spectra (3 spectra each from - and -anomers) obtained over a 6-month period. Ion trap CID with 5 mVpp was used for MS2 CID of m/z 341 step for (a) through (d).

(a) 1-2 linkage341

161

60 80 100 120 140 160 1800

100

Rel

. Int

., %

m/z(b) 1-3 linkage

341

161

60 80 100 120 140 160 1800

100

Rel

. Int

., %

m/z(c) 1-4 linkage

341

161

60 80 100 120 140 160 1800

100

Rel

. Int

., %

m/z

10183

73 133 16197 11587 143

(d) 1-6 linkage101

161

341

161

60 80 100 120 140 160 1800

100

Rel

. Int

., %

m/z

101

131161133

113

143

143161

113

131

Page 117: Structural Analysis of Carbohydrates by Mass Spectrometry

94

Figure 4.8 MS3 ion trap CID of the Z1 ion from 18O-labeled fructose containing disaccharides and MS3 ion trap CID of the Z1 ion generated by ion trap CID with low energy from native unlabeled fructose containing disaccharides, (a) α-D-Glcp-(1-3)-D-Fru-18O, (b) α-D-Glcp-(1-3)-D-Fru, (c) β-D-Galp-(1-4)-D-Fru-18O, and (d) β-D-Galp-(1-4)-D-Fru.

4.3.3 Linkage Determination for Oligosaccharides via MSn CID of Z1 Ions

Given that linkage positions can be determined from MS3 CID of Z1 ions from

deprotonated disaccharides, we further extended this method to linear oligosaccharides.

The approach involved optimizing glycosidic bond cleavages during early stages of MSn

so that an ordered set of overlapping disaccharide substructures could be formed,

generating Z1 ions from each disaccharide unit, and finally obtaining the fragmentation

fingerprints of Z1 ions for linkage determination. The concept is illustrated with a

pentasaccharide in Scheme 4.1, with the numbering of the sugar units starting from the

(b) α-D-Glcp-(1-3)-D-Fru

60 80 100 120 140 160m/z

0

100

Rel

. Int

. (%

)

50

113

143

161

341

161

(d) β-D-Galp-(1-4)-D-Fru

60 80 100 120 140 160m/z

0

100

Rel

. Int

. (%

)

50

10183

73133 161115 14397

341

161

87

(a) α-D-Glcp-(1-3)-D-Fru-18O

60 80 100 120 140 160m/z

0

100

Rel

. Int

. (%

)

50

115

145

163

(c) β-D-Galp-(1-4)-D-Fru-18O

60 80 100 120 140 160m/z

0

100

Rel

. Int

. (%

)

50

10383

75163135117 14587 99

343

163

343

163

Page 118: Structural Analysis of Carbohydrates by Mass Spectrometry

95

reducing-end (sugar 1). The key steps include: 1) dissociation of a precursor derivatized

at the reducing-end (M, Scheme 1) to a ladder of smaller oligosaccharides, each

successively one sugar unit shorter, containing the reducing-end tag (Ym ions); 2)

generation of disaccharide fragments (C2 ions) from each of the Ym ions; 3) dissociation

of the disaccharide units to form Z1 ions; 4) CID of Z1 ions and 5) spectral matching of

the Z1 CID data to the standard database (derived from disaccharides, Figure 4.3 or

Figure 4.7) for linkage determination. Linkage positions between the first and the second

sugar units could be directly obtained from 18O-labled oligosaccharides via MS3 CID of

m/z 163. The above approach has the advantage of knowing the exact origin of each

fragment with respect to its initial position within the oligomer. The reducing-end

modification M introduces a mass distinction between Y and C ions. In addition, the

group used to derivatize the reducing- end bears a negative charge and enabled the

selection of the ladder of charged Ym ions after MS2. Yet, through charge-transfer, it

permitted disaccharide (C2) ions to be formed and isolated. For example, to obtain

linkage information between sugar units 3 and 4, the following MS5 is needed: [[M-H]-

Y4 C2 Z1 fragments] (shown in Scheme 4.1).

Page 119: Structural Analysis of Carbohydrates by Mass Spectrometry

96

Scheme 4.1 The MSn approach for linkage determination of an oligosaccharide. “M” stands for reducing-end modification.

Two trisaccharides and one pentasaccharide were used to test the MSn approach

for extracting linkage information using this approach. The data in Figure 4.9 show a

series of experiments (MS2 to MS4) using 18O-labeled α-D-Glcp-(1-6)-α-D-Glcp-(1-4)-

D-Glc (Mw: 506 Da). In order to determine the linkage positions of linkages 1 and 2

(counting from the reducing- end), the following experiments were needed: MS3 CID

[505 163 fragments] and MS4 CID [505 341 161 fragments]. MS2 CID (CE =

15 V, beam-type) of the deprotonated molecular ions ([M-H]-) is shown in Figure 4.9a, in

which the formation of Z1 ions (m/z 163) and C2 ions (m/z 341) can be clearly seen. MS3

CID of Z1 ions (ion trap CID, 27 mVpp, Figure 4.9c) showed an almost identical spectrum

to the 1-4 linkage standard spectrum (Figure 4.3c). Therefore, linkage 1 can be

confidently identified as a 1-4 linkage. Note that the linkage could be assigned directly

from the m/z 163 ion derived from the labeled trisaccharide without requiring prior

2 1O M3O4 OO5

2 1O M3O4 OHO

3O4 OHHO

3 OH MS5 CID

Unique Dissociation

Pattern

DisaccharideStandardDatabaseSpectral Matching

LinkagePosition

Gas

-Pha

seD

isso

ciat

ion

OO 4OO 4Y4

OO3C2

3OO

Z1

Page 120: Structural Analysis of Carbohydrates by Mass Spectrometry

97

dissociation and isolation of the m/z 343 reducing-end disaccharide. This is particularly

worth emphasizing as the m/z 505 precursor ion gave rise to negligible quantities of the

m/z 343 product ion (Figure 4.9a), thus information could not be obtained for this linkage

either from the direct dissociation pattern of its reducing disaccharide (m/z 343) or from

the 163 product ion that may have been derived from it. Another important point is that

since m/z 161 ions are also formed in MS2 CID (due to sequential fragmentation of C1

ions), it is necessary to use m/z 163 ions to achieve correct linkage information for the

reducing end. Low energy MS3 CID (ion trap, 5 mVpp) of C2 ions (m/z 341, Figure 4.9a)

produced a reasonable intensity of m/z 161 ions (Figure 4.9b). Note that the best CID

conditions characterized from disaccharide studies (ion trap CID, 5 mVpp, 200 ms) were

used to favor the formation of Z1 ions relative to their structural isomers. Indeed, MS4

CID (ion trap, 35 mVpp) of m/z 161 showed a fragmentation pattern characteristic of the

1-6 linkage position (compare Figure 4.9d to the standard spectrum in Figure 4.7d). For

the other trisaccharide sample, 18O-labeled α-D-Galp-(1-3)-β-D-Galp-(1-4)-D-Gal, the

same set of MS3 and MS4 experiments were required to obtain information for linkages 1

and 2 : ([505 163 fragments] and [505 341 161 fragments]). Abundant Z1 and

C2 ions were formed under MS2 beam-type CID (CE = 15 V) as shown in Figure 4.10a.

MS3 CID (ion trap, 27 mVpp) of the Z1 ion clearly showed the distinct fragmentation

pattern of a 1-4 linkage position (Figure 4.9e). MS3 CID (ion trap, 5 mVpp) of C2 ions

(m/z 341) (Figure 4.10b) produced abundant Z1 ions (m/z 161). MS4 CID (ion trap, 35

mVpp) of the Z1 ions (m/z 161) optimized in abundance is shown in Figure 4.9f. The

fragmentation pattern matched well to the 1-3 linkage standard spectrum (Figure 4.7b)

and allowed confident assignment of this linkage position.

Page 121: Structural Analysis of Carbohydrates by Mass Spectrometry

98

Figure 4.9 MSn CID spectra for the determination of linkage types within trisaccharides. α-D-Glcp-(1-6)-α-D-Glcp-(1-4)-D-Glc-18O: (a) MS2 beam-type CID: [505 fragments], (b) MS3 ion trap CID: [505 341 fragments], (c) MS3 ion trap CID: [505 163 fragments], (d) MS4 ion trap CID: [505 341 161 fragments]. α-D-Galp-(1-3)-β-D-Galp-(1-4)-D-Gal-18O: (e) MS3 ion trap CID: [505 163 fragments] and (f) MS4 ion trap CID: [505 341 161 fragments].

(b)

60 140 220 300

179

323281

161221

341143311

m/z

2510

50

100

0

50

100

Rel

. Int

. (%

)

(a)

100 200 300 400 500

341 505

383179

323221143

161113

101 281 425

m/z

161 163

505

(c) 505

163

505

341

505

341

161

(d)

60 80 100 120 140 160m/z

101

161131

C2

C1Z1

α-D-Glcp-(1-6)-α-D-Glcp-(1-4)-D-Glc-18O, m/z 505

The parent ion The product ion for CID in the next stage

103

83

145133115 163738997

60 80 100 120 140 160m/z

0

50

100

Rel

. Int

. (%

)R

el. I

nt. (

%)

(e)

60 80 100 120 140 160m/z

0

50

100

Rel

. Int

. (%

)

83103

73 115 133145 1638997

505

163(f)

60 80 100 120 140 160m/z

113

143 161

505

341

161

α-D-Galp-(1-3)-β-D-Galp-(1-4)-D-Gal-18O, m/z 505

Page 122: Structural Analysis of Carbohydrates by Mass Spectrometry

99

Figure 4.10 MS2 and MS3 CID spectra from α-D-Galp-(1-3)-β-D-Galp-(1-4)-D-Gal-18O.(a) MS2 beam-type CID: [505 fragments], and (b) MS3 ion trap CID: [829 341 fragments].

In order to determine the individual linkages within the glucose homopentamer,

[β-D-Glcp-(1-4)]4-D-Glc using the MSn approach, formation of the following ions at the

MS2 stage was a prerequisite from the 18O-labeled sample: m/z 163 (Z1), 341 (C2), 505

(Y3), and 667 (Y4). MS2 CID of the deprotonated molecular ion (m/z 829, Figure 4.11a)

produced major peaks at m/z 341 (C2), 503 (C3), 665 (C4), and also a small amount of Z1

ions (m/z 163). However, no Y ions were detected above the noise level, which made it

impossible to conduct MSn to determine linkages 2 and 3. Reducing-end derivatization

with different types of functional groups has been shown to alter the fragmentation

patterns of oligosaccharides and generates X, Y, and Z product ions of

oligosaccharides.17 We tried reductive amination with ABA and ABEE at the reducing

α-D-Galp-(1-3)-β-D-Galp-(1-4)-D-Gal-18O, m/z 505

0

50

100

Rel

. Int

. (%

)

(a)

100 200

505161505

341

113 179

443143

323 425

163 C2Z1

C1

300 400 500

(b)

60 140 220 300m/z

0

50

100 505

341

161 341

179

113143 323281

Rel

. Int

. (%

)342

The parent ion The product ion for CID in the next stage

Page 123: Structural Analysis of Carbohydrates by Mass Spectrometry

100

sugar to form the reduced form of the corresponding Schiff base. Only data from ABEE

are discussed here since higher intensity of the desired ions were observed. The full series

of MS2 to MS5 spectra of ABEE modified [β-D-Glcp-(1-4)]4-D-Glc is shown in Figure

4.12. MS2 beam-type CID of the doubly deprotonated molecular ion, m/z 487 (Figure

4.12a), produced a variety of C and Y ions: m/z 341 (C2), 406 (Y42-), 503 (C3), 652 (Y3),

and 814 (Y4). The formation of reasonable intensities of Y3 and Y4 allowed further stages

of MS/MS. Finally, the following MSn CID sequences were employed for the

determination of linkages 1-4 individually: linkage 1, MS3 CID [829 163 fragments]

from the 18O-labeled sample; linkage 2, MS5 CID [487(2-

) 652 341 161 fragments] from the ABEE labeled sample; linkage 3, MS5 CID

[487(2-) 406(2-) 341 161 fragments] from the ABEE labeled sample; and linkage

4, MS4 CID: [829 341 161 fragments] from 18O-labeled sample. These four spectra

all showed characteristic fragmentation patterns of the 1-4 linkage position, allowing the

linkage position to be confidently assigned as 1-4. The data for individual stages of

MS/MS within each sequence of MSn CID can be found in Figure 4.11 (for the 18O-

labeled sample), and 4.12 (for the ABEE labeled sample).

Page 124: Structural Analysis of Carbohydrates by Mass Spectrometry

101

Figure 4.11 MS2 and MS4 CID spectra from [β-D-Glcp-(1-4)]4-D-Glc-18O. Experimental step for linkage 1 analysis is [829 163 fragments], (a) MS2 beam-type CID: [829 fragments], (b) MS3 ion trap CID: [829 163 fragments]. Experimental step for linkage 2 analysis is [829 341 161 fragments], (c) MS3 ion trap CID: [829 341 fragments], and (d) MS4 ion trap CID: [829 341 161 fragments].

--HZ1, m/z 163C2, m/z 341

60 140 220 300 380m/z

161

341

179281

(c)

0

50

100 83101

73 16113311597 14387 125

(d)

60 80 100 120 140 160m/z

0

50

100829

341(b)83 103

11589 1639775 13573 145

117

60 80 100 120 140 160m/z

0

50

100

Rel

. Int

. (%

)

829

163

829

341

161

(a)829

0

50

100

Rel

. Int

. (%

)

200 600 800m/z400

665503

341

587 829749425485

281323

161

161 163

C2

C3C4

Z1

[β-D-Glcp-(1-4)]4-D-Glc-18O, m/z 829

The parent ion The product ion for CID in the next stage

Page 125: Structural Analysis of Carbohydrates by Mass Spectrometry

102

Figure 4.12 MS2 to MS5 CID spectra from ABEE modified [β-D-Glcp-(1-4)]4-D-Glc. Experimental step for linkage 2 analysis is [487(2-) 652 341 161 fragments], (a) MS2 beam-type CID: [487(2-) fragments], (b) MS3 ion trap CID: [487(2-) 652 fragments], (c) MS4 ion trap CID: [487(2-) 652 341 fragments], and (d) MS5 ion trap CID: [487(2-) 652 341 161 fragments]. Experimental step for linkage 3 analysis is [487(2-) 406(2-) 341 161 fragments], (e) MS3 ion trap CID: [487(2-) 406(2-) fragments], (f) MS4 ion trap CID: [487(2-) 406(2-) 341 fragments], and (g) MS5 ion trap CID: [487(2-) 406(2-) 341 161 fragments].

As mentioned above, dissociation patterns of disaccharide ions (m/z 341) in the

negative ion mode have been used to assign linkages using sector instruments, triple

quadrupoles, Fourier transform ion cyclotrons and ion traps.7-11,13,24 Disaccharides having

2- or 6-linkages can be readily assigned in ion traps directly by MS2 as they yield

relatively abundant product ions of m/z 221 and 263, or m/z 221, 251 and 281,

respectively. For example, MS3 CID of the disaccharide substructure (α-D-Glcp-(1-6)-D-

(a)

200 400 600 800m/z

487310

652 814341 406 5035450

50

100

Rel

. Int

. (%

)-2H2-

Y3, m/z 652Y4, m/z 406(2-)

487(2-)

487(2-)

406(2-) 406(2-)

341

161

487(2-)487(2-)

406(2-)

341

200 400 600m/z

652

310

161407250

341

(e)

0

50

100

Rel

. Int

. (%

)

100 180 260 340m/z

161

341

179 281

(f)

0

50

100

60 80 100 120 140 160m/z

83

101

16173 115 143133

(g)

9787 1250

50

100

200 400 600m/z

652

328341 472 490310 606

(b)

0

50

100

Rel

. Int

. (%

)

100 180 260 340m/z

161

341

179

(c)

0

50

100

60 80 100 120 140 160m/z

83101

73133115 1619787 143125113

(d)

0

50

100487(2-)

652

487(2-)

652

341

652

341

161

487(2-)

C3C2

Y3Y42- Y4

[β-D-Glcp-(1-4)]4-D-Glc-ABEE, m/z 487(2-)

The parent ion The product ion for CID in the next stage

Page 126: Structural Analysis of Carbohydrates by Mass Spectrometry

103

Glc, m/z 341) from α-D-Glcp-(1-6)-α-D-Glcp-(1-4)-D-Glc-18O (Figure 4.9b) produced

abundant ions of m/z 221, 251, and 281, characteristic of the 1-6 linkage. However,

especially for 3-linkages and to some extent for 4 linkages, depending on the specific

sugars involved, the A ions derived by dissociation processes occurring within the

reducing sugar can be relatively low in abundance (frequently less than 5% of the base

product peak) with very abundant product ions observed at m/z 161 and 179 due to

cleavage on either side of the glycosidic oxygen. In these cases, further

isolation/dissociation of the m/z 161 ions are highly preferred specifically because, under

defined dissociation conditions, they are nearly entirely comprised of the diagnostic Z1

isomers derived from the reducing sugar (Fig. 4.6b and c). For example, MS3 CID of the

disaccharide substructure (α-D-Galp-(1-3)-D-Gal, m/z 341) derived from α-D-Galp-(1-3)-

β-D-Galp-(1-4)-D-Gal-18O (Figure 4.10b) only revealed one observable cross-ring

fragment at m/z 281 (signature fragments of the 1-3 linkage: m/z 251 and 281 but no

263), making it difficult to assign the linkage position. Thus for 3- and 4-linked

disaccharides, the CID data of their Z1 (m/z 161) ions are particularly useful. Another

key advantage is that the reducing sugar of an oligosaccharide can be selectively 18O-

labeled, whereby the Z1 product ion obtained directly from the reducing sugar (m/z 163)

can be used to determine the reducing-end linkage, without requiring prior isolation of a

reducing disaccharide fragment. In some cases this is very valuable because the reducing

disaccharide may only be produced as a trace product ion or not, apparently, at all.

Page 127: Structural Analysis of Carbohydrates by Mass Spectrometry

104

4.3.4 Linkage Determination for a Branched Oligosaccharide

In earlier discussions, we have shown that the Z1 ion approach can be applied to

linear oligosaccharides. However, in a real situation, branched oligosaccharides are often

encountered, such as N-linked glycans. We explored a way to convert a branched

oligosaccharide into linear structure using exoglycosidases and applied Z1 ion approach

for the smaller linear structures. Branched mannose 5 (Man5, having linkages of α1-3

and α1-6, structure shown in Scheme 4.2) is one of the common core structure in N-

glycans and was chosen as a model to test this method. First step is the chromophore

attachment to the reducing end to enable detection during later separation processes.

Schiff base formation with ABEE (without reduction) was employed for the initial tests.

Modified Man5 was purified by HPLC to eliminate substantial amount of impurities (i.e.

smaller oligosaccharides and isomers from the supply of commercial sample). ABEE

modified Man5 eluted around 10 min as shown in the chromatogram, Figure 4.13a.

There are two branching points within Man5 and they need to be cleaved to produce a

linear structure. An exoglycosidase, α1-2,3 Mannosidase, which cleaves off α1-2 and α1-

3 bonds was used to convert Man5 into a linear structure as shown in step 3. After

purification (step 4), a linear trisaccharide, α-D-Manp-(1-6)-α-D-Manp-(1-6)-D-Man-M,

was obtained. This ABEE modified Man3 came out around 7 min in Figure 4.13b. The

two HPLC separations (steps 2 and 4) are very important to eliminate possible impurities

and extract only the digested products originally from Man5. Step 5 is the delabeling of

chromophore and step 6 is 18O-labeling.

Page 128: Structural Analysis of Carbohydrates by Mass Spectrometry

105

Scheme 4.2 Sample preparation steps for branched oligosaccharide before MS analysis.

M

α1-2,3 Mannosidase digestion

Delabeling (M)

HPLC separation

MS analysis

Reducing end labeling (M)

HPLC separation

M

M

Step 1

Step 2

Step 3

Step 4

Step 5

Step 7

18O-labelingStep 618O

Page 129: Structural Analysis of Carbohydrates by Mass Spectrometry

106

A set of MS3 and MS4 experiments was required to obtain information for

linkages 1 and 2 within 18O-labeled α-D-Manp-(1-6)-α-D-Manp-(1-6)-D-Man:

([505 163 fragments] and [505 341 161 fragments]). Z1 and C2 ions were

formed under MS2 beam-type CID (CE = 20 V) as shown in Figure 4.14a. MS3 CID (ion

trap, 10 mVpp) of the Z1 ion clearly showed the distinct fragmentation pattern of a 1-6

linkage position (Figure 4.14c). MS3 CID (ion trap, 5 mVpp) of C2 ions (m/z 341) (Figure

4.14b) produced Z1 ions (m/z 161). MS4 CID (ion trap, 20 mVpp) of the Z1 ions (m/z 161)

is shown in Figure 4.14d. The fragmentation pattern of Figure 4.14c and d matched well

to the 1-6 linkage standard spectra (Figure 4.3d and 4.7d, respectively) and allowed

confident assignment of this linkage position.

The total amount of Man5 used for this experiment was only 1.2 nmol which is

compatible to permethylation method. Permethylation has the advantage of obtaining the

linkage information from the whole structure, it might experience the miss-assignment of

linkage positions due to ion suppression. In that case, Z1 method with a combination of

exoglycosidase can be used as a complimentary method to obtain confident linkage

information.

Page 130: Structural Analysis of Carbohydrates by Mass Spectrometry

107

Figure 4.13 HPLC chromatograms from (a) step 2 and (b) step 4 separations.

0

400

800

1200

1600

2000

0 3 6 9 12

mAU

Retention time (min)

0

20

40

60

80

100

120

140

0 3 6 9 12

mAU

Retention time (min)

Man5-M

Man3-M

(a) Step 2 HPLC sepration

(b) Step 4 HPLC sepration

Page 131: Structural Analysis of Carbohydrates by Mass Spectrometry

108

Figure 4.14 MSn CID spectra for the determination of linkage types within trisaccharides digested from Man5, α-D-Manp-(1-6)-α-D-Manp-(1-6)-D-Man-18O: (a) MS2 beam-type CID: [505 fragments], (b) MS3 ion trap CID: [505 341 fragments], (c) MS3 ion trap CID: [505 163 fragments], (d) MS4 ion trap CID: [505 341 161 fragments].

(b)

60 140 220 300m/z

0

50

100

0

50

100

Rel

. Int

. (%

)

(a)

100 200 300 400 500m/z

505

(c)505

163

505

341

505

341

161

(d)

60 80 100 120 140 160m/z

C2

α-D-Manp-(1-6)-α-D-Manp-(1-6)-D-Man-18O, m/z 505

60 80 100 120 140 160m/z

0

50

100

Rel

. Int

. (%

)R

el. I

nt. (

%)

383

341 443 505179 281 413

161 163

Z1

341

281

179

323221 251161

101

163103 135133

101

161131

Page 132: Structural Analysis of Carbohydrates by Mass Spectrometry

109

4.3.5 Algorithm-Assisted Linkage Determination

For the possibility of automation of the linkage analysis in future developments, it

is desirable to use an algorithm for spectral-matching. Spectral similarity scores can be

calculated between MSn CID spectra of Z1 ions from oligosaccharides following selected

dissociation pathways and the averaged CID spectra of Z1 ions from the disaccharide

standards originating from 2-, 3-, 4-, or 6-linkages. The same equation introduced in

Chapter 2 (Eq. 2.1) was used.

For disaccharide standards, two sets of standard spectra were prepared. One was

based on the averaged CID of m/z 163 spectra as shown in Figure 4.3 and the other was

generated based on the averaged CID of m/z 161 spectra formed under ion trap CID with

low energy from native glucose-dimers shown in Figure 4.7. The spectral similarity score

for CID of m/z 163 from unknowns was calculated against CID of m/z 163 from

standards while the spectral similarity score for CID of m/z 161 from unknowns was

calculated against CID of m/z 161 from standards. Table 4.2 summarizes the similarity

scores of MSn CID spectra of Z1 ions for each linkage derived from the three

oligosaccharides studied. The highest spectral similarity scores always corresponded to

the correct linkage types with the values very close to unity. Note that the similarity

scores for incorrect linkage assignments were typically smaller than 0.8.

Page 133: Structural Analysis of Carbohydrates by Mass Spectrometry

110

Table 4.2 Spectral similarity scores of MSn CID of Z1 ion spectra derived from oligosaccharides vs. disaccharide standards.

Oligosaccharides

Linkage (from

reducing-end)

Disaccharide Standards

1-2 1-3 1-4 1-6

α-D-Glcp-(1-6)-α-D-Glcp-(1-4)-D-Glc

1 0.7583 0.7053 0.9917 0.4566

2 0.8668 0.5212 0.7971 0.9944α-D-Galp-(1-3)-β-D-Galp-(1-4)-D-Gal

1 0.7603 0.7045 0.9912 0.4180

2 0.7223 0.9937 0.4805 0.4848

[β-D-Glcp-(1-4)]4-D-Glc

1 0.7722 0.6979 0.9702 0.5169

2 0.7504 0.4878 0.9917 0.7907

3 0.7619 0.5230 0.9875 0.7670

4 0.7511 0.5253 0.9938 0.7762α-D-Manp-(1-6)-α-D-Manp-(1-6)-D-Man

1 0.7807 0.5702 0.6247 0.9561

2 0.8907 0.5850 0.7598 0.9770

Page 134: Structural Analysis of Carbohydrates by Mass Spectrometry

111

4.4 Conclusions

A new approach for linkage determination of oligosaccharides has been evaluated

based on MSn (n = 3-5) CID of Z1 ions. MS3 CID of 18O-labeled disaccharides

(deprotonated ions) enabled the discovery of Z1 ions as “diagnostic ions” for linkage

determination. The fragmentation pattern of Z1 ions was only sensitive to their linkage

positions and not to their sugar identities and anomeric configurations. This unique

property allowed standard CID spectra of Z1 ions to be generated from a small set of

disaccharides (possibly from 4 disaccharides) that were representative of many other

possible isomeric structures. The formation of Z1 ions could be optimized using ion trap

CID at lower activation energies vs. their structural isomers. This enabled their analysis

to be performed on native disaccharides or disaccharide subunits formed by MS/MS of

larger oligosaccharides. This information can be used in conjunction with that from the

m/z 341 disaccharide ion dissociation patterns to provide highly confident assignments of

all linkages. It is worthy of note that for 3- and 4-linked disaccharides, information from

the Z1 ions of unlabeled disaccharides is especially valuable for their linkage assignments

as they highly predominate over other isomers under low or high energy dissociation

conditions. With 2- and 6-linked disaccharides, lower energy ion-trap conditions are

preferable to yield a preponderant Z1 ion. While a small amount of an isomeric species

cannot be eliminated, fragmentation patterns of the ions isolated under optimal conditions

still enabled the linkage to be assigned. Yet the 2- and 6-linkages yield distinct

fragmentation patterns directly from dissociation of their disaccharide precursors with

abundant and characteristic sets of product ions,5-9,28 so for a completely unknown,

unlabeled disaccharide, this information would be highly useful in conjunction with

Page 135: Structural Analysis of Carbohydrates by Mass Spectrometry

112

information from Z1 ion dissociation should there be any suspicion as to linkage

assignment. As we have not analyzed all stereochemical variants of Z1 product ions

arising from all linkages, it seems reasonable to surmise that higher statistical variance of

the linkage-characteristic dissociation patterns shown in Figures 4.3 and 4.7 may be

encountered, although observations so far indicate that linkage isomers dissociate with

dramatic differences. MSn CID of Z1 ions was applied to two trisaccharides and one

pentasaccharide to assess whether their linkage positions could be determined. By

comparing the MSn CID spectra of Z1 ions derived from different locations within

oligosaccharides with the standard CID spectra of Z1 ions from disaccharides, confident

assignments for individual linkages together with their locations within the oligomers

could be achieved for all three oligosaccharides studied as model compounds. This

method was further applied to a branched oligosaccharide, Man5. An exoglycosidase (α1-

2,3 Mannosidase) converted Man5 into a linear trisaccharide having the structure of α-D-

Manp-(1-6)-α-D-Manp-(1-6)-D-Man and two linkage information were successfully

obtained. The process of comparison of the Z1 ion spectra to standards was greatly

enhanced by a spectral similarity score algorithm, which provided numeric values in the

range of 0-1 with the highest scores indicative of the most likely assignment for the

linkage position. Although the examples were demonstrated on a hybrid triple

quadrupole/linear ion trap mass spectrometer, the MSn CID approach should in principle

be possible to implement on any standalone ion trap instrument, depending on the relative

abundances of isolated product ions derived from different oligosaccharide structures.

Page 136: Structural Analysis of Carbohydrates by Mass Spectrometry

113

4.5 References

(1) Domon, B.; Müller, D. R.; Richter, W. J., Org. Mass Spectrom. 1989, 24, 357-359.

(2) Spengler, B.; Dolce, J. W.; Cotter, R. J., Anal. Chem. 1990, 62, 1731-1737.

(3) Zhou, Z.; Ogden, S.; Leary, J. A., J. Org. Chem. 1990, 55, 5444-5446.

(4) Hofmeister, G. E.; Zhou, Z.; Leary, J. A., J. Am. Chem. Soc. 1991, 113, 5964-5970.

(5) Dongre, A. R.; Wysocki, V. H., Org. Mass Spectrom. 1994, 29, 700-702.

(6) Ashline, D.; Singh, S.; Hanneman, A.; Reinhold, V., Anal. Chem. 2005, 77, 6250-6262.

(7) Ballistreri, A.; Montaudo, G.; Garozzo, D.; Giuffrida, M.; Impallomeni, G.; Daolio, S., Rapid Commun. Mass Spectrom. 1989, 3, 302-304.

(8) Garozzo, D.; Giuffrida, M.; Impallomeni, G.; Ballistreri, A.; Montaudo, G., Anal. Chem. 1990, 62, 279-286.

(9) Dallinga, J. W.; Heerma, W., Biol. Mass Spectrom. 1991, 20, 215-231.

(10) Carroll, J. A.; Willard, D.; Lebrilla, C. B., Anal. Chim. Acta 1995, 307, 431-447.

(11) Mulroney, B.; Traeger, J. C.; Stone, B. A., J. Mass Spectrom. 1995, 30, 1277-1283.

(12) Guan, B.; Cole, R. B., J. Am. Soc. Mass Spectrom. 2008, 19, 1119-1131.

(13) Fang, T. T.; Zirrolli, J.; Bendiak, B., Carbohydr. Res. 2007, 342, 217-235.

(14) Guan, B.; Cole, R. B., Rapid Commun. Mass Spectrom. 2007, 21, 3165-3168.

(15) Carroll, J. A.; Ngoka, L.; Beggs, C. G.; Lebrilla, C. B., Anal. Chem. 1993, 65, 1582-1587.

(16) Reinhold, V. N.; Reinhold, B. B.; Costello, C. E., Anal. Chem. 1995, 67, 1772-1784.

(17) Chen, S.-T.; Her, G.-R., J. Am. Soc. Mass Spectrom. 2012, 1-11.

(18) Xie, Y.; Lebrilla, C. B., Anal. Chem. 2003, 75, 1590-1598.

(19) Adamson, J. T.; Håkansson, K., Anal. Chem. 2007, 79, 2901-2910.

(20) Zhao, C.; Xie, B.; Chan, S.-Y.; Costello, C.; O’Connor, P., J. Am. Soc. Mass Spectrom. 2008, 19, 138-150.

Page 137: Structural Analysis of Carbohydrates by Mass Spectrometry

114

(21) Wolff, J.; Laremore, T.; Aslam, H.; Linhardt, R.; Amster, I. J., J. Am. Soc. Mass Spectrom. 2008, 19, 1449-1458.

(22) Wolff, J.; Amster, I. J.; Chi, L.; Linhardt, R., J. Am. Soc. Mass Spectrom. 2007, 18, 234-244.

(23) Yu, X.; Huang, Y.; Lin, C.; Costello, C. E., Anal. Chem. 2012. 84, 7487-7494.

(24) Fang, T. T.; Bendiak, B., J. Am. Chem. Soc. 2007, 129, 9721-9736.

(25) Harvey, D. J., J. Am. Soc. Mass Spectrom. 2000, 11, 900-915.

(26) Chiesa, C.; Horváth, C., 1993, 645, 337-352.

(27) Konda, C.; Bendiak, B.; Xia, Y., J. Am. Soc. Mass Spectrom. 2012, 23, 347-358.

(28) Sheeley, D. M.; Reinhold, V. N., Anal. Chem. 1998, 70, 3053-3059.

Page 138: Structural Analysis of Carbohydrates by Mass Spectrometry

115

CHAPTER 5 THE EFFECTS OF REDUCING END MODIFICATION ON GAS-PHASE CEHMISTRY OF SMALL OLIGOSACCHARIDES

5.1 Introduction

Derivatization of sugars such as permethylation1,2 of hydroxyl groups and

reducing end modification3,4 is widely applied strategy in structural analysis of

carbohydrates prior to mass spectrometric analysis to achieve improved sensitivity. These

derivatizations largely affect gas-phase dissociation chemistry of glycan ions.5,6 One

popular approach for reducing end derivatization is amination (formation of Schiff base)

between an aromatic amine (e.g., 2-aminobenzamide,7 4-aminobenzoic acid ethyl ester8)

and a reducing sugar due to the high reaction efficiency and simple one-pot procedure.

Reductive amination is the reduced form of amination and the reducing sugar stays in

open-ring form.9,10 Her et al. previously reported that not only the type of derivatives but

also the structure of reducing sugar ring (open or closed) affected the fragmentation

patterns. They reported that closed-ring structure enhanced cross-ring cleavages as

compared to open-ring form.11

MSn (n>2) approach has been demonstrated on a modified triple-

quadrupole/linear ion trap instrument for stereo-structure characterization of individual

sugar units within small linear oligosaccharides in Chapter 3. This approach (as shown in

Scheme 3.1) involves optimizing the formation of an ordered set of overlapping

Page 139: Structural Analysis of Carbohydrates by Mass Spectrometry

116

disaccharide. Subsequently, the diagnostic product ion (m/z 221) having the structure of

non-reducing sugar glycosidically linked to a glycolaldehyde formed by cross-ring

cleavages were obtained. The success of this MSn approach largely depends on the

formation of the correct ladder of precursor ions within the tree of analysis, including Ym,

C2, and m/z 221 ions from C2. Stereochemistry and anomeric configuration of 3 sugar

units (3, 4, and 5 counting from the reducing-end sugar) in ABA modified (by reductive

amination) pentasaccharide anomeric isomers were successfully obtained. However,

sugar unit 2 information, which was theoretically possible to obtain by this approach, was

unavailable due to the absence of diagnostic m/z 221 ions from CID of Y2 ions.

Herein, a couple of N-derivatives (ABA, ABEE) were applied to model

disaccharides (α- and β-D-Glcp-(1-4)-D-Glc) by amination and reductive amination and

their fragmentation behavior, especially for the formation of m/z 221 ion, was studied.

Reducing-end derivatized disaccharides are considered to have the same structure with

the Y2 ions from oligosaccharides and can be used to simulate the condition without prior

MS/MS steps. Amination or reductive amination of N-derivatization showed a huge

difference on their fragmentation. Reductive N-derivatization produced no m/z 221 ions

while N-derivatization without reduction produced m/z 221 ions. This phenomenon is

consistent with that m/z 221 ion formation (from cross-ring cleavages) is favored by

having the reducing sugar in a closed-ring structure. N-derivatization without reduction

was then applied to trisaccharides. However CID of the deprotonated ion did not

produce Y2 ions. Since N-derivatization could not satisfy both criteria, we tested a new

derivatization: O-derivatization. O-derivatization gives the closed-ring structure of the

reducing-end sugar unit and the formation of m/z 221 ions was expected. To our

Page 140: Structural Analysis of Carbohydrates by Mass Spectrometry

117

knowledge, there is no systematic studies on the fragmentation behavior caused by

different types of derivatives and also this is the first time to introduce O-derivatized

sugars for mass spectrometric analysis. Among five derivatives (HBA, HBEE, HBSA,

1,2HNA, and 3,2HNA), HBA and HBSA showed a formation of m/z 221 ions and were

applied to trisaccharides to confirm whether they form Y2 ions or not. HBSA showed the

formation of Y2 ions seemed successful, however, this derivative failed when linkage

isomers of α- and β-D-Glcp-(1-4)-D-Glc were used.

5.2 Experimental

5.2.1 Materials

Eight types of reducing end modifications (shown in the table 1 with estimated

pKa values) were performed on a series of disaccharides and trisaccharides to study their

effect on the gas-phase fragmentation of oligosaccharide anions. β-D-Glcp-(1-4)-D-Glc,

α-D-Glcp-(1-6)-α-D-Glcp-(1-4)-D-Glc, [α-D-Glcp-(1-4)]2-D-Glc, and [β-D-Glcp-(1-4)]4-

D-Glc, H218O, chloroform, sodium cyanoborohydride, acetic acid, dimethyl sulfoxide

(DMSO), acetic anhydride, dichloromethane (DCM), boron trifluoride diethyl etherate

(BF3·Et2O), sodium methoxide were purchased from Sigma-Aldrich, Inc. (St. Louis, MO)

Octaacetyl α-D-Glcp-(1-4)-β-D-Glc from Carbosynth, Ltd. (Berkshire, UK). Detailed

procedures of 18O-labeling,12 reductive amination (R)3,13, and amination without

reduction (NR) of reducing saccharides were described in the previous chapters (18O-

labeling in chapter 2 and reductive amination in chapter 3, and amination without

reduction in chapter 4).

Page 141: Structural Analysis of Carbohydrates by Mass Spectrometry

118

Table 5.1 List of reducing end derivatives

.*Superscript e beside pKa value indicates estimation.

5.2.2 O-Derivatization at Reducing Sugars

Overall reaction steps are illustrated in Scheme 3.1 and (1) through (3) in

Scheme3.1 correspond to the each following section.

5.2.2.1 Acetylation of hydroxyl groups on sugars

Disaccharides or oligosaccharides (30 nmol) were dissolved in 60 to 100 μl of

acetic anhydride with 0.5 to 1 mg of iodine in an eppendorf tube and the reaction was

conducted with a stirring magnet for 3 h. The progress of the reaction was monitored by

MS for each step. After the completion of acetylation of hydroxyl groups on sugars,

DCM and aqueous sodium thiosulphate solution were added to the reaction mixture and

mixed thoroughly. The colorless organic layer was transferred to another eppendorf tube

and was washed with aqueous sodium carbonate solution to neutrality. The organic layer

was again transferred and dried completely by Speed-Vac.

CoreStructure R1 R2 (pKa of R2) Chemical Name (Abbreviation)

NH2COOH (4.87e)

COOCH2CH3 (-)4-aminobenzoic acid (ABA)

Ethyl-4-aminobenzoate (ABEE)

OHCOOH (4.57)

COOCH2CH3 (-)SO3H (2.54)

4-hydroxybenzoic acid (HBA)Ethyl 4-hydroxybenzoate (HBEE)

4-hydroxybenzenesulfonic acid (HBSA)

OH COOH (3.28e) 1-hydroxy-2-napthoic acid (1,2HNA)

OH COOH (3.02e) 3-hydroxy-2-napthoic acid (3,2HNA)

Page 142: Structural Analysis of Carbohydrates by Mass Spectrometry

119

5.2.2.2 O-derivatization to the acetylated sugars14

1 mg of acetylated sugar, 1.5 mg of derivatives (HBA, HBEE, HBSA, 1,2HNA,

or 3,2HNA), 1 μL of BF3·Et2O were dissolved in 500 μL of DCM under nitrogen. The

reaction mixture was placed in ultrasonic bath (B3500A-MT, VWR, Radnor, PA) for 1 to

2 h. After the completion of O-derivatization, aqueous sodium carbonate solution was

added and mixed thoroughly. The organic layer was transferred and dried completely by

Speed-Vac.

5.2.2.3 Deacetylation of O-derivatized Acetylated Sugars

1 mg of O-derivatized acetylated sugars and 1mg of sodium methoxide were

dissolved in 500 μL of methanol and incubated at 60 oC for 10-24 h.

5.2.2.4 Desalting of O-derivatized Sugars by HPLC

The reaction mixture of deacetylated O-derivatized sugars was desalted using

Agilent 1200 series HPLC system (Agilent Technologies, Santa Clara, CA). Desalting

was carried out on a HILIC (PolyLC Inc., Columbia, MD) column at a flow rate of 0.4

mL/min with a linear gradient of 20-35% solvent A in 30 min. Solvent A was water and

solvent B was acetonitrile. The eluent was detected at a different wavelength suitable for

individual chromophores (HBA: 310 nm, HBEE: 252 nm, HBSA: 271 nm, 1,2HNA: 352

nm, 3,2HNA: 352 nm).

Page 143: Structural Analysis of Carbohydrates by Mass Spectrometry

120

Scheme 5.1 Reaction steps for O-derivatization at reducing sugars

5.2.3 Mass Spectrometry

All samples were analyzed in the negative ion mode using a 4000Qtrap mass

spectrometer (Applied Biosystems/Sciex, Toronto, Canada) equipped with a home-built

nanoelectrospray (nanoESI) source. Two types of low energy collisional activation

methods were accessible on this instrument, i.e. beam-type CID and ion trap CID.

Analyst 1.5 software was used for instrument control, data acquisition, and processing.

The typical parameters of the mass spectrometer used in this study were set as follows:

spray voltage, -1.1 to -1.5 kV; curtain gas, 10; declustering potential, 50 V; beam-type

CID collision energy (CE), 5 to 30 V; ion trap CID activation energy (AF2), 5 to 60

(arbitrary units); scan rate, 1000 m/z/s; pressure in Q2, 5.0 x 10-3 Torr, and in Q3, 2.5 x

10-5 Torr.

Page 144: Structural Analysis of Carbohydrates by Mass Spectrometry

121

5.3 Results and Discussion

In order to investigate the major fragments formed from a native disaccharide,

MS2 beam-type and ion trap CID spectra of 18O-labeled β-D-Glcp-(1-4)-D-Glc were

collected and shown in Figure 5.1a. Product ions observed were m/z 161 (B1, -182 Da),

163 (Z1, -180 Da), 179 (C1, -164 Da), 221 (A2, -122 Da), 263 (A2, -78 Da), 283 (A2, -60

Da) and 325 (-18 Da). Product ions at m/z 161, 163, and 179 are formed from glycosidic

bond cleavages and m/z 221, 263, and 283 are formed from cross-ring cleavages. The

m/z 221 ion is the diagnostic ion for stereo-structure analysis and can be observed from

any disaccharides (although abundance may vary), however, Y1 ion (m/z 181) was never

observed in the product ions from CID of native disaccharide anions.

5.3.1 Gas-Phase Fragmentation Studies of N-Derivatized Disaccharides

MS2 beam-type and ion trap CID spectra of ABA derivatized disaccharide by

amination (β-D-Glcp-(1-4)-D-Glc-NR-ABA, MW: 461 Da) are shown in Figure 5.1b.

The major fragments included peaks at m/z 161 (B1), 178 (X0), 179 (C1), 192 (X0), 220

(X0), 221 (A2), 234 (X0), 262 (X0), 263 (A0), and 280 (Z1). The relatively intensity of m/z

161, 179, 221, 263 ions were similar to the β-D-Glcp-(1-4)-D-Glc-18O (Figure 5.1a),

however, a variety of X0 ions (cross-ring cleavages consisting of the reducing end) were

produced abundantly which suggested that the charge stayed on reducing end due to the

presence of carboxylic acid group in the derivative. As previously reported, beam-type

CID produced more fragments from cross-ring cleavages (A and X ions, especially m/z

221 ion) which required higher activation energies, while glycosidic bond cleavage (Z1

Page 145: Structural Analysis of Carbohydrates by Mass Spectrometry

122

ion formation) was enhanced by on-resonance ion trap CID (slow heating of the ion

which favors the lowest energy fragmentation channels).15

MS2 beam-type and ion trap CID spectra of ABA derivatized disaccharide by

reductive amination (β-D-Glcp-(1-4)-D-Glc-R-ABA, MW: 463) are shown in Figure

5.1c. These spectra showed major peaks at m/z 161 (B1), 179 (C1), 204 (X0), 222 (X0),

234 (X0), 246 (X0), 264 (X0), 282 (Z1), 300 (Y1), and 402 (X1). Unlike non-reduced

derivatization (Figure 5.1b), Y1 ion was formed, however, there were no A ions such as

m/z 221, 263, and 281. Similar phenomena were observed from the pare of β-D-Glcp-(1-

4)-D-Glc-NR-ABEE (MW: 489) vs. β-D-Glcp-(1-4)-D-Glc-R-ABEE (MW: 491) as

shown in Figure 5.1d and e, respectively. The A type ions were only observed from β-D-

Glcp-(1-4)-D-Glc-NR-ABEE while Y1 ion was only observed from β-D-Glcp-(1-4)-D-

Glc-R-ABEE. Comparing ABA and ABEE derivatives (e.g., Figure 5.1b and d, A, B,

and C type ions consisting of the non-reducing end were always observed in higher

intensities in ABEE than ABA which is consistent with the lower acidity of ABEE

derivative than ABA.

Page 146: Structural Analysis of Carbohydrates by Mass Spectrometry

123

Figure 5.1 MS2 beam-type vs. ion trap CIDs from (a) β-D-Glcp-(1-4)-D-Glc-18O, (b) β-D-Glcp-(1-4)-D-Glc-NR-ABA, (c) β-D-Glcp-(1-4)-D-Glc-R-ABA, (d) β-D-Glcp-(1-4)-D-Glc-NR-ABEE, and (e) β-D-Glcp-(1-4)-D-Glc-R-ABEE.

Z

Ion trap

X

Ion trap

282

222204 462300

179246

234

264C

Z

Y1X

282

462

222

179 300

246

402264

204161 CB

Z

Y1

X

Ion trap

248308

179 263290

488221161281

B

CZ

A

A

A

X

X

248

263 308179

488281220290161

ZC

B

A

AX

X

X

Ion trap310

179 490430328161

250 400Y1

Z

CBX

XX

310328

232250

179 490283

164

161

Z

Y1

CB XX

Ion trap

163

179343263283145

161

221AAA

CB

Z

163

343179325263 283

AAC

Z

Beam type

Beam type

Beam type Beam type

(a) 18O

343

(b) NR-ABA

460

(c) R-ABA

462

(d) NR-ABEE

488

(e) R-ABEE

490

100 200 300 400 5000

50

100R

el. I

nt. (

%)

280

263220

179 460442161192

Z

262C

B

XA

X 281A

161B

A221

100 200 300 400 5000

50

100

Rel

. Int

. (%

)

Beam type

B

C

220 280

179 262460

221

161 234192

178

A

263A

X

X

X

100 200 300 400 5000

50

100

Rel

. Int

. (%

)

100 200 300 400 5000

50

100

Rel

. Int

. (%

)

100 200 300 400 5000

50

100

234

100 200 300 400 5000

50

100

Rel

. Int

. (%

)

100 200 300 400 5000

50

100

Rel

. Int

. (%

)

164

100 200 300 400 5000

50

100

100 200 300 400 5000

50

100

100 200 300 400 5000

50

100

221A

X

X

X

X

m/z m/z

161

179 163

221

283

161

179 280

221 161

179 282

161

179 308

221

281

300

161

179 310

328

Page 147: Structural Analysis of Carbohydrates by Mass Spectrometry

124

5.3.2 Gas-Phase Fragmentation Studies of O-Derivatized Disaccharides

MS2 beam-type and ion trap CID spectra of β-D-Glcp-(1-4)-D-Glc-HBEE (MW:

489), β-D-Glcp-(1-4)-D-Glc-HBA (MW: 461), and β-D-Glcp-(1-4)-D-Glc-HBSA (MW:

497) are shown in Figure 5.2a, b, and c, respectively. β-D-Glcp-(1-4)-D-Glc-HBEE

showed major peaks at m/z 165 (Y0), 323 (B2), and 411 (X1) and small amount of m/z 221

ion was formed by beam-type CID (Figure 5.2a). α-D-Glcp-(1-4)-D-Glc-HBA (Figure

5.2b) showed major peaks at m/z 137 (Y0), 263 (A2), and 323 (B2). α-D-Glcp-(1-4)-D-

Glc-HBSA (Figure 5.2c) showed major peaks at m/z 173 (Y0), and 335 (Y1) while small

amount of m/z 221 (A2) ion was formed by both beam-type and ion trap CID. HBEE and

HBA derivatives showed abundant glycosidic bond cleavages between a reducing end

sugar and a derivative, forming Y0 and B2 ions depending on the charge locations. The

ratio of these two ions changes depending on the type of CID used. Bean-type CID

produced Y0 ions as the base peak while ion trap CID produced B2 ions as the base peak.

Based on these results, the majority of the original deprotonation sites can be considered

as located at the derivatized function groups and beam-type CID (fast reaction, < 1 ms)

produces abundant Y0 ions. The deprotonation sites are moved down to the sugar side

while ions were staying in the trap by ion trap CID (slow reaction, > 100 ms). HBSA

showed a little different effect on the formation of Y0 and B0 ions. Beam-type CID

produced 100% of Y0 ion (base peak) and about 40% of Y1 ion while no B2 ion was

formed. Ion trap CID reduced the formation of Y0 ion and increased the formation of Y1

ion, however, again, no B2 ion was generated. This is probably due to the higher acidity

of HBSA (pKa = 2.54) as compared to HBEE and HBA (pKa = 4.57) and the negative

charge tends to be localized on sulfonic acid group.

Page 148: Structural Analysis of Carbohydrates by Mass Spectrometry

125

MS2 beam-type and ion trap CID spectra of β-D-Glcp-(1-4)-D-Glc-1,2HNA

(MW: 511) and β-D-Glcp-(1-4)-D-Glc-3,2HNA (MW: 511) are shown in Figure 5.2d and

e, respectively. Both medication groups showed very similar fragments, having a base

peak at m/z 229 (X0), and fragments at m/z 187 (Y0), 313 (X0), and 331 (Z1). All of the

fragmentation appeared were X, Y, Z type of ions and no A, B, C ions were formed. This

is also due to the higher acidity of HNA (pKa=3.02 – 3.28) as compared to HBEE and

HBA (pKa = 4.57).

Page 149: Structural Analysis of Carbohydrates by Mass Spectrometry

126

Figure 5.2 MS2 beam-type vs. ion trap CID from (a) β-D-Glcp-(1-4)-D-Glc-HBEE, (b)β-D-Glcp-(1-4)-D-Glc-HBA, (c) β-D-Glcp-(1-4)-D-Glc-HBSA, (d) β-D-Glcp-(1-4)-D-Glc-1,2HNA, and (e) β-D-Glcp-(1-4)-D-Glc-3,2HNA.

100 200 300 400 5000

50

100R

el. I

nt. (

%)

100 200 300 400 5000

50

100

Rel

. Int

. (%

)165

323 489161221245178 411

B2BA A

Y0

X

323

411 489435165 244 453287

B2

Y0

X

Ion trap

Beam type

100 200 300 400 5000

50

100

Rel

. Int

. (%

)

100 200 300 400 5000

50

100

Rel

. Int

. (%

)

(a) HBEE

(b) HBA

(c) HBSA

100 200 300 400 5000

50

100

Rel

. Int

. (%

)

100 200 300 400 5000

50

100

Rel

. Int

. (%

)

100 200 300 400 5000

50

100

100 200 300 400 5000

50

100

100 200 300 400 5000

50

100

100 200 300 400 5000

50

100

(e) 3,2HNA

(d) 1,2HNA

Ion trap

Beam type

Ion trap

Beam type

137

323 461263245

B2A

Y0

323

461263137 245

B2

AY0

173

335497255

283 317

241

215

172Z

Y1

X

Y0

335

173

497417281215

221437

Y1

A

A XX

Y0

489

461

497

511

511

229

187 511313 331 391

ZX X

X

Y0

229

511187

361 375

157

443

313 467347 493

185

X

X

XY0

229157

185187 511479143167 473313

295

271

331493ZY0

X

XX

229

157 511187

331

313

493295

349271

483

185

Z

Y1Y0

XX

X

Ion trap

Beam type

Ion trap

Beam type X

X

m/z

m/z

161

179

221

323

165

161

179 323

137

335173

317

221

281

187331

187

331

A

A

Page 150: Structural Analysis of Carbohydrates by Mass Spectrometry

127

5.3.3 Comparisons between N-Derivatized (by Amination) and O-Derivatized Trisaccharides

Non-reduced form of ABA and ABEE, HBA and HBSA derivatized disaccharides

produced m/z 221 ions (diagnostic ions for stereo-chemical information) and these

derivatives are thus applied to trisaccharides to test whether Y2 ions can be formed or not.

The formation of Y2 ions is the key to obtain sugar 2 information from above listed

derivatized oligosaccharides because the MSn CID steps needed to follow for sugar 2 are:

[M-H]- Y2 221 fragments). MS2 beam-type and ion trap CID spectra of α-D-Glcp-

(1-6)-α-D-Glcp-(1-4)-Glc-NR-ABA and α-D-Glcp-(1-6)-α-D-Glcp-(1-4)-D-Glc-NR-

ABEE are shown Figure 5.3a and b, respectively. α-D-Glcp-(1-6)-α-D-Glcp-(1-4)-D-

Glc-NR-ABA showed major peaks at m/z 161 (B1), 179 (C1), 192 (X0), 220 (X0), 221

(A2), 262 (X0), 280 (Z1), 323 (A2), 341 (C2), 383 (A3), 425 (A3), and 562 (X2). α-D-Glcp-

(1-6)-α-D-Glcp-(1-4)-D-Glc-NR-ABEE showed very similar fragmentation patterns,

having different masses for X, Y, Z ions due to the different masses in derivatives.

However, no Y2 ions were produced from CID of α-D-Glcp-(1-6)-α-D-Glcp-(1-4)-D-

Glc-ABA and α-D-Glcp-(1-6)-α-D-Glcp-(1-4)-D-Glc-ABEE and these derivatizations

were not applicable to our MSn (n>2) approach.

O-derivatized trisaccharides were then examined to see the formation of Y2 ions

from MS2 CID. MS2 beam-type and ion trap CID spectra of [α-D-Glcp-(1-4)]2-D-Glc-

HBEE and [α-D-Glcp-(1-4)]2-D-Glc-HBSA are shown Figure 5.4b and c, respectively.

Spectra from [α-D-Glcp-(1-4)]2-D-Glc-HBEE showed a base peak at m/z 485 (B3) and

small amount of product ions at m/z 165(Y0), 323 (B2), 383 (A3), 425 (A3), and 573(X2).

As was previously observed in α-D-Glcp-(1-4)-D-Glc-HBEE, glycosidic bond cleavages

Page 151: Structural Analysis of Carbohydrates by Mass Spectrometry

128

between a reducing end sugar and a derivative was the main fragmentation site. α-D-

Glcp-(1-4)-D-Glc-HBEE showed different base peaks (Y0 for beam-type and B2 for ion

trap), there was no significant difference between beam-type and ion trap CIDs for [α-D-

Glcp-(1-4)]2-D-Glc-HBEE. Spectra from [α-D-Glcp-(1-4)]2-D-Glc-HBSA showed major

peaks at m/z 173 (Y0), 335, (Y1), 497 (Y2) with small amount of product ions at m/z 215

(X0), 221 (A2), 317 (Z2), 383 (A3), 479 (Z2), and 581 (A3). HBSA derivatized

trisaccharide produced a variety of product ions, most importantly having Y2 ion and m/z

221 ions.

Figure 5.3 MS2 beam-type CID from (a) α-D-Glcp-(1-6)-α-D-Glcp-(1-4)-D-Glc-NR-ABA, and (b) α-D-Glcp-(1-6)-α-D-Glcp-(1-4)-D-Glc-NR-ABEE.

0

50

100

Rel

. Int

. (%

)

100 200 300 400 500 600 700

0

50

100

Rel

. Int

. (%

)

100 200 300 400 500 600 700

383

220 622

341280

179

425

323

221161

562

C1

B1X2

A2

A2

C2

Z1

192 X0

248

650383308

341220

179 425

321

323

C1

A2

A2

C2

Z1

(a) NR-ABA

(b) NR-ABEE

622

650

m/z

221A2

Beam type

Beam type

A2

A2

X0

B2

161179

221323

341280

161179

221323

341308

B1161

323B2

m/z

B2

Page 152: Structural Analysis of Carbohydrates by Mass Spectrometry

129

Figure 5.4 MS2 beam-type vs. ion trap CID from (a) [α-D-Glcp-(1-4)]2-D-Glc-18O, (b)[α-D-Glcp-(1-4)]2-D-Glc-HEE, (c) [α-D-Glcp-(1-4)]2-D-Glc-HBA, and (d) [α-D-Glcp-(1-4)]2-D-Glc-HBSA.

100 200 300 400 5000

50

100

Rel

. Int

. (%

)

161341

425383179

505443281263221 323

163

B1

C1

A2 B2

C2Z1

100 200 300 400 5000

50

100

Rel

. Int

. (%

)

341

443 505161425179 281 487

163

C2

Z1

C1

B1

A2A2

100 200 300 400 500 6000

50

100 383

425

137 407485 623545221

179323161

311

587281

100 200 300 400 500 6000

50

100 485

545407

425383 623

605

587323

341305

461281

501221

381

B1

C1

Y2

C2

B3

B3 X2

Y0

B2

B2

100 200 300 400 500 6000

50

100

Rel

. Int

. (%

)

485

651383 512425 623573323165

100 200 300 400 500 600m/z0

50

100

Rel

. Int

. (%

)

485

651573425383 512 623

B3

B3

B2 X2

X2

A3Y0A3

100 200 300 400 500 6000

50

100 173

497335659

215 317 419 581221 479383

100 200 300 400 500 6000

50

100497

335 659215 383281

Y2

Y2

Y1

Y1

Z2Z1

Y0

X0

X0

X1

505 623

651

(c) HBA

(d) HBSA

659

(a) 18O

(b) HBEE

Ion trap

Beam type

Ion trap

Beam type

Ion trap

Beam type

Ion trap

Beam type

m/z

A3A3

A3

A3A3

A2

A2A2/Z2

A2

A3

A3

X2A3A2 A2

A3A3

A3

A2 A3

A2 A3

X2

161179

323

341163

161179

323

341281

137

323

165485 317173

335

479

497

Page 153: Structural Analysis of Carbohydrates by Mass Spectrometry

130

The relative intensity (normalized to the base peak) of the specific product ions

formed by different types of derivatizations was summarized in Table 5.2. Specific

product ions of our interest were m/z 221 ions from disaccharide-derivatives and Y2 ions

from trisaccharide-derivatives. Both criteria were only satisfied by HBSA. HBSA was

then derivatized to disaccharide linkage isomers (e.g., α-D-Glcp-(1-2)-D-Glc, α-D-Glcp-

(1-3)-D-Glc, β-D-Glcp-(1-3)-D-Glc, α-D-Glcp-(1-6)-D-Glc, and β-D-Glcp-(1-6)-D-Glc)

to test the applicability of this derivatives to a variety of disaccharides. Unfortunately,

m/z 221 ion could not be obtained from linkage isomers listed above.

Table 5.2 Summary of the normalized product ion intensity of characteristic ions by a variety of derivatizations

Relative intensity of product ions (%)

Type of derivativesDisaccharide Trisaccharide

m/z 221 ion Y2 ion

Native 5 0NR-ABA 15 0R-ABA 0 -

NR-ABEE 14 0R-ABEE 0 -

HBA 0 4HBEE 3 0HBSA 2 100

1,2HNA 0-

3,2HNA 0

Page 154: Structural Analysis of Carbohydrates by Mass Spectrometry

131

5.4 Conclusions

The fragmentation behavior of oligosaccharides can be affected by the chemical

nature of reducing-end derivatizations. HBSA was the only derivative which produced

both Y2 and m/z 221 ions from a derivatized model trisaccharide, [α-D-Glcp-(1-4)]2-D-

Glc-HBSA. However, this derivative only worked for 1-4 linked sugars and could not be

applied for other linkage isomers.

We continue to explore different derivatives, especially having the pKa value of

between 3.5 and 4.5 would be a good choice since HBSA (pKa = 2.54) was too acidic

and produced mainly Y ions while HBA (pKa = 4.57) was not acidic enough to produce

Y ions.

Page 155: Structural Analysis of Carbohydrates by Mass Spectrometry

132

5.5 References

(1) Ciucanu, I.; Kerek, F., Carbohydr. Res. 1984, 131, 209-217.

(2) Mechref, Y.; Kang, P.; Novotny, M. V., Solid-Phase Permethylation for Glycomic Analysis. In Glycomics, Packer, N.; Karlsson, N., Eds. Humana Press: 2009; Vol. 534, pp 53-64.

(3) Harvey, D. J., J. Am. Soc. Mass Spectrom. 2000, 11, 900-915.

(4) Harvey, D. J., J. Chromatogr. B 2011, 879, 1196-1225.

(5) Cancilla, M. T.; Penn, S. G.; Carroll, J. A.; Lebrilla, C. B., J. Am. Chem. Soc. 1996,118, 6736-6745.

(6) Orlando, R.; Allen Bush, C.; Fenselau, C., Biol. Mass Spectrom. 1990, 19, 747-754.

(7) Ruhaak, L. R.; Huhn, C. H.; Deelder, A. M.; Wuhrer, M., Anal. Chem. 2008, 80,6119-6126.

(8) Cheng, H.; Pai, P.; Her, G.-R., 2007, 18, 248-259.

(9) Her, G. R.; Santikarn, S.; Reinhold, V. N.; Williams, J. C., 1987, 6, 129-139.

(10) Nishikaze, T.; Kaneshiro, K.; Kawabata, S.; Tanaka, K., Anal. Chem. 2012, 84,9453-9461.

(11) Chen, S.-T.; Her, G.-R., J. Am. Soc. Mass Spectrom. 2012, 23, 1408-1418.

(12) Fang, T. T.; Bendiak, B., J. Am. Chem. Soc. 2007, 129, 9721-9736.

(13) Chiesa, C.; Horváth, C., 1993, 645, 337-352.

(14) Smits, E.; Engberts, J. B. F. N.; Kellogg, R. M.; van Doren, H. A., J. Chem. Soc. Perkin Trans. 1 1996, 0, 2873-2877.

(15) Zaia, J., Mass Spectrom. Rev. 2004, 23, 161-227.

Page 156: Structural Analysis of Carbohydrates by Mass Spectrometry
Page 157: Structural Analysis of Carbohydrates by Mass Spectrometry

133

VITA

Chiharu Konda was born on December 22, 1979, in Nagano, Japan. She is the

daughter of Yoshie and Hiromich Konda. While enrolling at Meiji Gakuin

Higashimurayama High School (Tokyo, Japan), she transferred to Notre Dame College

School (Ontario, Canada) and graduated from both high schools in 1998. Chiharu started

working as a sales engineer at one of the largest cell phone companies in Japan, NTT

Docomo in 1999. After working for 6 years, she pursued her dream to study abroad for

higher education and enrolled at Valdosta State University (Georgia, USA) in 2005. She

worked in Professor De La Garza’s lab from 2006-2008, focusing on the fabrication of

TiO2 film by electro-deposition. In August of 2009, Chiharu graduated with a Bachelor of

Science in Chemistry with Magna Cum Laude and began her graduate career in the

Chemistry Department at Purdue Univeristy. She joined Professor Yu Xia’s lab in the fall

of 2009, where her research focused on structural analysis of carbohydrates using mass

spectrometry. She defended her Ph. D. thesis in October 2013. Following graduate

school, she will be working at the new division of NTT Docomo where developing the

synergetic technology involving analytical science and cell phone network.

Page 158: Structural Analysis of Carbohydrates by Mass Spectrometry
Page 159: Structural Analysis of Carbohydrates by Mass Spectrometry

134

PUBLICATIONS

1. Du, Y. M.; Konda, C.; Xia, Y.; Ouyang, Z. “Statistical Analysis Model for Classifying Stereo-Structures of Oligosaccharides Using Tandem Mass Spectrometry”, InPreparation.

2. Konda, C.; Bendiak, B.; Londry, F. A.; Xia, Y. “Stereochemistry and AnomericConfiguration with Single-Sugar Resolution for Oligosaccharides via MSn (n>2)”, InPreparation.

3. Konda, C.; Bendiak, B.; Xia, Y. “Linkage Determination of Linear Oligosaccharides by MSn (n>2) Collision-Induced Dissociation (CID) of Z1 Ions in Negative Ion Mode Tandem Mass Spectrometry”, J. Am. Soc. Mass Spectrom., accepted.

4. Konda, C.; Bendiak, B.; Xia, Y. “Differentiation of the Stereochemistry and Anomeric Configuration for 1-3 linked Disaccharides via Tandem Mass Spectrometry and 18O-Labeling”. J. Am. Soc. Mass Spectrom., 2012, 23, 347-358.

Page 160: Structural Analysis of Carbohydrates by Mass Spectrometry

B American Society for Mass Spectrometry, 2011DOI: 10.1007/s13361-011-0287-5J. Am. Soc. Mass Spectrom. (2012) 23:347Y358

RESEARCH ARTICLE

Differentiation of the Stereochemistryand Anomeric Configuration for 1-3 LinkedDisaccharides Via Tandem Mass Spectrometryand 18O-labeling

Chiharu Konda,1 Brad Bendiak,2 Yu Xia1

1Department of Chemistry, Purdue University, West Lafayette, IN 47907-1393, USA2Department of Cell and Developmental Biology, and Program in Structural Biology and Biophysics, University of ColoradoDenver, Anschutz Medical Campus, Aurora, CO, USA

AbstractCollision-induced dissociation (CID) of deprotonated hexose-containing disaccharides (m/z 341)with 1–2, 1–4, and 1–6 linkages yields product ions at m/z 221, which have been identified asglycosyl-glycolaldehyde anions. From disaccharides with these linkages, CID of m/z 221 ionsproduces distinct fragmentation patterns that enable the stereochemistries and anomeric config-urations of the non-reducing sugar units to be determined. However, only trace quantities ofm/z 221ions can be generated for 1–3 linkages in Paul or linear ion traps, preventing further CID analysis.Here we demonstrate that high intensities of m/z 221 ions can be built up in the linear ion trap (Q3)from beam-type CID of a series of 1–3 linked disaccharides conducted on a triple quadrupole/linearion trap mass spectrometer. 18O-labeling at the carbonyl position of the reducing sugar allowedmass-discrimination of the “sidedness” of dissociation events to either side of the glycosidic linkage.Under relatively low energy beam-type CID and ion trap CID, anm/z 223 product ion containing 18Opredominated. It was a structural isomer that fragmented quite differently than the glycosyl-glycolaldehydes and did not provide structural information about the non-reducing sugar. Underhigher collision energy beam-type CID conditions, the formation of m/z 221 ions, which have theglycosyl-glycolaldehyde structures, were favored. Characteristic fragmentation patterns wereobserved for each m/z 221 ion from higher energy beam-type CID of 1–3 linked disaccharidesand the stereochemistry of the non-reducing sugar, together with the anomeric configuration, weresuccessfully identified both with and without 18O-labeling of the reducing sugar carbonyl group.

Key words: Oligosaccharides, Stereochemistry, Anomeric configuration, Collision-induceddissociation, Tandem mass spectrometry

Introduction

Carbohydrates play important roles in biological systems,such as providing energy to cells and functioning as

structural components for plant cell walls. By conjugatingwith proteins and lipids, carbohydrates are widely involvedin cell–cell interactions, cell signaling, and self and non-selfrecognition events [1, 2]. Carbohydrates can form almostunlimited variations in their structures due to their structuralcomplexity. In order to elucidate the structure of anoligosaccharide, it is necessary to characterize the stereo-chemistry of each monosaccharide unit, the anomericconfiguration of the glycosidic bonds, linkage positions,Received: 23 August 2011Revised: 20 October 2011Accepted: 21 October 2011Published nline: 18 November 2011

Electronic supplementary material The online version of this article(doi:10.1007/s13361-011-0287-5) contains supplementary material, whichis available to authorized users.

Correspondence to: Yu Xia; e-mail: [email protected]

o

135

Page 161: Structural Analysis of Carbohydrates by Mass Spectrometry

and the sequence of the individual monosaccharides inthe oligomer. When enough sample is available, higher-dimensional NMR is a powerful tool to obtain detailedstructural information [3–5].

Mass spectrometry is a widely applied method instructural analysis of carbohydrates, due to its capabilityof providing detailed molecular information and highsensitivity [6]. The molecular weight information ofcarbohydrates is readily obtained from soft ionizationmethods such as electrospray ionization (ESI) [7, 8] andmatrix-assisted laser desorption/ionization (MALDI) [9–11]. Tandem mass spectrometry based on collision-induced dissociation (CID) is heavily relied on to obtainstructural information for carbohydrates. Glycosidic bondcleavages and/or cross-ring cleavages are typically ob-served from collisional activation. Following the nomen-clature proposed by Domon and Costello , A, B, and Cions are fragments containing the non-reducing terminuswhile X, Y, and Z ions include the reducing end [12].Four types of fragments, B, C, Y, and Z ions, areformed from cleavages on either side of the glycosidicoxygen. B and Y ions are cleaved at the non-reducingside of a glycosidic oxygen and C and Z ions arecleaved at the reducing side of a glycosidic oxygen.Based on the specific mass differences of fragmentsresulting from glycosidic bond cleavages, sequenceinformation for both linear and branching oligosacchar-ides as well as glycotypes such as the complex, hybrid,or high-mannose can be determined for methylatedoligosaccharides [8, 13].

A and X ions result from cross-ring cleavages and theyare typically more informative for the structural analysis ofcarbohydrates. It has been shown that the relative abundan-ces of these ions can be correlated to the linkage position,and in some cases, the anomeric configuration and stereo-chemistry of each monosaccharide. For example, CIDspectra of deprotonated di- and oligosaccharide alkoxyanions in the negative ion mode showed distinguishablefragmentation patterns for each linkage position, which wassuccessfully applied to di-, tri-, and hexasaccharides [14–17]. Negative ion adducts [18, 19] and positive adducts [20,21] of deprotonated di- and oligosaccharides have also

enabled linkage positions to be determined, both in thenegative and positive ion modes, and linkage sites can beestablished for neutral disaccharides as positive ion adductsof lithium or sodium ions or as protonated adducts [22–27].Determination of anomeric configuration has been demon-strated for underivatized disaccharides in the negative ionmode [15, 17, 18], for derivatized 1-4- and 1-6-linkeddisaccharides [28], and in the positive ion mode with alkalimetals [22–27] and lead cationization [20]. Prior knowledgesuch as linkage position, ring form, and stereochemistry [15,17, 18, 20, 22–28], however, was typically required toassign anomeric configuration, which was difficult to applyto larger oligosaccharide systems [18]. CID of metalcationized monosaccharides derivatized as a Schiff basewith diethylenetriamine [29] and CID of metal cationized N-acetylhexosamine diastereomers [30] have been shown toproduce distinct fragmentation patterns according to thestereochemistry. The assignment of stereochemistry has alsobeen obtained by matching the CID spectrum of acetylatedmonosaccharides in oligosaccharides with reference spectra[31, 32]. Using additional statistical analysis, anomericconfiguration has been identified for metal cationizedglucopyranosyl-glucose disaccharides (1–2, –3, –4, and –6linkages) [26]. Also, the linkage positions and stereo-chemistries of non-reducing units can be discriminated forglucose, galactose, and mannose containing disaccharideshaving 1–2, –3, and –4 linkages [27].

Recently, a tandem mass spectrometry approach has beendeveloped to differentiate the stereochemistry and anomericconfiguration for the non-reducing unit of hexose-containingdisaccharides having any of the 16 possible stereochemicalvariants [33, 34]. In this method, diagnostic ions at m/z 221were formed from CID of deprotonated disaccharide ions(m/z 341). It was established that the m/z 221 ions consistedof the intact non-reducing sugar glycosidically linked toglycolaldehyde, as indicated in Scheme 1 (where GAabbreviates glycolaldehyde). Note that an open-chain formfor the reducing sugar is indicated in Scheme 1 and also forother disaccharides discussed later. This is based on theobservation of absorbance in the carbonyl stretch region invariable wavelength infrared radiation photo-dissociation ofdeprotonated monosaccharide anions in the gas phase [35].

H --HH

-H --HCID CIDCID CIDCID CID

D Gl (1 4) Gl / 341 D Gl GA / 221α-D-Glcp-(1-4)-Glc m/z 341 α-D-Glcp-GA m/z 221α D Glcp (1 4) Glc, m/z 341 α D Glcp GA, m/z 221

Scheme 1. Formation of glycosyl-GA anions at m/z 221 from CID of deprotonated 1–4 linked disaccharides

348 C. Konda et al.: Structure Determination for Disaccharides

136

Page 162: Structural Analysis of Carbohydrates by Mass Spectrometry

When m/z 221 ions were further dissociated by collisionalactivation, disaccharides having different non-reducing sugarunits and anomeric configurations showed distinct fragmen-tation patterns that matched synthetic glycosyl-GAs. Thismethod was shown to be useful for assigning the stereo-chemistry as well as the anomeric configuration of theglycosidic bond for the non-reducing sugar in disaccharideshaving 1–2, 1–4, and 1–6 linkages [33, 34]. However, due tothe low abundance of m/z 221 ions produced from 1–3linked disaccharides, MS3 CID of m/z 221 ions could not beperformed, and it was unclear whether the fragmentationpatterns could be used for assigning either their stereochem-istry or anomeric configuration.

Herein, a series of 1–3 linked disaccharides were studiedon a triple quadrupole-linear ion trap mass spectrometer(QTRAP 4000). MS3 CID data of m/z 221 ions from the 1–3linked disaccharides were obtained for the first time. Theformation of m/z 221 ions was examined using differentcollisional activation methods, i.e., beam-type CID and iontrap CID of the deprotonated disaccharides. 18O-labeling ofthe reducing carbonyl oxygen in 1–3 linked disaccharideswas used to enable mass-discrimination of structural isomersof the (usually) m/z 221 ions. By choosing the proper CIDconditions, the diagnostic m/z 221 ions (the glycosyl-GAs)could be formed as the dominant isomer. Their CIDfragmentation patterns could be used to establish thestereochemistry and anomeric configuration of the non-reducing sugar unit from 1–3 linked disaccharides.

ExperimentalMaterials

α-D-Glcp-(1–2)-D-Glc (kojibiose), β-D-Glcp-(1–2)-D-Glc(sophorose), α-D-Glcp-(1–3)-D-Glc (nigerose), α-D-Glcp-(1–3)-D-Fru (turanose), and H2

18O were purchased fromSigma-Aldrich, Inc. (St. Louis, MO, USA); β-D-Glcp-(1–3)-D-Glc (laminaribiose), α-D-Manp-(1–3)-D-Man (3α-manno-biose), and α-D-Galp-(1–3)-D-Gal (3α-galactobiose) werepurchased from Carbosynth, Ltd. (Berkshire, UK). α- and β-monosaccharide-glycolaldehyde standards, glucopyranosyl-glycolaldehydes (Glcp-GA), galactopyranosyl-glycolaldehydes(Galp-GA), and mannopyranosyl-glycolaldehydes (Manp-GA) were synthesized as previously described [33]. Dis-accharides and synthetic standards were dissolved in meth-anol to a final concentration of 0.01 mg/mL and NH4OHwas added to a final concentration of 1% immediately beforeuse.

18O-Labeling of Reducing Disaccharides

Disaccharides (1 mg) were dissolved in 100 μL of H218O for

3 to 10 d at room temperature. The solution was furtherdiluted to 0.1 mg/mL with methanol before mass spectro-metric analysis.

Mass Spectrometry

All samples were analyzed in the negative-ion mode on aQTRAP 4000 mass spectrometer (Applied Biosystems/SCIEX, Toronto, Canada) equipped with a home-builtnanoelectrospray ionization (nanoESI) source. A schematicpresentation of the instrument ion optics is shown in theSupporting Information, Figure S1. Two types of low energycollisional activation methods were accessible on thisinstrument, i.e., beam-type CID and ion trap CID. Inbeam-type CID, the precursor ions (m/z 341 or 343) wereisolated in Q1, accelerated in the Q2 collision cell forcollisional activation, and all products were analyzed in theQ3 linear ion trap. Collision energy (CE) was defined by thepotential difference (absolute value) between Q0 and Q2. Inion trap CID, the precursor ions were isolated in the Q3linear ion trap via the RF/DC mode and a dipolar excitationwas used for collisional activation. In order to perform iontrap CID at different Mathieu q-parameters, an AC (alter-nating current) generated from an external waveformgenerator (Agilent Technologies, Santa Clara, CA, USA)was used for resonance excitation. Frequency and the lowmass cut-off were calculated by SxStability (Pan GalacticScientific, Omemee, Ontario, Canada). MS3 CID experi-ments were carried out by first performing beam-type CIDof precursor ions in Q2. The fragment ions of interest wereisolated in Q3 and then subjected to ion trap CID. Analyst1.5 software was used for instrument control, data acquisi-tion, and processing. The typical parameters of the massspectrometer used in this study were set as follows: sprayvoltage, –1.1 to −1.5 kV; curtain gas, 10; declusteringpotential, 50 V; beam-type CID collision energy (CE), 5 to30 V; ion trap CID activation energy (AF2), 5 to 60(arbitrary units); scan rate, 1000 m/z; pressure in Q2, 5.0×10–3 Torr, and in Q3, 2.5×10–5 Torr. Ion injection time wascontrolled to keep a similar parent ion intensity: typically 3×106 counts per second (cps) for MS2 CID experiments and1×106 cps for MS3 CID experiments. Activation time waskept constant at 200 ms for all ion trap CID experiments.Seven spectra were collected for CID of m/z 221 ions fromsynthesized monosaccharide-GA standards (deprotonatedmolecules) and disaccharides over a 1 y period. Standarddeviations of peak heights were calculated for major frag-ments such as m/z 87, 99, 101, 113, 129, 131, 159, 161, 203,and 221, which were observed from all the standards anddisaccharides studied here except β-D-Glcp-GA and β-D-Glcp-(1–2)-D-Glc, which showed no peaks at m/z 99.

Results and DiscussionIon trap CID of deprotonated 1–3 linked disaccharides (m/z341) typically generates ions at m/z 221 in trace abundanceon a Paul trap instrument, and isolation or further CID of m/z221 ions have not been achieved before [14, 15, 34, 36]. The4000QTRAP mass spectrometer used in this study has aunique triple quadrupole-linear ion trap configuration,

C. Konda et al.: Structure Determination for Disaccharides 349

137

Page 163: Structural Analysis of Carbohydrates by Mass Spectrometry

offering high sensitivity due to the large capacity of thelinear ion trap, and allowing either beam-type or ion trapcollisional activation. In beam-type CID, the precursor ionswere isolated in Q1 and accelerated in Q2 for collisionalactivation, while ion trap CID was conducted in Q3 with adipolar excitation for collisional activation. Since CIDfragmentation patterns can be sensitive to the means ofactivation, the formation of m/z 221 ions from five 1–3linked disaccharides was investigated via both beam-typeand ion trap CID. Figure 1 compares the MS2 beam-type andion trap CID of deprotonated β-D-Glcp-(1–3)-D-Glc (m/z341) using low energy CID conditions. A relatively low CE(6 V) was used for beam-type CID; in ion trap CID, the AF2for an AC dipolar excitation was set to 25 (arbitrary units)for 200 ms. Under either activation condition, the absolute

intensities of m/z 221 ions (indicated by an arrow inFigure 1) were very low and their relative intensities wereless than 1% (normalized to the base peak in the spectrum).This phenomenon was generally observed for all 1–3 linkeddisaccharides studied herein. The insets in Figure 1 demon-strate the isolated m/z 221 ions (with a 2 m/z isolationwindow) from each set of dissociation conditions. For beam-type CID, 1×106 cps of m/z 221 ions could be accumulatedwith an injection time of 1 s, which was sufficient forperforming the next stage of tandem mass spectrometry(MS3 in this case) with reasonable ion statistics andsensitivity. Far lower abundance of the m/z 221 ions (4.6×104 cps) could be isolated from ion trap CID of m/z 341,even after doubling the injection time to 2 s. As a result, itwas not feasible to obtain MS3 CID for m/z 221 ionsgenerated from m/z 341 precursor ions initially isolatedwithin the trap. In experiments described below, beam-typeCID was used to dissociate disaccharide precursor anions inthe Q2 collision cell thereby generating m/z 221 product ionsin high enough abundance to acquire their spectra in thelinear trap reproducibly.

CID of m/z 221 Ions Generated from 1–3 LinkedDisaccharides

Previous studies have demonstrated that m/z 221 productions formed from collisional activation of disaccharideanions typically consist of an intact non-reducing sugar witha 2-carbon aglycone derived from the reducing sugar [34].Three dominant fragment peaks are commonly observedfrom CID of m/z 221: m/z 101, 131, and 161. The relativeintensities of these peaks, together with some other fragmentions, can be used to establish the fragmentation patterns andto distinguish the stereochemistry and anomeric configura-tion of the non-reducing sugar. Given that the CID patternsof m/z 221 ions will be used for structural identification,spectral reproducibility is an important issue. Similar to thefindings from a Paul trap instrument [33, 34], we noticedthat the number of ions (m/z 221) in the linear ion trap andthe energy input into an ion were among the most importantparameters affecting spectral reproducibility. To ensurereasonable ion statistics and avoid adverse space chargeeffects, the intensity of the m/z 221 ions was kept at 1×106 cps before MS3 CID. Based on previous studies, the CIDenergies were tuned so that the ratio of remaining precursorion to the most abundant product ion was kept around 18%±3% [33]. Figure 2a, b, e, and f were the averaged spectrafrom seven repetitions collected over a 1 y period, and theywere further used to make spectral comparisons in laterdiscussion. Error bars in the spectra indicate the standarddeviation of the peak intensity for 10 major fragment ions,which were frequently observed for all the disaccharidesstudied herein (m/z 87, 99, 101, 113, 129, 131, 159, 161,203, and 221). The standard deviations for these peaks wereless than 5% in most cases, indicating high reproducibility ofthe spectra from day to day by controlling the ion counts in

341100 341

( ) 221100(a) 1.05e6

2211.05e6

ps

% cp

y, % y,

sity

sity

ns ns

ten

te

Int

222In

e I 222

ive

at 161

Rel

161179

R

113113143143

100 140 180 220 260 300 3400

60 100 140 180 220 260 300 340m/z

060

m/z

179100 179100161(b) 4 6e4 221161 4.6e4 221

s

341% cps

341

y,%

y, c

ty ity

nsi nsi

en en

nte nte

e In In

veat

iv

113ela

113143

Re

143

R

060 100 140 180 220 260 300 340

060

Figure 1. MS2 CID spectra in the negative ion modeobtained from deprotonated β-D-Glcp-(1–3)-D-Glc (m/z 341)under low energy dissociation conditions: (a) beam-type CID(CE=6 V), and (b) ion trap CID (AF2=25). Insets in (a) and (b)show the isolation of m/z 221 ions generated from beam-typeCID (injection time = 1 s) and ion trap CID (injection time = 2 s),respectively

350 C. Konda et al.: Structure Determination for Disaccharides

138

Page 164: Structural Analysis of Carbohydrates by Mass Spectrometry

the trap before CID and the energy input to the ions. Sincethe CID patterns upon dissociation of m/z 221 ions can differto some extent from instrument to instrument [37], CIDspectra of the synthetic monosaccharide-GA were collectedas standards for comparisons. Figure 2a and b show the CIDdata of α- and β-D-Glcp-GA, respectively. The abundantpeaks at m/z 101 and 131 in Figure 2a are a signature of anon-reducing glucose with an α anomeric configuration.Note that a distinct fragmentation pattern is observed for theβ configuration (Figure 2b), where m/z 131 and 161 ions aredominant. The same collisional activation conditions wereapplied to m/z 221 ions derived from α-D-Glcp-(1–3)-D-Glcand β-D-Glcp-(1–3)-D-Glc, anomeric isomers containing anon-reducing glucose. It is obvious that the spectra from thetwo anomeric isomers (Figure 2c and d) were drasticallydifferent from their corresponding D-Glcp-GA standards,however, were similar to each other. This indicates that them/z 221 ions generated using low collision energies from m/z341 precursors have different structures from the D-Glcp-GAstandards, and that their CID patterns cannot be used to assign

either the stereochemistry or anomeric configurations of theions. Note that beam-type CID was used to generate the m/z221 ions from disaccharides shown in Figure 2, a conditiondiffering from previous studies where ion trap CID had beenused [33]. This difference in activation could have contributedto the formation of structural isomers observed for the m/z 221product ions. In order to test this hypothesis, m/z 221 ionsof 1–2 linked disaccharides, α-D-Glcp-(1–2)-D-Glc and β-D-Glcp-(1–2)-D-Glc, were formed using similar beam-type CIDconditions and further subjected to MS3 CID (Figure 2eand f). Except for a larger fluctuation in peak intensityfor m/z 203, almost identical fragmentation patterns tothe standards were observed (compare Figure 2a to e andb to f), strongly indicating that the expected D-Glcp-GAstructures were formed. We further investigated a widevariety of disaccharides and found that the CID patternsof m/z 221 ions matched with their corresponding monosac-charide-GA standards with the exception of 1–3 linkeddisaccharides when low collision energy beam-type CIDconditions were used to dissociate the disaccharides.

-

60 100 140 180 220m/z m/z

161

203131113

221

-H

α-D-Glcp-(1-3)-Glc, m/z 341

(c)

-H -α-D-Glcp-(1-2)-Glc, m/z 341

m/z 221 m/z 221

0

100

Rel

ativ

e In

ten

sity

, %

60 100 140 180 2200

100

Rel

ativ

e In

ten

sity

, %

87

101

113

129

131

159161

203

221

99

(e)

-H -α-D-Glcp-GA, m/z 221

60 100 140 180 220m/z

0

100

Rel

ativ

e In

ten

sity

, %

87

99101

113

129

131

159161203

221

(a)

(b)

-H -β-D-Glcp-GA, m/z 221

β-D-Glcp-(1-3)-Glc, m/z 341-H -

m/z 221

060 100 140 180 220

m/z

100

Rel

ativ

e In

ten

sity

, %

161

203131113

221

(d)

60 100 140 180 220m/z

0

100

Rel

ativ

e In

ten

sity

, %

87101 113

129

131

159

161

203

221

--H

β-D-Glcp-(1-2)-Glc, m/z 341

m/z 221

60 100 140 180 220m/z

0

100

Rel

ativ

e In

ten

sity

, %

87101113

129159

161

221

131

(f)

Figure 2. MS2 ion trap CID spectra of m/z 221 ions derived from synthetic standards (a) α-D-Glcp-GA, AF2=25 and (b) β-D-Glcp-GA, AF2=18. MS3 CID spectra of m/z 221 ions derived from low energy beam-type CID of glucose-containingdisaccharides: (c) α-D-Glcp-(1–3)-Glc, CE=6 V for MS2 and AF2=25 for MS3, (d) β-D-Glcp-(1–3)-Glc, CE=6 V for MS2

and AF2=25 for MS3, (e) α-D-Glcp-(1–2)-Glc, CE=5 V for MS2, and AF2=27 for MS3, and (f) β-D-Glcp-(1–2)-Glc, CE=5 V for MS2, and AF2=25 for MS3. The error bars in the spectra show the standard deviation of the peak intensitybased on seven spectra collected over a 1 y period

C. Konda et al.: Structure Determination for Disaccharides 351

139

Page 165: Structural Analysis of Carbohydrates by Mass Spectrometry

18O-labeling at the carbonyl position of the reducingsugar was used to mass-discriminate the “sidedness” ofdissociation events to either side of the glycosidic linkageand thus the origins of the m/z 221 and/or potential 223product ions. Figure 3 compares relatively low energy beam-type and ion trap CID of 18O-labeled deprotonated α-D-Glcp-(1–3)-D-Glc, m/z 343. Similar fragments were ob-served for both conditions; however, the ion abundance form/z 283 (loss of 60 Da, C2H4O2) was much higher in beam-type CID relative to ion trap CID. It is possible that thisfragmentation channel requires higher activation energy andis promoted, even in lower-energy beam-type CID, sincehigher collision energies (several eV) may have beenobtained as compared to ion trap CID (hundreds of meV).

The inset in Figure 3a shows data collected using a wideisolation window (6 m/z units) around m/z 221 after CID ofm/z 343 precursor ions. A peak at m/z 223, due toincorporation of 18O, appeared with much higher abundancethan m/z 221 ions for both low-energy beam-type and iontrap CID. Note that if the expected D-Glcp-GA structurewere formed, it should consist of the intact non-reducingsugar unit with a 2-carbon aglycon derived from thereducing sugar (C-2 and C-3 or C-3 and C-4). In this case,18O should not be incorporated into the product ion and itshould still appear at m/z 221. Therefore, the observationof abundant m/z 223 ions indicated that under relativelylow energy CID conditions, most of the m/z 221 ionsformed from α-D-Glcp-(1–3)-D-Glc do not have the D-Glcp-GA structure which is the structural isomer neededto distinguish the stereochemistry and anomeric config-uration of the non-reducing sugar.

Ions at m/z 221 and m/z 223 were further subjected to iontrap CID. Figure 4 compares the CID spectra of the isolatedm/z 221 ions and the m/z 223 ions generated from 18O-labeled α-D-Glcp-(1–3)-D-Glc and β-D-Glcp-(1–3)-D-Glc.The CID spectra of the m/z 221 ions (Figure 4a and c) fromthe two anomeric isomers are distinct from each other andalmost identical to those of the corresponding α and β-D-Glcp-GA standards (Figure 2a and b). The fragmentationpatterns of m/z 223 (Figure 4b and d), however, were similarto each other yet were very different from the syntheticglucosyl-GA standards.

The major fragments that resulted from CID of the m/z223 ions included product ions at m/z 205, 163, 131, and113. These ions are likely due to losses of water (−18 Da), a2-carbon piece, C2H4O2 (−60 Da), a 3-carbon pieceincluding 18O, C3H6O2+

18O (−92 Da), and sequential orconcerted losses of a three-carbon piece plus waterincluding 18O (−110 Da), respectively. Interestingly, neitherthe loss of water nor the loss of 60 Da significantly involvesloss of the 18O oxygen. The m/z 223 ions are hypothesizedto have a structure in which the reducing sugar is connectedto a two-carbon piece from the non-reducing sugar as shownin the scheme above Figure 4b and d. Note that C-1 is nolonger chiral on the piece from the (former) non-reducingsugar, which also explains the similarity in the CID data ofthe m/z 223 ion derived from the two anomeric isomers(Figure 4b and d). We also noticed some subtle differencesbetween Figure 4b and d. For example, the relativeintensities of m/z 205 and 159 are higher (more than 10%)in Figure 4b than in d. These differences may be due to theexistence of a small fraction of structural isomers other thanthat hypothesized for the m/z 223 ions.

The Effect of CID Conditions on the Formationof m/z 221/223 Product Ions from 3-LinkedDisaccharides and their 18O-Labeled Isotopomers

As demonstrated in Figure 4, abundant structural isomers ofm/z 221 product ions, (m/z 223 ions from the 18O-labeled

343100 343223100 (a)

(b)

2.9e5

s

% ps

y, % cp

ity

ty,

ns sit 221

ten

ens

nt te

eI In

ive

0

ati

2830

ela 283

Re

163179179

0100 140 180 220 260 300 340

060

m/z

343223100

4 0 4223

4.0e4

s

% ps

y, % c

sity ty,

ns sit

221ten

ens 221

Int

nte

e I In

ive

0at 0

Rel 179R 179163

100 140 180 220 260 300 340060 100 140 180 220 260 300 340

m/z

060

m/z

Figure 3. MS2 spectra of 18O-labeled α-D-Glcp-(1–3)-D-Glcunder (a) relatively low-energy beam-type CID (CE=6 V), and(b) ion trap CID (AF2=25). Insets in (a) and (b) show isolationof m/z 221 and 223 ions generated from beam-type CID(injection time = 1 s) and ion trap CID (injection time = 2 s),respectively

352 C. Konda et al.: Structure Determination for Disaccharides

140

Page 166: Structural Analysis of Carbohydrates by Mass Spectrometry

disaccharides) were observed under relatively low-energydissociation conditions of 1–3 linked disaccharides, eitherusing beam-type or ion trap CID. This prevents theassignment of the stereochemistry or anomeric configurationof the non-reducing sugar in a typical scenario where either adisaccharide is unlabeled or when it is isolated (unlabeled)from a larger oligosaccharide structure. It would be highlydesirable to optimize CID conditions or, for that matter, tofind any dissociation conditions whereby the relatively pure,structurally informative glycosyl-GA (m/z 221) ions couldbe formed predominantly. Figure 5 shows the effect ofcollision energies on the formation of m/z 221 and 223 ionsunder beam-type and ion-trap CID, using 18O-labeled β-D-Glcp-(1–3)-D-Glc as an example. The data were collectedusing a wide isolation window around m/z 221 to observeboth m/z 221 and 223 ions.

It is clear from Figure 5a–c that the collision energy inbeam-type CID affects the absolute and relative intensities of

m/z 221 ions. When the CE was relatively low (CE=5 V), m/z223 ions were predominantly formed, with four times higherintensity than that of m/z 221. At a higher CE (CE=10 V), m/z221 and 223 ions were seen at nearly equal intensities. Once theCE was increased to 15 V, m/z 221 ions became the dominantpeak, accounting for 80% of the total intensities from m/z 221and 223. Further increasing CE, however, resulted in a hugeloss of ion abundance possibly due to competitive ion ejectionthus the ratio was not improved.

Parameters that might affect the formation of m/z 221 ionsversus m/z 223 ions were also examined for ion trap CID.When ion trap CID of m/z 343 was performed under theinstrument default Mathieu q-parameter (q=0.235), m/z 223ions were formed exclusively independent of activationenergies (data not shown). By changing the activation Mathieuq-parameter to a higher value, precursor ions are placed under ahigher potential well depth, and higher activation energies canbe applied. An AC generated from an external waveform

60 100 140 180 220m/z

0

100

Rel

ativ

e In

ten

sity

, %

131

101

99

22111387159

203

(b)(a)

-H - m/z 223 -H -

α-D-Glcp-(1-3)-Glc, m/z 343 m/z 221 -H

-H -

-α-D-Glcp-(1-3)-Glc, m/z 343

60 100 140 180 220m/z

163

205

113 131

223

0

100

Rel

ativ

e In

ten

sity

, %

159161

1818

(d)

-H - m/z 223 -H -

β-D-Glcp-(1-3)-Glc, m/z 343m/z 221 -H

-H --

β-D-Glcp-(1-3)-Glc, m/z 343

(c)

0

100

Rel

ativ

e In

ten

sity

, %

60 100 140 180 220m/z

0

100R

elat

ive

Inte

nsi

ty, %

163

113 205131

223159

60 100 140 180 220m/z

131

161

22187

203159101

18

18

Figure 4. MS3 spectra of m/z 221 and 223 ions derived from 18O-labeled α-D-Glcp-(1–3)-D-Glc and β-D-Glcp-(1–3)-D-Glc. (a)CID of m/z 221 ions, CE=15 V (MS2), AF2=15 (MS3), (b) CID of m/z 223 ions, CE=5 V (MS2), AF2=14 (MS3) from 18O-labeled α-D-Glcp-(1–3)-D-Glc, and (c) CID of m/z 221 ions, CE=15 V (MS2), AF2=28 (MS3), (d) CID of m/z 223 ions, CE=5 V (MS2), AF2=24 (MS3) from 18O-labeled β-D-Glcp-(1–3)-D-Glc

C. Konda et al.: Structure Determination for Disaccharides 353

141

Page 167: Structural Analysis of Carbohydrates by Mass Spectrometry

generator was used for resonance excitation at q=0.4. Asshown in Figure 5d to f, the ratio ofm/z 221 tom/z 223 ions wasincreased from almost zero to about 1 as the activationamplitude was increased from 100 mVp-p to 400 mVp-p

(activation time: 50 ms for all cases). Further increasing theactivation amplitude resulted in a decrease in m/z 221 to 223ratio as well as a huge ion loss.

The data in Figure 5 suggest that m/z 221 ions, whichhave the desired monosaccharide-GA structures, are gener-ated more favorably under relatively high collision energyconditions in both beam-type and ion trap CID. Comparedwith ion trap CID, beam-type CID provided more abundantand higher relative intensities of the m/z 221 ions that werewanted for discrimination of the stereochemistry andanomeric configuration of the non-reducing sugar. Evidently,a higher activation energy is needed for the formation ofthese m/z 221 product ions, and the pathway to generate theglycosyl-GAs is favored when the internal energies of themolecular ions increase. In beam-type CID, much highercollision energies can be applied (typically more than 10 V)as compared to ion trap CID (hundreds of mV), which leadsto a shift in the internal energy distribution of the molecularions to the high energy direction [38]. It is interesting topoint out that the glycosyl-glycolaldehyde product ions arevirtually the only isomeric species generated under ion trapor low-energy beam-type CID of the 1–2 linked disaccharideanions [34]. Since much higher relative intensities of theglycosyl-glycolaldehyde product ions (10%–40%, normal-ized to the most abundant peak) can be formed from 1–2linkages compared with that of 1–3 linkages (typically G1%

relative intensity) under ion trap CID conditions, it isreasonable to conclude that the formation of these ions from1–2 linkages needs less energy than required for theirgeneration from 1–3 linkages. Therefore, the formation ofthe glycosyl-glycolaldehyde product ions is a much lowerenergy dissociation channel for 1–2 linked disaccharides buta fragmentation pathway for this isomeric species can onlybe promoted for 1–3 linked disaccharides when the collisionenergy is higher.

Given the high pressure in the collision cell (~5 mTorr),multiple collisions happen in beam-type CID, and the first-generation product ions may also be subjected to collisionalactivation once they are formed within the collision cellespecially under higher CE conditions. In this sense, beam-type CID is less selective than ion trap CID, where fragmentions are not typically further activated. Indeed, MS3 CIDstudies in the ion trap showed that many fragment ions,including m/z 325, 323, 283, 281, 253, and 251 generatedm/z 221 ions, which might contribute to the observation ofhigher intensity m/z 221 ions under beam-type CID due tosecondary dissociation.

Identification of the Non-Reducing Sugarand its Anomeric Configuration for 1–3 LinkedDisaccharides

Since relatively pure m/z 221 ions containing the intact non-reducing sugars could be formed using beam-type CID withhigh collision energies, it was possible to differentiate thestereochemistry and anomeric configuration of the non-

223 221 2213.9e5 223 2.5e5221

223 3.9e5 221( ) (b)

2.5e5 223 3.9e5(a) (b) (c)

s

(a) (b) (c)

s s

ps ps

cps

c c

y, c

ty,

ty,

ty,

sit

sit

nsi

ens

221 ens

en

nte 221

nte nte

In 223In In

0 0 0221 223 221 223 221 223

0 0 03

m/z221 223

m/z221 223

m/z

223 223 2239 1e4

2232 0e5

2231 0e5 223

2219.1e4 2.0e5 1.0e5 221(d) (e) (f)(d) (e) (f)

ps

ps

ps

cp cp cp

y,

y,

y,

221sity

sity

sity

221ns

ns

ns

ten

ten

ten

Int

Int

Int

I I

0 0 0221 223 221 223 221 2230 0 0221 223

m/z221 223

m/z221 223

m/zm/z m/z m/z

Figure 5. Isolation of m/z 221 and 223 ions derived from 18O-labeled β-D-Glcp-(1–3)-D-Glc under different collisionalactivation conditions. Beam-type CID: (a) CE=5 V, (b) CE=10 V, (c) CE=15 V. Ion trap CID at q=0.4, f=119.248 kHz, excitationtime = 50 ms: (d) 100 mVp-p, (e) 250 mVp-p, (f) 400 mVp-p

354 C. Konda et al.: Structure Determination for Disaccharides

142

Page 168: Structural Analysis of Carbohydrates by Mass Spectrometry

reducing sugar in disaccharides without 18O-labeling. Figure 6shows the MS3 CID spectra of m/z 221 ions generated bybeam-type CID with relatively high CE (13 to 22 V) from five1–3 linked disaccharides. Each spectrum was an average ofseven spectra and the error bars indicate standard deviations ofthe peak intensities. The standard deviations were found to behigher (0%–12%) for the disaccharide samples than those fromstandards (0%–4%). This larger degree of spectral variation islikely contributed by the fluctuation in ion intensity of the lowabundance m/z 221 isomers under slightly different instrumentconditions, and these isomers fragment differently from thediagnostic and more abundant m/z 221 ions that have themonosaccharide-GA structures.

Note that α-D-Glcp-(1–3)-D-Glc, α-D-Glcp-(1–3)-D-Fru,and β-D-Glcp-(1–3)-Glc are disaccharide isomers containinga glucose as the non-reducing sugar; however, each haseither a different anomeric configuration or reducing sugar.The characteristic fragmentation profile for disaccharideshaving glucose as the non-reducing sugar and an α-anomeric

configuration can be clearly identified for Figure 6a (α-D-Glcp-(1–3)-D-Glc) and Figure 6c (α-D-Glcp-(1–3)-D-Fru),which is distinct from the β-anomeric isomer as shown inFigure 6b (β-D-Glcp-(1–3)-D-Glc, compare all three to thesynthetic standards shown in Figure 2a and b). Figure 6dshows the characteristic fragmentation profile for the m/z221 ion of disaccharides having galactose as the non-reducing sugar and having an α-anomeric configuration(compare to Figure S3a, CID of m/z 221 from α-D-Galp-GAstandard, Supporting Information). The MS3 CID of α-D-Manp-(1–3)-Man (Figure 6e) was similar to that of the α-D-Manp-GA (Supporting Information, Figure S3b). It is alsoimportant to note that under low-energy dissociation condi-tions, the spectra of the m/z 221 product ions derived from thedisaccharides α-D-Glcp-(1–3)-D-Fru, α-D-Galp-(1–3)-D-Galand α-D-Manp-(1–3)-Man did not match those of theirrespective glycosyl-glycolaldehydes (Supporting Informa-tion, Figure S2b and Figure S3c and d). We concludethat this is due to the presence of alternate isomers, pos<

(b)

-H -α-D-Glcp-(1-3)-Glc, m/z 341

m/z 221

β-D-Glcp-(1-3)-Glc, m/z 341-H -

m/z 221

60 100 140 180 220m/z

0

100

Rel

ativ

e In

ten

sity

, %

87

101113

129

131

159

161

203

221

60 100 140 180 220m/z

0

100

Rel

ativ

e In

ten

sity

, %

87

99101

113

129

131

159161

203

221

(a) (c)

-H -α-D-Glcp-(1-3)-Fru, m/z 341

m/z 221

60 100 140 180 220m/z

0

100

Rel

ativ

e In

ten

sity

, %

87

99

101

113

131

129 159

161

203

221

-H -α-D-Galp-(1-3)-Gal, m/z 341-H -

α-D-Manp-(1-3)-Man, m/z 341

m/z 221 m/z 221

(d)

60 100 140 180 220m/z

0

100

Rel

ativ

e In

ten

sity

, %

8799

101

113129131

159

161

203

221

60 100 140 180 220m/z

0

100

Rel

ativ

e In

ten

sity

, %

8799101

113

129

131

159

161

203221

(e)

Figure 6. MS3 CID of m/z 221 ions generated via using high CE (CE=13 to 22 V) for the dissociation of deprotonateddisaccharide ions (m/z 341). (a) α-D-Glcp-(1–3)-D-Glc, CE=15 V (MS2), AF2=26 (MS3), (b) β-D-Glcp-(1–3)-D-Glc, CE=13 V(MS2), AF2=30 (MS3), (c) α-D-Glcp-(1–3)-D-Fru, CE=18 V (MS2), AF2=25 (MS3), (d) α-D-Galp-(1–3)-D-Gal (3α-Gal-Gal),CE=22 V (MS2), AF2=35 (MS3), and (e) α-D-Manp-(1–3)-D-Man (3α-Man-Man), CE=20 V (MS2), AF2=36 (MS3). The error barsin the spectra show the standard deviation of peak intensities based on seven spectra collected over a one year period

C. Konda et al.: Structure Determination for Disaccharides 355

143

Page 169: Structural Analysis of Carbohydrates by Mass Spectrometry

sibly related in their origins to the hypothetical structuresshown in Figure 4b and d but having different reducingmonosaccharides.

The methodology for assigning the stereochemistry andanomeric configuration for the non-reducing sugar unitwithin a disaccharide is based on the comparison of theCID patterns of m/z 221 ions to those of the syntheticmonosaccharide-GA standards [33]. A high similaritybetween the compared spectra indicates a large likelihoodof them sharing the same structure. Spectral similarityscores, which have been widely used in mass spectral librarysearch for both small molecules [39], peptides [40–42], andoligosaccharides [43], were chosen to facilitate thesecomparisons. The spectral similarity scores were calculatedbetween each of the averaged spectra in Figure 6 and theaveraged spectra from the monosaccharide-GA standardsbased on the following equation [44],

Spectral similarity score ¼P

kI1mI2m

� �1=2P kI1mþI2m

2

; and k¼P

I2mPI1m

where I1m and I2m are the normalized intensities of an ion atm/z = m for the two spectra. Note that the spectral similarityscore always has a value between 0 and 1. If two spectra areexactly the same, the spectral similarity score becomes 1. Ingeneral, a large similarity score indicates close similaritybetween the two spectra and a large degree of structuralsimilarity. As shown in Table 1, the spectral similarityscores between a standard and a disaccharide having thesame stereochemistry and anomeric configuration for thenon-reducing side were the highest scores, ranging between0.9838 and 0.9977. When a disaccharide’s stereochemistryand anomeric configuration on the non-reducing side did notmatch with the standard, the spectral similarity score wassignificantly lower, between 0.6942 and 0.9178. Clearly, bycomparing the spectral similarity scores, assigning thestereochemistry and anomeric configuration for the non-reducing side of the 1–3 linked disaccharides could beachieved with high confidence. Note that this was onlypossible under high-energy beam-type CID conditions wherethe m/z 221 product anions containing the intact non-reducing sugars were optimally generated from precursordisaccharides.

ConclusionsCollisional activation of deprotonated 1–3 linked hexose-containing disaccharides (m/z 341) generated a low-abundancem/z 221 product ion. By 18O-labeling the reducing sugarcarbonyl oxygen of these disaccharides, at least twostructural isomers of the m/z 221 ion with the main portionderived from either side of the glycosidic linkage could bemass-discriminated (m/z 221 vs. m/z 223), which enabledthe isomers to be isolated and independently studied. Them/z 221 isomer containing the intact non-reducing sugarattached in glycosidic linkage to a glycolaldehyde aglyconwas found to be analytically useful, since CID of thisspecies provided the structural information that identifiedthe stereochemistry and anomeric configuration of the non-reducing sugar. No structural information could be obtainedfrom m/z 223 isomer(s) to determine the stereochemistry ofthe non-reducing sugar or its anomeric configuration. Theformation of the diagnostic m/z 221 isomer was found tobe affected by CID conditions and was favored underhigher energy beam-type CID. It was demonstrated thatunder optimized CID conditions, this structural isomercould be generated predominantly from five different 1–3linked disaccharides without requiring 18O-labeling of thereducing sugar. Identification of the non-reducing sugar andthe anomeric configuration, therefore, were achieved at ahigh confidence level by statistically comparing the CIDdata of m/z 221 ions generated from the disaccharidesamples with those of the synthetic standards via spectralsimilarity scores. This study also demonstrated that beam-type CID was a more desirable activation method comparedwith ion trap CID to characterize disaccharides using themethodology based on the CID patterns of m/z 221 ions.This method now enables the anomeric configuration andstereochemistry of the m/z 221 ions derived from 1–2, –3, –4,or –6 linked disaccharides to be assigned in the negative ionmode. This capability was afforded due to the specificarrangement of the triple quadrupole-linear ion trap instru-ment. It combined (1) selection of the precursor (m/z 341) inthe first quadrupole with (2) higher energy dissociation in thesecond quadrupole collision cell followed by (3) buildup ofthe desired low abundance m/z 221 product ion in the lineartrap, all three of which were necessary to obtain thesestructural details for 1–3 linked disaccharides.

Table 1. Spectral Similarity Scores for 1–3 Linked Disaccharides Versus Monosaccharide-GA Standards

Disaccharides Synthesized Standards

α-D-Glcp-GA β-D-Glcp-GA α-D-Galp-GA α-D-Manp-GA

α-D-Glcp-(1–3)-Glc 0.9977 0.8845 0.8823 0.7968β-D-Glcp-(1–3)-Glc 0.8572 0.9840 0.7608 0.8568α-D-Glcp-(1–3)-Fru 0.9930 0.8899 0.9027 0.8045α-D-Galp-(1–3)-Gal 0.9178 0.7530 0.9838 0.6942α-D-Manp-(1–3)-Man 0.8461 0.8702 0.7527 0.9891

356 C. Konda et al.: Structure Determination for Disaccharides

144

Page 170: Structural Analysis of Carbohydrates by Mass Spectrometry

AcknowledgmentB.B. acknowledges support from the National ScienceFoundation CHE-0137986.

References1. Taylor, M.E., Drickamer, K.: Introduction to Glycobiology, 2nd edn.

Oxford University Press, Oxford (2006)2. Varki, A., Cummings, R.D., Esko, J.D., Freeze, H.H., Stanley, P.,

Bertozzi, C.R., Hart, G.W., Etzler, M.E.: Essentials of Glycobiology.Cold Spring Harbor Laboratory Press, Cold Spring Harbor (2009)

3. Bush, C.A., Martin-Pastor, M., Imberty, A.: Structure and conformationof complex carbohydrates of glycoproteins, glycolipids, and bacterialpolysaccharides. Annu. Rev. Biophys. Biomolec. Struct. 28, 269–294(1999)

4. Bendiak, B., Fang, T.T., Jones, D.N.: An effective strategy for structuralelucidation of oligosaccharides through NMR spectroscopy combinedwith peracetylation using doubly 13C-labeled acetyl groups. Can. J.Chem. 80, 1032–1050 (2002)

5. Armstrong, G.S., Mandelshtam, V.A., Shaka, A.J., Bendiak, B.: Rapidhigh-resolution four-dimensional NMR spectroscopy using the filterdiagonalization method and its advantages for detailed structuralelucidation of oligosaccharides. J. Magn. Reson. 173, 160–168 (2005)

6. Zaia, J.: Mass spectrometry of oligosaccharides. Mass Spectrom. Rev.23, 161–227 (2004)

7. Fenn, J., Mann, M., Meng, C., Wong, S., Whitehouse, C.: Electrosprayionization for mass spectrometry of large biomolecules. Science 246,64–71 (1989)

8. Reinhold, V.N., Reinhold, B.B., Costello, C.E.: Carbohydrate molecularweight profiling, sequence, linkage, and branching data: ES-MS andCID. Anal. Chem. 67, 1772–1784 (1995)

9. Karas, M., Bachmann, D., Bahr, U., Hillenkamp, F.: Matrix-assistedultraviolet laser desorption of non-volatile compounds. Int. J. MassSpectrom. 78, 53–68 (1987)

10. Tanaka, K., Waki, H., Ido, Y., Akita, S., Yoshida, Y., Yoshida, T.,Matsuo, T.: Protein and polymer analyses up to m/z 100,000 by laserionization time-of-flight mass spectrometry. Rapid Commun. MassSpectrom. 2, 151–153 (1988)

11. Harvey, D.J.: Quantitative aspects of the matrix-assisted laser desorp-tion mass spectrometry of complex oligosaccharides. Rapid Commun.Mass Spectrom. 7, 614–619 (1993)

12. Domon, B., Costello, C.E.: A systematic nomenclature for carbohydratefragmentations in FAB-MS/MS spectra of glycoconjugates. Glycoconj.J. 5, 397–409 (1988)

13. Domon, B., Costello, C.E.: Structure elucidation of glycosphingolipidsand gangliosides using high-performance tandem mass spectrometry.Biochemistry 27, 1534–1543 (1988)

14. Garozzo, D., Giuffrida, M., Impallomeni, G., Ballistreri, A., Montaudo,G.: Determination of linkage position and identification of the reducingend in linear oligosaccharides by negative ion fast atom bombardmentmass spectrometry. Anal. Chem. 62, 279–286 (1990)

15. Dallinga, J.W., Heerma, W.: Reaction mechanism and fragment ionstructure determination of deprotonated small oligosaccharides, studiedby negative ion fast atom bombardment (tandem) mass spectrometry.Biol. Mass Spectrom. 20, 215–231 (1991)

16. Carroll, J.A., Ngoka, L., Beggs, C.G., Lebrilla, C.B.: Liquid secondaryion mass spectrometry/Fourier transform mass spectrometry of oligo-saccharide anions. Anal. Chem. 65, 1582–1587 (1993)

17. Mulroney, B., Traeger, J.C., Stone, B.A.: Determination of both linkageposition and anomeric configuration in underivatized glucopyranosyldisaccharides by electrospray mass spectrometry. J. Mass Spectrom. 30,1277–1283 (1995)

18. Guan, B., Cole, R.B.: MALDI linear-field reflectron TOF post-sourcedecay analysis of underivatized oligosaccharides: Determination ofglycosidic linkages and anomeric configurations using anion attach-ment. J. Am. Soc. Mass Spectrom. 19, 1119–1131 (2008)

19. Jiang, Y., Cole, R.B.: Oligosaccharide analysis using anion attachmentin negative mode electrospray mass spectrometry. J. Am. Soc. MassSpectrom. 16, 60–70 (2005)

20. Firdoussi, A.E., Lafitte, M., Tortajada, J., Kone, O., Salpin, J.-Y.:Characterization of the glycosidic linkage of underivatized disaccharidesby interaction with Pb2+ ions. J. Mass Spectrom. 42, 999–1011 (2007)

21. Staempfli, A., Zhou, Z., Leary, J.A.: Gas-phase dissociation mecha-nisms of dilithiated disaccharides: Tandem mass spectrometry andsemiempirical calculations. J. Org. Chem. 57, 3590–3594 (1992)

22. Laine, R.A., Pamidimukkala, K.M., French, A.D., Hall, R.W., Abbas,S.A., Jain, R.K., Matta, K.L.: Linkage position in oligosaccharides byfast atom bombardment ionization, collision-activated dissociation,tandem mass spectrometry and molecular modeling. L-Fucosylp-(↦1→X)-D-N-acetyl-D-glucosaminylp-(β1→3)-D-galactosylp-(β1-O-methyl) where X=3, 4, or 6. J. Am. Chem. Soc. 110, 6931–6939 (1988)

23. Hofmeister, G.E., Zhou, Z., Leary, J.A.: Linkage position determinationin lithium-cationized disaccharides: tandem mass spectrometry andsemiempirical calculations. J. Am. Chem. Soc. 113, 5964–5970 (1991)

24. Asam, M.R., Glish, G.L.: Tandem mass spectrometry of alkalicationized polysaccharides in a quadrupole ion trap. J. Am. Soc. MassSpectrom. 8, 987–995 (1997)

25. Polfer, N.C., Valle, J.J., Moore, D.T., Oomens, J., Eyler, J.R., Bendiak,B.: Differentiation of isomers by wavelength-tunable infrared multiple-photon dissociation-mass spectrometry: Application to glucose-contain-ing disaccharides. Anal. Chem. 78, 670–679 (2005)

26. Simoes, J., Domingues, P., Reis, A., Nunes, F.M., Coimbra, M.A.,Domingues, M.R.: Identification of anomeric configuration of under-ivatized reducing glucopyranosyl-glucose disaccharides by tandemmass spectrometry and multivariate analysis. Anal. Chem. 79, 5896–5905 (2007)

27. Zhang, H., Brokman, S.M., Fang, N., Pohl, N.L., Yeung, E.S.: Linkageposition and residue identification of disaccharides by tandem massspectrometry and linear discriminant analysis. Rapid Commun. MassSpectrom. 22, 1579–1586 (2008)

28. Mendonca, S., Cole, R., Zhu, J., Cai, Y., French, A., Johnson, G.,Laine, R.: Incremented alkyl derivatives enhance collision inducedglycosidic bond cleavage in mass spectrometry of disaccharides. J. Am.Soc. Mass Spectrom. 14, 63–78 (2003)

29. Gaucher, S.P., Leary, J.A.: Stereochemical differentiation of mannose,glucose, galactose, and talose using zinc(II) diethylenetriamine and ESI-ion trap mass spectrometry. Anal. Chem. 70, 3009–3014 (1998)

30. Desaire, H., Leary, J.A.: Differentiation of diastereomeric N-acetylhex-osamine monosaccharides using ion trap tandem mass spectrometry.Anal. Chem. 71, 1997–2002 (1999)

31. Mueller, D.R., Domon, B.M., Blum, W., Raschdorf, F., Richter, W.J.:Direct stereochemical assignment of sugar subunits in naturallyoccurring glycosides by low energy collision induced dissociation.Application to papulacandin antibiotics. Biol. Mass Spectrom. 15, 441–446 (1988)

32. Domon, B., Muller, D.R., Richter, W.J.: Determination of interglyco-sidic linkages in disaccharides by high performance tandem massspectrometry. Int. J. Mass Spectrom. Ion Process. 100, 301–311 (1990)

33. Fang, T.T., Bendiak, B.: The stereochemical dependence of unim-olecular dissociation of monosaccharide-glycolaldehyde anions in thegas phase: A basis for assignment of the stereochemistry and anomericconfiguration of monosaccharides in oligosaccharides by mass spec-trometry via a key discriminatory product ion of disaccharidefragmentation, m/z 221. J. Am. Chem. Soc. 129, 9721–9736 (2007)

34. Fang, T.T., Zirrolli, J., Bendiak, B.: Differentiation of the anomericconfiguration and ring form of glucosyl-glycolaldehyde anions in thegas phase by mass spectrometry: isomeric discrimination between m/z221 anions derived from disaccharides and chemical synthesis of m/z221 standards. Carbohydr. Res. 342, 217–235 (2007)

35. Brown, D.J., Stefan, S.E., Berden, G., Steill, J.D., Oomens, J., Eyler, J.R., Bendiak, B.: Direct evidence for the ring opening of monosaccha-ride anions in the gas phase: photodissociation of aldohexoses andaldohexoses derived from disaccharides using variable-wavelengthinfrared irradiation in the carbonyl stretch region. Carbohydr. Res.346, 2469–2481 (2011)

36. Carroll, J.A., Ngoka, L., Beggs, C.G., Lebrilla, C.B.: Liquid secondaryion mass spectrometry/Fourier transform mass spectrometry of oligo-saccharide anions. Anal. Chem. 65, 1582–1587 (1993)

37. Bendiak, B., Fang, T.T.: Assignment of the stereochemistry andanomeric configuration of structurally informative product ions derivedfrom disaccharides: infrared photodissociation of glycosyl-glycolalde-hydes in the negative ion mode. Carbohydr. Res. 345, 2390–2400(2010)

38. Wells, J.M., McLuckey, S.A. Collision-Induced Dissociation (CID) ofPeptides and Proteins. In: Biol. Mass Spectrom. Burlingame, A.L. (ed.)Academic Press, Amsterdam, pp. 148–185 (2005)

C. Konda et al.: Structure Determination for Disaccharides 357

145

Page 171: Structural Analysis of Carbohydrates by Mass Spectrometry

39. Stein, S., Scott, D.: Optimization and testing of mass spectral librarysearch algorithms for compound identification. J. Am. Soc. MassSpectrom. 5, 859–866 (1994)

40. Frewen, B.E., Merrihew, G.E., Wu, C.C., Noble, W.S., MacCoss, M.J.:Analysis of peptide MS/MS spectra from large-scale proteomics experi-ments using spectrum libraries. Anal. Chem. 78, 5678–5684 (2006)

41. Lam, H., Deutsch, E.W., Eddes, J.S., Eng, J.K., King, N., Stein, S.E.,Aebersold, R.: Development and validation of a spectral librarysearching method for peptide identification from MS/MS. Proteomics7, 655–667 (2007)

42. Lam, H., Deutsch, E.W., Eddes, J.S., Eng, J.K., Stein, S.E., Aebersold,R.: Building consensus spectral libraries for peptide identification inproteomics. Nat. Methods 5, 873–875 (2008)

43. Kameyama, A., Kikuchi, N., Nakaya, S., Ito, H., Sato, T.,Shikanai, T., Takahashi, Y., Takahashi, K., Narimatsu, H.: Astrategy for identification of oligosaccharide structures usingobservational multistage mass spectral library. Anal. Chem. 77,4719–4725 (2005)

44. Zhang, Z.: Prediction of low-energy collision-induced dissociationspectra of peptides. Anal. Chem. 76, 3908–3922 (2004)

358 C. Konda et al.: Structure Determination for Disaccharides

146