09.02.2013 Views

Electrospray ionization source geometry for mass spectrometry: past ...

Electrospray ionization source geometry for mass spectrometry: past ...

Electrospray ionization source geometry for mass spectrometry: past ...

SHOW MORE
SHOW LESS

Create successful ePaper yourself

Turn your PDF publications into a flip-book with our unique Google optimized e-Paper software.

Trends in Analytical Chemistry, Vol. 25, No. 3, 2006 Trends<br />

<strong>Electrospray</strong> <strong>ionization</strong> <strong>source</strong><br />

<strong>geometry</strong> <strong>for</strong> <strong>mass</strong> <strong>spectrometry</strong>:<br />

<strong>past</strong>, present, and future<br />

Irina Manisali, David D.Y. Chen, Bradley B. Schneider<br />

The <strong>geometry</strong> of an electrospray ion <strong>source</strong> plays important roles in the<br />

processes of analyte desolvation, <strong>ionization</strong>, transportation, and detection in<br />

a <strong>mass</strong> spectrometer. We provide a brief account of the scientific principles<br />

involved in developing an electrospray ion <strong>source</strong>, and in the various geometries<br />

used to improve the sensitivity of <strong>mass</strong> <strong>spectrometry</strong>. We also present<br />

some popular configurations currently available and outline future<br />

trends in this research area.<br />

ª 2005 Elsevier Ltd. All rights reserved.<br />

Keywords: Atmospheric pressure <strong>ionization</strong>; Commercial ion <strong>source</strong> design; <strong>Electrospray</strong><br />

<strong>ionization</strong>; Ion <strong>source</strong>; Ion <strong>source</strong> <strong>geometry</strong>; Mass <strong>spectrometry</strong><br />

Irina Manisali,<br />

David D.Y. Chen*<br />

Department of Chemistry,<br />

University of British Columbia,<br />

2036 Main Mall, Vancouver,<br />

BC, Canada V6T 1Z1<br />

Bradley B. Schneider<br />

MDS SCIEX, 71 Four Valley Dr.,<br />

Concord, ON, Canada L4K 4V8<br />

* Corresponding author.<br />

Tel.: +1 604 822 0887;<br />

Fax: +1 604 822 2847;<br />

E-mail: chen@chem.ubc.ca<br />

1. Brief introduction to electrospray<br />

<strong>Electrospray</strong> <strong>ionization</strong> (ESI) is a soft <strong>ionization</strong><br />

method, allowing the <strong>for</strong>mation of<br />

gas phase ions through a gentle process<br />

that makes possible the sensitive analysis<br />

of non-volatile and thermolabile compounds.<br />

Consequently, the use of the ESI<br />

<strong>source</strong> in the field of <strong>mass</strong> <strong>spectrometry</strong><br />

(MS) has greatly facilitated the study of<br />

large biomolecules, as well as pharmaceutical<br />

drugs and their metabolites. Thus,<br />

the ESI <strong>source</strong> has evolved with the<br />

growth of proteomics and drug discovery<br />

research [1]. The analysis of carbohydrates,<br />

nucleotides, and small polar molecules<br />

comprises a few of the many other<br />

applications in which ESI-MS is currently<br />

used. Together with matrix assisted laser<br />

desorption <strong>ionization</strong> (MALDI), another<br />

soft <strong>ionization</strong> technique, ESI has revolutionized<br />

biomedical analysis. In 2002, half<br />

the Nobel Prize <strong>for</strong> Chemistry was awarded<br />

to ESI developer John Fenn, who<br />

shared it with Koichi Tanaka, who made<br />

an outstanding contribution to the development<br />

of MALDI.<br />

The concept of the ESI <strong>source</strong> is<br />

deceivably simple, as some aspects of its<br />

functioning are still not well understood.<br />

The technique involves a number of steps,<br />

including the <strong>for</strong>mation of charged droplets,<br />

desolvation, ion generation, declustering,<br />

and ion sampling. As the name<br />

suggests, the basis of this technique lies in<br />

using a strong electric field to create an<br />

excess of charge at the tip of a capillary<br />

containing the analyte solution. Charged<br />

droplets exit the capillary as a spray and<br />

travel at atmospheric pressure down an<br />

electrical gradient to the gas conductancelimiting<br />

orifice or tube. Gas phase ions are<br />

then transported through different vacuum<br />

stages to the <strong>mass</strong> analyzer and<br />

ultimately the detector. ESI has successfully<br />

been coupled with a variety of <strong>mass</strong><br />

analyzers. Each analyzer has different<br />

advantages and the choice of which to use<br />

is based on the requirements of the application.<br />

Comprehensive in<strong>for</strong>mation about<br />

the coupling of ESI with various <strong>mass</strong><br />

analyzers can be found in Cole’s 1997<br />

compilation of review articles on the topic<br />

[2].<br />

2. Milestones in the evolution of<br />

electrospray<br />

Since 1968, when Dole used ESI and a<br />

nozzle-skimmer system to produce<br />

charged gas-phase polystyrene [3], the ESI<br />

<strong>source</strong> has continued to undergo various<br />

changes in its size, material, and <strong>geometry</strong>.<br />

These trans<strong>for</strong>mations were made to<br />

optimize both the <strong>ionization</strong> efficiency and<br />

the efficiency of transferring gas phase<br />

ions into the <strong>mass</strong> analyzer.<br />

The evolution of the electrospray-<strong>source</strong><br />

design has also reflected the coupling of<br />

the ion <strong>source</strong> to separation systems, such<br />

0165-9936/$ - see front matter ª 2005 Elsevier Ltd. All rights reserved. doi:10.1016/j.trac.2005.07.007 243


Trends Trends in Analytical Chemistry, Vol. 25, No. 3, 2006<br />

as high per<strong>for</strong>mance liquid chromatography (HPLC) and<br />

capillary electrophoresis (CE). This has brought new<br />

challenges in generating ions at different flow rates,<br />

minimizing contamination of the interface, and optimizing<br />

overall efficiency. Leading instrument manufacturers<br />

are directly involved in providing innovative<br />

solutions to the industry’s demand <strong>for</strong> more sensitive,<br />

reliable <strong>mass</strong> spectrometers.<br />

The purpose of this review is to present some milestones<br />

in the evolution of the ESI <strong>source</strong> to achieve the<br />

designs and the capabilities currently offered by commercial<br />

manufacturers. We discuss some important<br />

developments achieved by research groups with respect<br />

to ESI design. We also present a basic theoretical background<br />

on droplet <strong>for</strong>mation and current theories of gas<br />

phase ion creation in order to provide the reader with a<br />

better understanding of the reasoning behind the modifications<br />

in <strong>source</strong> design. While other atmospheric<br />

pressure ion <strong>source</strong>s use similar geometries, this review<br />

will focus on only electrospray <strong>ionization</strong>.<br />

3. Theory<br />

A number of publications have presented details of the<br />

electrochemical nature of the electrospray process [4–6].<br />

When a large potential difference is applied between an<br />

electrode shaped as a wire and a counter electrode, a<br />

strong electric field is created at the tip. In the case of ESI,<br />

a high voltage is applied to a capillary containing the<br />

analyte solution. For simplicity, we consider the application<br />

of positive potential only. Due to the electric field<br />

gradient at the tip, charge separation occurs in the<br />

solution as anions migrate towards the capillary walls,<br />

and cations travel towards the meniscus of the droplet<br />

<strong>for</strong>med at the tip (Fig. 1) [7].<br />

In order to direct charged species into the <strong>mass</strong> spectrometer,<br />

a series of counter electrodes is used in order of<br />

decreasing potential. Typically, the principal counter<br />

sample<br />

solution<br />

HV<br />

_ _ _ _ _ _ _ _ _<br />

+<br />

+<br />

+ +<br />

_ _ _ _ _ _ _ _ _ + +<br />

+<br />

+ +<br />

+ + + +<br />

_ _<br />

_ +<br />

_ + _+<br />

+ +<br />

+ _ +<br />

+<br />

_<br />

Taylor<br />

cone<br />

+ +<br />

E<br />

evaporation<br />

+ +<br />

+ +<br />

+ +<br />

+<br />

+<br />

+<br />

_<br />

_<br />

_<br />

offspring<br />

droplets<br />

ions<br />

and<br />

neutrals<br />

Figure 1. Schematic diagram of the electrospray process.<br />

244 http://www.elsevier.com/locate/trac<br />

counter<br />

electrode<br />

electrode can be either the curtain plate of the <strong>mass</strong><br />

spectrometer or a transport capillary, which will be<br />

discussed later. The optimal potential difference between<br />

the sprayer and the principal counter electrode depends<br />

on experimental parameters, such as the charge state of<br />

the analyte, the solution flow rate, the solvent composition,<br />

and the distance between the tip and the counter<br />

electrode. While different <strong>mass</strong> spectrometers may require<br />

different applied voltages on the sprayer and the<br />

counter electrode, the potential difference between them<br />

is similar (typically 2–5 kV) [8]. In the presence of an<br />

electric field, liquid emerges from the tip of the capillary<br />

in the shape of a cone, also known as the ‘‘Taylor cone’’<br />

(Fig. 1) [9]. When the electrostatic repulsion between<br />

charged molecules at the surface of the Taylor cone<br />

approaches the surface tension of the solution – known<br />

as reaching the Rayleigh limit – charged droplets are<br />

expelled from the tip. The droplets containing excess<br />

positive charge generally follow the electric field lines at<br />

atmospheric pressure toward the counter electrode.<br />

However, trajectories will also be affected by space<br />

charge and gas flow.<br />

The mechanism of <strong>for</strong>ming the Taylor cone is not<br />

entirely understood, but it is known that, under certain<br />

conditions, the morphology of the spray emitted from the<br />

capillary tip can change [10]. The various spray modes<br />

strongly depend on the capillary voltage and are related<br />

to pulsation phenomena observed in the capillary current<br />

[10]. Juraschek and Röllgen showed that liquid flow<br />

rate, capillary diameter, and electrolyte concentration<br />

can all impact the spray mode. Controlling the spray<br />

mode is thus crucial in achieving a stable spray and<br />

an optimal signal. However, this becomes particularly<br />

difficult when the mobile phase composition changes<br />

during gradient elution conditions necessary in many<br />

LC-MS applications. To address this problem, Valaskovic<br />

and Murphy have developed an orthogonal optoelectronic<br />

system capable of identifying many spray modes<br />

under different conditions [11]. The system automatically<br />

optimizes the ESI potential in response to changes<br />

in flow rate and solvent composition.<br />

The size of the spray droplets released from the Taylor<br />

cone, highly dependent on flow rate and capillary<br />

diameter, is critical to the efficient <strong>ionization</strong> of the<br />

analyte.<br />

Since a small droplet contains less solvent, desolvation<br />

and <strong>ionization</strong> can be more efficient. Because less fission<br />

is required to produce ions, the salt concentration in the<br />

final offspring droplets may be lower compared to a<br />

droplet that has undergone more evaporation-fission<br />

cycles. As a result, the background noise in the <strong>mass</strong><br />

spectrum may be reduced [12]. In addition, with smaller<br />

droplets, analytes that are not surface active will have a<br />

greater chance of being transferred to the gas phase<br />

rather than being lost in the bulk of the parent droplet<br />

residue [12].


Trends in Analytical Chemistry, Vol. 25, No. 3, 2006 Trends<br />

The solvent is typically a combination of acidified<br />

water and an organic modifier. The role of the organic<br />

solvent is to lower the surface tension of the liquid,<br />

facilitating the <strong>for</strong>mation of gas phase ions. As the solvent<br />

evaporates, the droplet shrinks and the electrostatic<br />

repulsion between charges within the droplet increases.<br />

When the Rayleigh limit is approached, offspring droplets<br />

begin to break away unevenly in a process also<br />

known as ‘‘coulombic fission’’ [7]. Evaporation of the<br />

offspring droplets leads to a new fission series and the<br />

process repeats itself to produce smaller and smaller<br />

droplets. The eventual creation of gas phase ions from<br />

these droplets is thought to be a combination of two<br />

mechanisms known as the ‘‘ion evaporation mechanism’’<br />

(IEM), initially proposed by Iribarne and Thomson<br />

[13], and the ‘‘charged residue model’’ (CRM), put <strong>for</strong>ward<br />

by Dole et al. [3,14] and supported by Rollgen et al.<br />

[15]. The fundamental difference between the two lies in<br />

the mechanism of <strong>for</strong>ming gas phase ions.<br />

IEM suggests that, when the electric field on a charged<br />

droplet is high enough, single, solvated, analyte molecules<br />

carrying some of the droplet charge are ejected into<br />

the gas phase. This happens because the potential energy<br />

of the ions near the surface becomes high enough to<br />

allow evaporation to occur [16].<br />

By contrast, the CRM maintains that gas phase ions<br />

are <strong>for</strong>med when successive fissions lead to a charged<br />

droplet containing a single analyte.<br />

Although it is difficult to determine which of the two<br />

mechanisms is more accurate, extensive studies [17–19]<br />

seem to suggest that CRM is the preferred mechanism in<br />

the case of <strong>for</strong>ming charged globular proteins in the gas<br />

phase.<br />

Kebarle also concluded that charging a single protein in<br />

the evaporating droplet is due to small ions found at the<br />

surface of the droplet [18]. The mechanism of <strong>for</strong>ming<br />

small analyte ions is still not clear. Charging analyte<br />

molecules can occur through more than one process [20].<br />

Charge separation (in the ESI <strong>source</strong>), adduct <strong>for</strong>mation,<br />

gas phase molecular reactions, and electrochemical<br />

reactions may also contribute to <strong>ionization</strong> during the<br />

electrospray process.<br />

Efficient transport of ions and charged droplets from<br />

the sprayer into the <strong>mass</strong> spectrometer is challenging<br />

and depends on parameters such as interface arrangement<br />

and gas throughput into the instrument [8]. As gas<br />

and ions are transported from atmospheric pressure into<br />

vacuum, strong cooling of the mixture occurs during<br />

expansion. Under these conditions, polar neutral molecules<br />

may cluster with analyte ions and it is there<strong>for</strong>e<br />

very important to achieve efficient desolvation within<br />

the atmospheric pressure region. One instrument manufacturer<br />

(MDS SCIEX) introduced a method of desolvation<br />

(described in more detail in Section 5.1.2.),<br />

whereby heated auxiliary nitrogen gas is directed at an<br />

angle from the direction of the spray (TurboIonSpraye)<br />

[21,22]. In addition, counter current nitrogen gas flow<br />

(Curtain Gase, further illustrated in Fig. 5) emanating<br />

from in front of the gas conductance limiting orifice<br />

provided additional desolvation [8]. The advantage of<br />

the Curtain Gas is that neutral molecules, such as solvent,<br />

are carried away from the sampling orifice by the<br />

nitrogen flow, while charged molecules are directed<br />

through the gas flow under the influence of the electric<br />

field. Counter current gas flows can also aid in declustering<br />

as a result of collisions between ions and gas<br />

molecules. Since nitrogen is inert, it cannot <strong>for</strong>m covalent<br />

bonds with the ions and clustering during transfer<br />

to vacuum is prevented. While its name varies depending<br />

on the manufacturer, counter current gas flow is an<br />

important feature of the ESI ion <strong>source</strong>. In addition to<br />

using inert gas flow, heating the ion <strong>source</strong> or gas conductance<br />

limiting orifice or capillary is also a common<br />

way to aid declustering and desolvation [23].<br />

4. Historical highlights<br />

4.1. Development of the traditional ESI <strong>source</strong><br />

Although the work of Fenn and colleagues demonstrated<br />

the applicability of ESI to generation of biomolecular<br />

ions, it was Dole and co-workers who led the way in the<br />

late 1960s by building the first electrospray based <strong>mass</strong><br />

analysis system [3], as shown in Fig. 2.<br />

A sharp hypodermic needle was used to spray polystyrene<br />

solution into an evaporation chamber. A Teflon<br />

plate supported the needle as well as a gas inlet introducing<br />

nitrogen as a bath gas. A nozzle-skimmer system<br />

and two stage differential pumping were used to reduce<br />

pressure in the analyzer region and permit sampling of<br />

the supersonic free jet without disturbing the molecular<br />

flow [24]. The voltage applied on the needle was maintained<br />

at approximately 10 kV (negative polarity<br />

used), while the voltages on the first and second collimating<br />

plates were 3 kV, and 1.4 kV, respectively.<br />

Dole measured <strong>mass</strong>-to-charge ratios (m/z) by determining<br />

the retarding potential of ions through a grid<br />

preceding a Faraday cup detector.<br />

Intrigued by Dole’s ESI work, and having himself been<br />

a pioneer of supersonic free jets as molecular beam<br />

<strong>source</strong>s, Fenn reproduced Dole’s experiments with the<br />

same observations but discontinued research in this area<br />

due to technical limitations [25]. Years later, Fenn<br />

reprised his experiments on molecular beams and used<br />

ESI coupled to a quadrupole <strong>mass</strong> analyzer <strong>for</strong> ion<br />

detection. Recognizing flaws in Dole’s ion <strong>source</strong> design<br />

as well as in the interpretation of results, Fenn modified<br />

the electrospray ion <strong>source</strong> by reducing the distance<br />

between the hypodermic needle and end plate containing<br />

the nozzle [26] (Fig. 3). As a result, the applied<br />

sprayer voltage could be significantly lower.<br />

http://www.elsevier.com/locate/trac 245


Trends Trends in Analytical Chemistry, Vol. 25, No. 3, 2006<br />

Faraday Cage<br />

Skimmer<br />

Nozzle<br />

Pump Pump<br />

N 2 gas out<br />

In contrast with Dole’s ion <strong>source</strong>, Fenn’s ESI chamber<br />

was built with metal walls to prevent charge build up,<br />

and had significantly smaller dimensions. Fenn’s ESI<br />

<strong>source</strong> later evolved into what is known as the Fenn-<br />

Whitehouse design [27]. The additional characteristic is<br />

a glass capillary that allows ions from atmospheric<br />

pressure to enter the first vacuum stage of the instrument.<br />

The capillary dimensions can be chosen to allow<br />

the same flux of drying gas and ions as a regular thinplate<br />

orifice [24]. Increasing the capillary diameter may<br />

be desirable in order to increase the acceptance and to<br />

decrease the chance of blockage. However, <strong>for</strong> a fixed gas<br />

throughput, this must be associated with a corresponding<br />

increase in capillary length, potentially lowering<br />

per<strong>for</strong>mance.<br />

An alternative to using the counter current gas <strong>for</strong> ion<br />

desolvation is the heated metal capillary, introduced by<br />

Chowdhury et al. [23]. With this configuration, gas<br />

throughput depends on both capillary dimensions and<br />

temperature. The heated metal capillary inlet (ion pipe), as<br />

Spray Chamber<br />

N 2 gas in<br />

Liquid in<br />

Figure 2. Configuration of Dole’s first electrospray based <strong>mass</strong> analysis system. (Reprinted with permission from [3], ª1968, American Institute<br />

of Physics).<br />

Oxygen<br />

Liquid<br />

Sample<br />

Cylindrical Electrode<br />

Hypodermic Needle<br />

shown in Fig. 4, is currently used on Thermo Electron<br />

instruments. However, a number of other manufacturers<br />

have developed modified versions of this device, including<br />

heated resistive tubes [28] and heated metal capillaries<br />

with different inlet/outlet dimensions [29]. The combination<br />

of a heated metal capillary and the counter current<br />

gas makes possible even more efficient ion desolvation.<br />

ESI-MS interfaces comprising multiple capillaries have<br />

been developed <strong>for</strong> various applications, such as ion ion<br />

reaction monitoring [30,31], separate introduction of<br />

analyte and calibrant using a multiple sprayer <strong>source</strong><br />

[32], and improved ion transmission when used in<br />

combination with an electrodynamic ion funnel [33].<br />

Other modifications to the heated capillary inlet include<br />

offsetting the exit of the capillary from downstream<br />

ion optics devices (such as a skimmer) [34,35]. The<br />

advantage of this design is that some large solvent<br />

clusters exiting the transfer capillary are prevented from<br />

entering the next vacuum stage, while the lighter analyte<br />

ions follow the gas flow through the ion optics.<br />

N 2<br />

End Plate<br />

Nozzle<br />

Skimmer<br />

Quadrupole<br />

Figure 3. Configuration of the electrospray ion <strong>source</strong> coupled with a quadrupole <strong>mass</strong> analyzer designed by Fenn and co-workers in 1984.<br />

(Reprinted with permission from [22], ª1984, American Chemical Society).<br />

246 http://www.elsevier.com/locate/trac


Trends in Analytical Chemistry, Vol. 25, No. 3, 2006 Trends<br />

ESI<br />

Needle<br />

ESI Nozzle<br />

Ion Sweep<br />

Cone<br />

Recently, Varian also adopted this modification by<br />

positioning a heated capillary outlet off axis from a<br />

subsequent RF multipole ion guide [36].<br />

In 1987, Bruins et al. introduced the pneumatically<br />

assisted nebulizer <strong>for</strong> the electrospray interface, also<br />

known as ion spray. Fig. 5 shows the details of the ion<br />

spray interface [37].<br />

The nebulizer gas was beneficial <strong>for</strong> coupling LC with<br />

ESI-MS, because it helped to stabilize electrospray <strong>for</strong><br />

flow rates up to approximately 0.2 mL/min. It also tolerated<br />

larger distances between the sprayer and the<br />

counter electrode, reducing the occurrence of corona<br />

Heater<br />

Ion<br />

Transfer<br />

Capillary<br />

To<br />

Skimmer<br />

Figure 4. Schematic of the Ion Max <strong>source</strong> from Thermo Electron using a heated metal capillary <strong>for</strong> ion desolvation and transfer.<br />

silica capillary<br />

containing<br />

sample<br />

solution<br />

stainless steel<br />

capillary holder<br />

sample solution<br />

stainless steel<br />

tubing<br />

discharge [37]. Bruins observed that the spraying<br />

process was less dependent on the sprayer position relative<br />

to the orifice than without nebulizer assistance, and<br />

that better sensitivity was obtained if the sprayer was<br />

pointed off axis instead of directly at the orifice. The<br />

reasoning behind the off axis <strong>geometry</strong> was that, by<br />

sampling the periphery of the spray, finer droplets entered<br />

the <strong>mass</strong> spectrometer while the larger droplets struck the<br />

curtain plate. This led to improved per<strong>for</strong>mance because<br />

finer droplets are easier to desolvate. Currently, the off<br />

axis sprayer <strong>geometry</strong> has evolved to an orthogonal<br />

sampling position. More detail is given in Section 5 below.<br />

Air / N2<br />

Air / N2<br />

E<br />

neutrals<br />

charged<br />

droplets<br />

ions<br />

atmospheric pressure<br />

Figure 5. Schematic of the ion spray configuration developed by Bruins in 1987.<br />

N 2<br />

N<br />

2<br />

vacuum<br />

http://www.elsevier.com/locate/trac 247


Trends Trends in Analytical Chemistry, Vol. 25, No. 3, 2006<br />

4.2. The CE-MS interface<br />

While ion spray became standard <strong>for</strong> LC-MS at higher<br />

flows, CE-MS coupling was more complicated due to the<br />

large variety of solvent compositions required <strong>for</strong> separation<br />

and problems associated with electrochemistry. In<br />

addition, differences in flow rate requirements also<br />

presented problems. Three common ways of interfacing<br />

CE-MS are shown in Fig. 6.<br />

The most common method of CE-MS interfacing is the<br />

coaxial sheath flow interface, first described by Smith<br />

et al. in 1988 [38]. In 2003, approximately 77% of CE-<br />

MS users employed this type of interface, due to its<br />

reproducibility and robustness [39]. By allowing a conducting<br />

liquid to provide electrical contact between the<br />

stainless steel capillary maintained at high voltage and<br />

the fused silica capillary carrying the analyte solution,<br />

the interfacing of capillary zone electrophoresis (CZE)<br />

with ESI-MS could be achieved with many buffer systems<br />

that were previously impractical. This is significant because<br />

aqueous buffers of high ionic strength are normally<br />

not tolerated by ESI-MS. Buffer ions can lead to<br />

charge neutralization and corona discharge as well as<br />

<strong>for</strong>mation of undesirable adducts with the analyte<br />

resulting in loss of signal and high sensitivity and high<br />

background. By using organic solvents, such as acetonitrile<br />

or methanol, as the sheath liquid, buffers of ionic<br />

strength up to 0.2 M can be used. The sheath liquid<br />

prevents direct contact between the high voltage elec-<br />

a nebulizer gas<br />

sheath liquid<br />

stainless steel<br />

capillary/electrode <strong>for</strong><br />

CZE / ESI voltage<br />

application<br />

b<br />

nebulizer gas<br />

c<br />

H.V.<br />

nebulizer gas tube<br />

analyte/buffer solution<br />

analyte/buffer solution<br />

separation capillary<br />

separation capillary with conductive<br />

coating <strong>for</strong> ESI voltage application<br />

liquid junction<br />

nebulizer gas<br />

to MS<br />

analyte/buffer solution<br />

separation capillary<br />

to MS<br />

to MS<br />

Figure 6. Schematic of the three most common ways to interface<br />

CE-MS: a) coaxial sheath flow; b) sheathless; and, c) liquid junction.<br />

248 http://www.elsevier.com/locate/trac<br />

trode and the analyte solution, thus avoiding electrochemical<br />

modification of analytes. An additional<br />

advantage of using the sheath flow interface is minimization<br />

of corona discharge as a result of reducing the<br />

ionic strength of the sprayed solvent [38].<br />

Olivares et al. developed the sheathless interface in<br />

1987 by using a stainless steel capillary to create electrical<br />

contact with the analyte solution [39]. Accounting<br />

<strong>for</strong> about 11% of CE-MS interfaces in 2003 [40], this<br />

method evolved with a focus on low flow rates <strong>for</strong> improved<br />

sensitivity.<br />

Other sheathless methods were subsequently developed,<br />

such as sharpening the CE-capillary tip in order to<br />

produce a stable spray, while the ESI electrode was made<br />

by installing a piece of gold wire [41] or coating the<br />

capillary end with a metal or alloy [42]. The advantages<br />

of the sheathless interface are that analyte dilution and<br />

ion suppression due to the sheath flow can be eliminated.<br />

Sensitivity can also be improved by up to an order of<br />

magnitude over sheath-flow designs [43].<br />

A third method of interfacing CE-MS was the liquidjunction<br />

interface developed by Henion and colleagues<br />

[44,45]. A liquid junction between the CE capillary and<br />

the ESI emitter was built using a stainless steel tee<br />

equipped with a buffer reservoir. This interface reduced<br />

the complication brought about by the different flow rate<br />

requirements of CE and ESI [46], but peak broadening<br />

and mechanical difficulties limited the general applicability<br />

of this technique [42].<br />

4.3. Miniaturization of electrospray<br />

In an attempt to reduce flow rates and thus produce<br />

smaller droplets to improve the <strong>ionization</strong> efficiency,<br />

various miniaturized ESI <strong>source</strong>s were developed [47].<br />

Caprioli et al. [48,49] developed an on line ESI <strong>source</strong><br />

equipped with a sprayer tip of 10–20 lm internal<br />

diameter and capable of operating at flow rates of 300–<br />

800 nL/min. The emitter was created by burning off the<br />

polyimide coating of a fused silica capillary tip and<br />

tapering it by etching with hydrofluoric acid to give<br />

droplets in the micron-diameter range. Caprioli named<br />

his technique microelectrospray, a term that currently<br />

describes a flow rate range of 0.2–4 lL/min [50].<br />

Operation in this flow-rate regime may involve various<br />

means <strong>for</strong> pulling silica capillaries and applying the<br />

electrospray voltage [51].<br />

In 1995, Caprioli introduced the idea of using the LCcolumn<br />

tip combined with microelectrospray technology<br />

[52] to spray directly towards the <strong>mass</strong>-spectrometer<br />

inlet. This was significant because it triggered the evolution<br />

of analytical LC columns towards miniaturized<br />

versions of microbore and nanobore capillary columns<br />

necessary <strong>for</strong> low flow rate electrospray [50].<br />

Nanoelectrospray, introduced by Wilm and Mann<br />

more than a decade ago [53], was a particularly<br />

successful technique due to its ability to <strong>for</strong>m droplets


Trends in Analytical Chemistry, Vol. 25, No. 3, 2006 Trends<br />

100–1000 times smaller in diameter than droplets<br />

<strong>for</strong>med by conventional ESI. The technique was named<br />

to reflect the nanometer (nm) sized droplets as well as<br />

the flow rate of nanoliters per minute (nL/min) [54].<br />

Essentially, the solution infusion system was eliminated,<br />

as a few microliters of solution were transferred to the tip<br />

of a gold coated glass capillary. The emitter tip was<br />

located within a few mm of the inlet aperture, and a<br />

stereomicroscope was used to accurately position the<br />

sprayer in front of the inlet. By applying a low potential<br />

onto the conductive capillary surface, a fine spray was<br />

generated and the charged droplets were directed towards<br />

the counter electrode under the effect of the<br />

electric field gradient. Despite an estimated efficiency of<br />

500 times greater than traditional ESI [54], nanoelectrospray<br />

has some limitations, including plugging of the<br />

emitter tip and lower reproducibility associated with<br />

differences in tip morphology.<br />

While the term ‘‘nanoelectrospray’’ initially referred to<br />

Wilm and Mann’s off line technique, evolving technology<br />

has made possible the use of on line flow rates down to tens<br />

of nL/min. The term nanoelectrospray has thus expanded<br />

to encompass both on line and off line techniques.<br />

For nanoelectrospray, the effects of tip <strong>geometry</strong>, flow<br />

rate and solvent composition have been investigated by a<br />

number of groups [55–57].<br />

Motivated by the fact that, at certain solvent compositions,<br />

the sensitivity decreased dramatically, Vanhoutte<br />

et al. compared the effects of various mobile phase<br />

compositions using several types of electrospray capillaries.<br />

These included combinations of uncoated or gold<br />

coated, tapered or non-tapered fused silica capillary tips,<br />

as well as stainless steel emitters. They found that the<br />

gold coated, tapered tips were the most effective, because<br />

they covered the most extensive range of solvent composition<br />

<strong>for</strong> the tested conditions without loss of sensitivity<br />

[55].<br />

Schmidt and Karas used combinations of model compounds<br />

to examine the effect of flow rate on ion signal<br />

[56]. They determined that, at flow rates of a few<br />

nL/min, signal suppression caused by differences in<br />

surface activity of the analytes was insignificant compared<br />

to that at higher flow rates (> 50 nL/min).<br />

Li and Cole studied the often observed shift in charge<br />

state as a result of changing experimental parameters <strong>for</strong><br />

nanoelectrospray [57]. Parameters such as tip diameter,<br />

flow rate, analyte concentration and solvent composition<br />

can all affect the observed ions and the charge states.<br />

5. Currently available commercial <strong>source</strong>s<br />

Since recognition of ESI as an invaluable bioanalytical<br />

tool in the late 1980s, research groups have attempted<br />

to exploit the capabilities of ESI by modifying the <strong>source</strong><br />

<strong>geometry</strong> in order to allow a wider range of flow rates<br />

and in <strong>source</strong> fragmentation, as well as improved sensitivity,<br />

efficiency and practicality. Niessen’s 1998 and<br />

2003 review articles [58,59] are a good starting point<br />

<strong>for</strong> those interested in the trends that have stimulated<br />

ESI <strong>source</strong> evolution. Despite numerous improvements,<br />

commercial ESI <strong>source</strong>s still retain many features of the<br />

original configuration.<br />

5.1. Source configuration and its relationship to flow<br />

rate<br />

A parameter of critical importance in ESI-MS is the flow<br />

rate of the solution. Depending on the application <strong>for</strong><br />

which the <strong>mass</strong> spectrometer is used, the ESI <strong>source</strong><br />

needs to be adapted to handle the in coming flow of<br />

solution, whether it is direct infusion, or LC or CE eluent.<br />

5.1.1. Low flow rate. While flow rates from approximately<br />

1 lL/min up to several mL/min are typically used<br />

with the assistance of a nebulizer, traditional nanoelectrospray<br />

relied simply on the electrostatic attraction between<br />

the mobile phase and the counter electrode to<br />

generate and disperse the charged droplets. The nanoflow<br />

regime can generally be defined as using flow rates<br />

of less than 1 lL/min and can extend to levels of less<br />

than 1 nL/min [55]. As mentioned previously, nanoflow<br />

ESI can be used independently (off line) or coupled to a<br />

separation system (on line), such as CE or HPLC.<br />

Depending on experimental requirements, the commercially<br />

available nanoelectrospray emitters vary in material,<br />

tip shape, diameter, and the configuration used <strong>for</strong><br />

electrical contact. On line analyses typically use higher<br />

flow rates (0.1–1 lL/min), and the tips are usually fabricated<br />

from fused silica or metal. Off line analyses typically<br />

use flow rates in the low nL/min range, with<br />

coated or uncoated glass or quartz emitters. A number of<br />

groups have also described modified nanoflow sprayers<br />

with various coatings [60–62] or inserted fibers [63].<br />

There are many different configurations <strong>for</strong> nanoflow<br />

LC-MS coupling [64], although they all include an LC<br />

column and a narrow diameter emitter tip. The simplest<br />

combination involves coupling an upstream LC column<br />

to the emitter tip using some type of low dead volume<br />

union [65]. The union can also serve as the electrical<br />

contact <strong>for</strong> the electrospray process, although it is also<br />

possible to use conductive tips or other electrode configurations<br />

[64,66]. The advantage of decoupling the LC<br />

column from the sprayer is simple, inexpensive replacement<br />

of sprayer tips, should they become damaged or<br />

plugged. However, the coupling must be done carefully<br />

to maintain chromatographic per<strong>for</strong>mance.<br />

An alternative combination involves packing the LCcolumn<br />

material directly into the sprayer tip to eliminate<br />

any dead volume after the column [67–70]. Tapered<br />

sprayers containing LC column packing are available<br />

commercially from companies such as New Objective<br />

Inc. The column preparation typically involves passing a<br />

http://www.elsevier.com/locate/trac 249


Trends Trends in Analytical Chemistry, Vol. 25, No. 3, 2006<br />

slurry containing column packing material through a<br />

tapered sprayer [71,72]. Sometimes, a small frit is included<br />

inside the tip of the sprayer to help confine the<br />

packing material. The elimination of post column dead<br />

volume helps to ensure optimal chromatographic per<strong>for</strong>mance.<br />

In addition, it has been suggested that the<br />

presence of packing material in the tapered sprayer acts<br />

similar to a filter, extending tip lifetimes by preventing<br />

plugging. The drawback with this configuration is that<br />

tip replacement due to blockage or damage requires the<br />

entire column to be replaced. Electrical contact can be<br />

established at the sprayer tip or distal to the tip end.<br />

Amirkhani and co-workers recently compared the<br />

efficiency of four different sheathless electrospray emitter<br />

configurations <strong>for</strong> a nanoflow LC system [73]. Two of the<br />

configurations used on-column emitters with the applied<br />

voltage either at the outlet or the inlet end of the<br />

column, and the other two had emitters coupled to<br />

columns, with the electrical connection at either the<br />

sprayer tip or the low dead volume union. It was<br />

demonstrated that all the configurations worked equally<br />

well, provided the connections were made with minimal<br />

dead volume [73].<br />

Currently, most major MS manufacturers (e.g., Thermo<br />

Electron, MDS SCIEX, Agilent Technologies, and Waters/<br />

Micro<strong>mass</strong>) make the nanoelectrospray <strong>source</strong> available.<br />

The number of applications using electrospray at low flow<br />

rates is very large, and outside the scope of this paper.<br />

Wood et al.’s 2003 review article is a good starting point<br />

and a <strong>source</strong> of references [47].<br />

5.1.2. High flow rate. If the flow rate is in the range<br />

0.05–3 mL/min, sensitivity can be an issue, due to the<br />

decrease in <strong>ionization</strong> efficiency resulting from large<br />

droplet size. ESI-MS is widely interfaced with LC, so high<br />

flow rates are often necessary. One solution to this<br />

problem is to employ a splitter that allows only a limited<br />

volumetric flow rate to reach the MS interface. Under<br />

these conditions, part of the eluent is wasted and connection<br />

dead volume is a common problem. A more<br />

practical approach adopted by manufacturers is to<br />

sample the spray from a region, peripheral to the main<br />

droplet trajectory, where the mist is much finer. This is<br />

done by simply re-orienting the sprayer relative to the<br />

interface so that the fine droplets from the exterior of<br />

the spray plume can enter the sampling inlet, while the<br />

majority of large droplets are directed away from the<br />

entrance [74]. To minimize contamination, many major<br />

MS manufacturers now sample orthogonally from the<br />

spray plume <strong>for</strong> many applications (e.g., Agilent Technologies,<br />

Waters/Micro<strong>mass</strong>, and MDS SCIEX). An<br />

example of this is shown in Fig. 7 <strong>for</strong> an Agilent Technologies<br />

<strong>source</strong> incorporating an additional asymmetrical<br />

lens.<br />

The asymmetrical lens is essentially a half-full, partial<br />

cylinder electrode operated at high voltage during the<br />

250 http://www.elsevier.com/locate/trac<br />

asymmetrical<br />

lens<br />

orthogonal<br />

nebulizer<br />

nebulized<br />

analyte solution<br />

HPLC eluent<br />

atmospheric pressure<br />

1.5 torr<br />

ion transfer capillary<br />

to MS<br />

Figure 7. Schematic of an ESI <strong>source</strong> configuration developed by<br />

Agilent Technologies.<br />

electrospray process [75]. The purpose of the electrode is<br />

to help initiate and sustain electrospray. Conversely,<br />

other groups have attempted to use various types of<br />

auxiliary electrodes [76–78] to focus ions at atmospheric<br />

pressure.<br />

Also exploiting the advantages of orthogonal sampling<br />

is the Waters/Micro<strong>mass</strong> Zspraye [79,80] interface,<br />

shown schematically in Fig. 8. The spray is first sampled<br />

orthogonally through the sampling cone into a low<br />

pressure chamber. An extraction cone (skimmer) is oriented<br />

at a right angle relative to the axis of the spray to<br />

sample <strong>for</strong> a second time into the next differentially<br />

pumped vacuum stage. The double orthogonal sampling<br />

system prevents solvent and neutral molecules from<br />

entering the analyzer, resulting in reduced chemical<br />

background [81]. Larger cone apertures are used to<br />

Figure 8. Schematic of the Zspray TM <strong>source</strong> configuration used by<br />

Waters/Micro<strong>mass</strong>.


Trends in Analytical Chemistry, Vol. 25, No. 3, 2006 Trends<br />

compensate <strong>for</strong> transmission losses. The ability to disassemble<br />

and to clean the sampling cone without<br />

breaking vacuum is an additional advantage contributing<br />

to the ruggedness of the Zspray <strong>source</strong>.<br />

Although some MS manufacturers (e.g., MDS SCIEX,<br />

Waters/Micro<strong>mass</strong>, and Agilent Technologies) have opted<br />

<strong>for</strong> orthogonal sampling, Thermo Electron uses a<br />

sampling angle at 60° from the ion optics axis. In its Ion<br />

Max <strong>source</strong>, the function of the angled position of the<br />

sprayer is similar to the orthogonal one.<br />

Sprayer orientation is not the only important parameter<br />

when it comes to optimizing sensitivity while using<br />

high flow rates. For conventional high flow rate ESI<br />

<strong>source</strong>s, the nebulizer gas is an essential component. Air<br />

or nitrogen is typically used to disperse the emerging<br />

solution into small droplets and to direct the droplets on<br />

the trajectory chosen <strong>for</strong> optimal sampling.<br />

To vaporize the large amounts of solvent emerging<br />

from the sprayer as efficiently as possible, manufacturers<br />

have introduced additional features to assist the nebulizer<br />

gas. MDS SCIEX developed the TurboIonSpray<br />

<strong>source</strong> [21,22], in which heated nitrogen gas that is<br />

released from a unit external to the sprayer is used to<br />

assist evaporation of the spray droplets at atmospheric<br />

pressure. More recently, MDS SCIEX took this design a<br />

step further, creating the TurboVe <strong>source</strong>, shown in<br />

Fig. 9. In this case, two heated auxiliary nitrogen <strong>source</strong>s<br />

are oriented to achieve very efficient desolvation. The<br />

improved desolvation permits stable, sensitive operation<br />

<strong>for</strong> flow rates of greater than 1 mL/min [82].<br />

Thermo Electron also integrated auxiliary nitrogen gas<br />

<strong>for</strong> desolvation purposes, except that it flows directly<br />

from the nozzle of the Ion Max <strong>source</strong> [83]. Similar to the<br />

TurboV, the Ion Max uses a high temperature <strong>source</strong><br />

heater incorporated into the housing of the ion <strong>source</strong>.<br />

sampling orifice<br />

exhaust<br />

ESI<br />

probe<br />

heater<br />

spray<br />

heated nitrogen<br />

Figure 9. Schematic of the TurboV TM <strong>source</strong> used by MDS SCIEX.<br />

This probe configuration is referred to as Heated <strong>Electrospray</strong><br />

Ionization (H-ESI) and it tolerates the use of<br />

higher flow rates than previously possible on Thermo<br />

Electron instruments.<br />

5.2. Interface<br />

Various ion sampling interface configurations have been<br />

developed [8]. The sampling orifice can be built in a disc,<br />

in the top or the bottom of a cone, or in glass or metal<br />

tubes. Some MS manufacturers now equip their instruments<br />

with conical interfaces. Intuitively, the conical<br />

shape is meant to improve the electric field density at the<br />

sampling orifice in an attempt to improve ion transmission.<br />

To avoid potential contamination, the sampling<br />

cone can be heated to evaporate residuals from its<br />

surface. Since high temperatures are required <strong>for</strong> this<br />

purpose, ceramic materials have often been used <strong>for</strong> the<br />

interface. Ceramic produces no outgassing, and is<br />

thought to have low sample adsorptivity, thus reducing<br />

memory effects on the interface surface [84].<br />

6. Multiple sprayers<br />

Commercial multiple sprayer systems <strong>for</strong> ESI have<br />

evolved mainly as a result of industry’s increasing<br />

requirements <strong>for</strong> high throughput sample analysis, particularly<br />

<strong>for</strong> pharmaceutical samples. With a multiple<br />

sprayer <strong>source</strong> and suitable equipment, LC-MS analyses<br />

can be done quickly and efficiently. However, the first<br />

attempts at building multiple ESI sprayers were motivated<br />

by other reasons. Smith et al. [30,85] were the first<br />

to describe the coupling of two ESI sprayers in order to<br />

generate ions of opposite polarity and study ion-ion<br />

chemistry. McLuckey et al. also interfaced up to three ion<br />

<strong>source</strong>s (two ESI <strong>source</strong>s and one atmospheric sampling<br />

glow discharge <strong>ionization</strong> (ASGDI) <strong>source</strong>) onto a<br />

quadrupole ion trap <strong>mass</strong> spectrometer in order to control<br />

the charge states of reactant and product ions<br />

[86,87]. Other multiple <strong>source</strong> combinations have also<br />

been described, incorporating ESI with <strong>source</strong>s such as<br />

atmospheric pressure chemical <strong>ionization</strong> (APCI)<br />

[88,89].<br />

Using a linear array of capillary electrodes, Rulison<br />

et al. demonstrated the feasibility of increasing sample<br />

throughput by using parallel capillaries [90]. They also<br />

observed that the Taylor cones at the ends of the array<br />

were being deflected due to end effects caused by the<br />

electric field.<br />

In 1994, Kostiainen and Bruins proposed that multiple<br />

sprayers could be used to improve ion current stability<br />

under high flow rate conditions (i.e., above 200 lL/min)<br />

[91]. Four years later, Kassel and Zeng developed a dual<br />

ESI <strong>source</strong> <strong>for</strong> purification and simultaneous analysis of<br />

combinatorial libraries using two separate LC-MS eluent<br />

streams [92]. Although this development paved the way<br />

http://www.elsevier.com/locate/trac 251


Trends Trends in Analytical Chemistry, Vol. 25, No. 3, 2006<br />

<strong>for</strong> further ef<strong>for</strong>ts in high throughput parallel analysis,<br />

Kassel’s system required knowledge of the particular<br />

analyte that each sprayer was generating. Dual ESI<br />

<strong>source</strong>s have also been used to introduce an internal<br />

calibration solution separately from the analyte solution<br />

[93–96]. This configuration provides benefits compared<br />

with a single solution containing both analyte and<br />

calibrant because it avoids preferential <strong>ionization</strong>.<br />

In 1999, Kassel et al. [97] and De Biassi et al. [98]<br />

described indexed multiple sprayer systems that allow<br />

sequential sampling of each spray by a <strong>mass</strong> spectrometer,<br />

thus making it possible to identify ions from their<br />

respective eluent stream.<br />

Currently, publications in which fast LC-MS analyses<br />

of samples are discussed indicate that the Waters/<br />

Micro<strong>mass</strong> multiplexed ESI <strong>source</strong> (MUX) is the most<br />

commonly used <strong>for</strong> the analysis of multiple LC eluent<br />

streams [99–101]. Tiller’s 2003 review paper provides<br />

useful references regarding biological applications using<br />

the MUX system as well as other types of parallel analysis<br />

[102]. The MUX system attaches to the Zspray<br />

<strong>source</strong> of the <strong>mass</strong> spectrometer and is fitted with either<br />

four or eight channels. As a rotating aperture allows<br />

only one spray at a time to reach the sampling cone,<br />

spectra of compounds eluting from individual separation<br />

columns can be quickly acquired.<br />

However, a significant problem with MUX-type<br />

systems is a decrease in sensitivity of approximately<br />

three-fold at high flows [103]. Yang et al. attributed the<br />

drop in sensitivity to several reasons, such as difficulty in<br />

optimizing sprayer position and lower desolvation<br />

efficiency due to large amounts of solvent introduced<br />

into the <strong>source</strong> in combination with a desolvation gas<br />

that was counter-current relative to the spray instead of<br />

the traditional coaxial flow. The sensitivity decrease is<br />

expected to be larger <strong>for</strong> nanoflow operation, where<br />

sprayer position can be more critical.<br />

In 2002, Schneider et al. [104] developed an alternative<br />

multiple sprayer system that appeared to solve the sensitivity<br />

problem. By installing an atmospheric pressure ion<br />

lens [76,105] on each sprayer of a dual sprayer and a four<br />

sprayer ESI <strong>source</strong>, they were able to maintain high sensitivity<br />

by optimizing the applied lens potential as well as<br />

sprayer position. An additional advantage of the system<br />

was that sprayers could be enabled or disabled by<br />

changing the lens potential, potentially a faster method<br />

than using mechanical devices.<br />

7. Recent developments in ESI <strong>source</strong>s<br />

The quest <strong>for</strong> a superior <strong>source</strong> design capable of effectively<br />

focusing more of the electrospray generated ion<br />

cloud through the sampling aperture of the instrument is<br />

still on going. Due to gas phase collisions occurring<br />

under atmospheric pressure conditions and space charge<br />

252 http://www.elsevier.com/locate/trac<br />

effects, only a small percentage of the spray plume is<br />

actually sampled. There is still significant room <strong>for</strong><br />

improvement in ion sampling efficiency.<br />

7.1. High velocity nebulizers<br />

One approach developed by Zhou et al. [106] involved<br />

installing an industrial air amplifier between the electrospray<br />

probe and the sampling orifice. The concentric,<br />

high velocity gas flow was meant to control the expansion<br />

of the spray plume as well as assist in the desolvation<br />

of ions. In addition, focusing of ions could be<br />

achieved by applying an optimized voltage of up to 3 kV<br />

onto the amplifier. Thus, when both the air amplifier and<br />

the high voltage were used, Zhou’s ion <strong>source</strong> configuration<br />

seemed to provide some improvement in ionsampling<br />

efficiency. Suppression of background chemical<br />

noise was also observed.<br />

7.2. Improved nanoflow configurations<br />

Schneider et al. developed a novel interface <strong>for</strong> nanoflow<br />

ESI (Fig. 10) that provides two separate atmospheric<br />

pressure regions <strong>for</strong> droplet desolvation as well as two<br />

regions <strong>for</strong> removing unwanted particles [107].<br />

The first desolvation stage makes use of the traditional<br />

counter-current gas to eliminate neutral particles and<br />

solvent. Next, charged particles, ions and gas travel<br />

through the heated laminar flow chamber, where additional<br />

desolvation takes place at optimal temperature.<br />

Once they reach the particle discriminator area, ions are<br />

carried by the gas streamlines through the discriminator<br />

space and into the sampling orifice, while large charged<br />

particles with high momentum escape from the gas<br />

streamline trajectory under the influence of the strong<br />

electric field. This <strong>source</strong> design provided improved signal<br />

and ion current stability <strong>for</strong> flow rates from a few nL/<br />

min up to 1 lL/min as well as improved per<strong>for</strong>mance <strong>for</strong><br />

gradient elution nanoLC-MS in both positive-ion and<br />

negative ion mode [108].<br />

7.3. Desorption electrospray <strong>ionization</strong><br />

Cooks and co-workers recently described a new method<br />

of <strong>ionization</strong> in which analytes are desorbed from<br />

surfaces by nebulizing the charged droplets of an<br />

electrospray plume at a surface [109]. This technique is<br />

referred to as desorption electrospray <strong>ionization</strong> (DESI).<br />

The exact mechanism of ion <strong>for</strong>mation in DESI is not<br />

well understood, although the technique has been<br />

successfully used <strong>for</strong> analyzing a number of samples,<br />

including pills and TLC plates [110]. Efficient sampling of<br />

desorbed ions requires careful optimization of the angles<br />

and the spacings between the surface, the electrospray<br />

emitter, and the <strong>mass</strong> spectrometer inlet.<br />

7.4. Coupling ESI to micro-analytical systems<br />

In the rapidly expanding area of microchip chemical<br />

analysis, currently known as ‘‘Micro-Total Analysis


Trends in Analytical Chemistry, Vol. 25, No. 3, 2006 Trends<br />

Curtain Plate<br />

Tee and Nanospray<br />

Tip with Nebulizer<br />

Curtain Gas<br />

Systems’’ (l-TASs) and the related ‘‘Lab-on-a-chip’’, ESI<br />

has become the preferred interface method <strong>for</strong> MS<br />

detection [111–113]. More than a decade after being<br />

introduced [114], l-TASs have been used in a wide<br />

range of biological applications, drug discovery, and<br />

clinical diagnostics. References regarding applications<br />

can be found in review papers [115–120].<br />

The interest in miniaturized analytical systems rose<br />

due to the potential <strong>for</strong> integrating time consuming<br />

steps, such as sample preparation, chemical reaction,<br />

separation, and detection, into a fast, highly automated<br />

technique, while at the same time using minute<br />

sample quantities. Constantly evolving microfabrication<br />

techniques adapted from the semiconductor<br />

industry are used to produce the microfluidic devices<br />

[117,121,122].<br />

Although on-chip optical detection is often used in<br />

microfluidic applications, MS detection is desirable due to<br />

its suitability <strong>for</strong> analyzing biological molecules. In<br />

addition, the low flow rates (nL–lL/min) used with<br />

microfluidic devices are optimal <strong>for</strong> the ESI <strong>source</strong>.<br />

However, integrating ESI-MS with a microchip is not<br />

easy. Initial work involved using traditional ESI emitters,<br />

such as a section of fused silica capillary or a nanoelectrospray<br />

needle attached to a specific microchannel exit,<br />

but these types of <strong>source</strong>s were prone to contamination<br />

problems from bonding agents, poor reproducibility, and<br />

dead volume problems [121,123,124]. Work has<br />

there<strong>for</strong>e focused on building the ESI emitter as part of<br />

the microdevice and a range of materials, from silicon<br />

based to polymers, has been tested <strong>for</strong> fabrication of the<br />

ESI emitter [117,121].<br />

Agilent Technologies has introduced a polymer based<br />

HPLC-MS microfluidic chip [125]. The chip contains<br />

(TM)<br />

Particle Discriminator<br />

Space<br />

Heated Laminar Flow<br />

Chamber<br />

Teflon Spacer<br />

Orifice<br />

Figure 10. Schematic of the particle discriminator interface developed by MDS SCIEX.<br />

enrichment and analytical columns packed with reversephase<br />

material, a face-seal rotary valve <strong>for</strong> switching<br />

between loading and separation positions, and a conical<br />

ESI emitter created directly on the chip structure by laser<br />

ablation. Gradient reverse phase separation of peptides<br />

has been achieved at flow rates of 100–400 nL/min. The<br />

system was reported to be stable with various solvents,<br />

sensitive, and robust. Since the various connections<br />

necessary <strong>for</strong> traditional nanoelectrospray were eliminated,<br />

dead volume problems as well as analyte dispersion<br />

were minimized [125].<br />

While the combination of microfluidic devices with MS<br />

may have great potential, many barriers still need to be<br />

overcome. Currently, CE methods appear to be most<br />

common <strong>for</strong> sample separation on microfluidic chips,<br />

because they are less complex to miniaturize than HPLC<br />

[126]. Difficulties, such as integrating pumping systems<br />

<strong>for</strong> controlling flow and creating gradients, still limit<br />

widespread commercialization of microfluidic-LC systems<br />

[127].<br />

For high throughput analyses, Advion BioSciences<br />

successfully introduced a fully automated nanoelectrospray<br />

system, called the NanoMatee. This microchip<br />

device contains an array of nozzles to which the sample<br />

is automatically delivered <strong>for</strong> MS analysis. Short analysis<br />

times are possible due to a simplified spray optimization<br />

procedure [128,129]. Although the NanoMate was initially<br />

intended <strong>for</strong> use in combination with microfluidic<br />

separation devices [128], it has been integrated with<br />

<strong>source</strong>s from other manufacturers (e.g., Waters/Micro<strong>mass</strong>,<br />

MDS SCIEX, and Thermo Electron). The key<br />

advantages of the NanoMate include elimination of<br />

sample carry over and simplification of the tuning procedure<br />

[130].<br />

http://www.elsevier.com/locate/trac 253


Trends Trends in Analytical Chemistry, Vol. 25, No. 3, 2006<br />

8. Conclusion and future outlook<br />

Because ESI-MS is used in many areas of chemistry,<br />

there is an immense number of publications covering<br />

modifications to the ion <strong>source</strong> <strong>for</strong> various applications.<br />

This article has attempted to sketch a basic picture of<br />

how ESI works and how it has evolved, and continues to<br />

do so, towards new designs. Undoubtedly, commercial<br />

ESI <strong>source</strong>s will continue to evolve to meet growing<br />

demands <strong>for</strong> improved sensitivity and robustness. Further<br />

improvements are likely to include increasing the<br />

gas throughput to sample a larger number of ions, even<br />

though these types of approaches are typically associated<br />

with higher costs due to the increased pumping<br />

requirements. Research is also likely to continue to increase<br />

the ion sampling efficiency by electrostatic and<br />

aerodynamic focusing, without the need to increase<br />

system pumping. With MS more popular than ever, it is<br />

fascinating to see that the capabilities of ESI-MS are still<br />

expanding.<br />

Acknowledgement<br />

The authors thank Dr. Damon Robb <strong>for</strong> valuable discussions<br />

on many topics included in this article.<br />

References<br />

[1] N. Mano, J. Goto, Anal. Sci. 19 (2003) 3.<br />

[2] R.B. Cole (Editor), <strong>Electrospray</strong> Ionization Mass Spectrometry<br />

Fundamentals Instrumentation and Applications, John Wiley &<br />

Sons, Inc., New York, USA, 1997.<br />

[3] M. Dole, L.L. Mack, R.L. Hines, J. Chem. Phys. 49 (1968) 2240.<br />

[4] A.T. Blades, M.G. Ikonomou, M.G. Kebarle, Anal. Chem. 63<br />

(1991) 2109.<br />

[5] G.J. Van Berkel, F. Zhou, Anal. Chem. 64 (1992) 1586.<br />

[6] G.J. Van Berkel, in: R.B. Cole (Editor), <strong>Electrospray</strong> Ionization<br />

Mass Spectrometry, John Wiley & Sons, Inc., New York, USA,<br />

1997 Chap. 1, p. 65.<br />

[7] R.B. Cole, J. Mass. Spectrom. 35 (2000) 763.<br />

[8] A.P. Bruins, in: R.B. Cole (Editor), <strong>Electrospray</strong> Ionization Mass<br />

Spectrometry: Fundamentals, Instrumentation, and Applications,<br />

John Wiley & Sons, Inc., New York, USA, 1997, p. 107.<br />

[9] G. Taylor, Proc. R. Soc. London, Ser. A 280 (1964) 383.<br />

[10] R. Juraschek, F.W. Rollgen, Int. J. Mass Spectrom. 177 (1998)<br />

1.<br />

[11] G.A. Valaskovic, J.P.I. Murphy, J. Am. Soc. Mass Spectrom. 15<br />

(2004) 1201.<br />

[12] M. Karas, U. Bahr, T. Dulcks, Fresenius’ J. Anal. Chem. 366<br />

(2000) 669.<br />

[13] J.V. Iribarne, B.A. Thomson, J. Chem. Phys. 64 (1976) 2287.<br />

[14] L.L. Mack, P. Kralik, A. Rheude, M. Dole, J. Chem. Phys. 52<br />

(1969) 4977.<br />

[15] G. Schmelzeisen-Redeker, L. Butfering, F.W. Rollgen, Int. J. Mass<br />

Spectrom. Ion Processes 90 (1989) 139.<br />

[16] B.A. Thomson, J.V. Iribarne, J. Chem. Phys. 71 (1979) 4451.<br />

[17] P. Kebarle, M. Peschke, Anal. Chim. Acta 406 (1999) 11.<br />

[18] N. Felitsyn, M. Peschke, P. Kebarle, Int. J. Mass Spectrom. 219<br />

(2002) 39.<br />

254 http://www.elsevier.com/locate/trac<br />

[19] M. Gamero-Castano, F.J. de la Mora, J. Mass. Spectrom. 35<br />

(2000) 790.<br />

[20] N.B. Cech, C.G. Enke, Mass Spectrom. Rev. 20 (2001) 362.<br />

[21] L.Y.T. Li, D.A. Campbell, P.K. Bennett, J. Henion, Anal. Chem. 68<br />

(1996) 3397.<br />

[22] J.P. Allanson, R.A. Biddlecombe, A.E. Jones, S. Pleasance, Rapid<br />

Commun. Mass Spectrom. 10 (1996) 811.<br />

[23] S.K. Chowdhury, V. Katta, B.T. Chait, Rapid Commun. Mass<br />

Spectrom. 4 (1990) 81.<br />

[24] C.N. McEwen, B.S. Larsen, in: R.B. Cole (Editor), <strong>Electrospray</strong><br />

Ionization Mass Spectrometry, John Wiley & Sons, Inc., New<br />

York, USA, 1997, p. 177.<br />

[25] J.B. Fenn, M. Mann, C.K. Meng, S.F. Wong, C.M. Whitehouse,<br />

Mass Spectrom. Rev. 9 (1990) 37.<br />

[26] M. Yamashita, J.B. Fenn, J. Phys. Chem. 88 (1984) 4671.<br />

[27] C.M. Whitehouse, R.N. Dreyer, M. Yamashita, J.B. Fenn, Anal.<br />

Chem. 57 (1985) 675.<br />

[28] J. Franzen, Bruker Franzen Analytik GmbH, US Patent 5736740,<br />

(1998).<br />

[29] M.J. Tomany, J.A. Jarrell, Millipore Corp., US Patent 5304798,<br />

(1994).<br />

[30] L.R.R. Ogorzalek, H.R. Udseth, R.D. Smith, J. Phys. Chem. 95<br />

(1991) 6412.<br />

[31] L. Ogorzalek, R.D. Smith, J. Am. Chem. Soc. 5 (1994) 207.<br />

[32] L. Jiang, M. Moini, Anal. Chem. 72 (2000) 20.<br />

[33] K. Taeman, H.R. Udseth, R.D. Smith, Anal. Chem. 72 (2000)<br />

5014.<br />

[34] I.C. Mylchreest, M.E. Hail, Finnigan Corp., US Patent 5171990,<br />

(1992).<br />

[35] I.C. Mylchreest, M.E. Hail, Finnigan Corp., US Patent re-issue<br />

RE35413E, (1996).<br />

[36] A. Mordehai, S.J. Buttrill, Varian Associates, US Patent<br />

5672868, (1997).<br />

[37] A.P. Bruins, T.R. Covey, J.D. Henion, Anal. Chem. 59 (1987)<br />

2642.<br />

[38] R.D. Smith, C.J. Barinaga, H.R. Udseth, Anal. Chem. 60 (1988)<br />

1948.<br />

[39] J.A. Olivares, N.T. Nguyen, C.R. Yonker, R.D. Smith, Anal.<br />

Chem. 59 (1987) 1230.<br />

[40] P. Schmitt-Kopplin, M. Frommberger, Electrophoresis 24 (2003)<br />

3837.<br />

[41] M. Moini, Anal. Bioanal. Chem. 373 (2002) 466.<br />

[42] A. von Brocke, G. Nicholson, E. Bayer, Electrophoresis 22 (2001)<br />

1251.<br />

[43] M. Mazereeuw, A.J. Hofte, U.R. Tjaden, J. van der Greef, Rapid<br />

Commun. Mass Spectrom. 11 (1997) 981.<br />

[44] E.D. Lee, W. Muck, J.D. Henion, T.R. Covey, J. Chromatogr. 458<br />

(1988) 313.<br />

[45] E.D. Lee, W. Muck, J.D. Henion, T.R. Covey, Biomed. Environ.<br />

Mass Spectrom. 18 (1989) 844.<br />

[46] J. Cai, J. Henion, J. Chromatogr. A 703 (1995) 667.<br />

[47] T.D. Wood, M.A. Moy, A.R. Dolan, P.M.J. Bigwarfe, T.P. White,<br />

D.R. Smith, D.J. Higbee, Appl. Spectrosc. Rev. 38 (2003)<br />

187.<br />

[48] M.R. Emmett, R.M. Caprioli, J. Am. Soc. Mass Spectrom. 5<br />

(1994) 605.<br />

[49] E. Andren, M.R. Emmett, R.M. Caprioli, J. Am. Soc. Mass<br />

Spectrom. 5 (1994) 867.<br />

[50] E. Gelpi, J. Mass Spectrom. 37 (2002) 241.<br />

[51] J. Abian, A.J. Oosterkamp, E. Gelpi, J. Mass Spectrom. 34 (1999)<br />

244.<br />

[52] P.E. Andren, R.M. Caprioli, J. Mass Spectrom. 30 (1995) 817.<br />

[53] M.S. Wilm, M. Mann, Int. J. Mass Spectrom. Ion Processes 136<br />

(1994) 167.<br />

[54] M. Wilm, M. Mann, Anal. Chem. 68 (1996) 1.<br />

[55] K. Vanhoutte, W. Van Dongen, E.L. Esmans, Rapid Commun.<br />

Mass Spectrom. 12 (1998) 15.


Trends in Analytical Chemistry, Vol. 25, No. 3, 2006 Trends<br />

[56] A. Schmidt, M. Karas, T. Dulcks, J. Am. Soc. Mass Spectrom. 14<br />

(2003) 492.<br />

[57] Y. Li, R.B. Cole, Anal. Chem. 75 (2003) 5739.<br />

[58] W.M.A. Niessen, J. Chromatogr. A 794 (1998) 407.<br />

[59] W.M.A. Niessen, J. Chromatogr. A 1000 (2003) 413.<br />

[60] D.R. Barnidge, S. Nilsson, K.E. Markides, Anal. Chem. 71 (1999)<br />

4115.<br />

[61] Y.Z. Chang, Y.R. Chen, G.R. Her, Anal. Chem. 73 (2001) 5083.<br />

[62] E.P. Maziarz III, S.A. Lorenz, T.P. White, T.D. Wood, J. Am. Soc.<br />

Mass Spectrom. 11 (2000) 659.<br />

[63] J. Liu, K.W. Ro, M. Busman, D.R. Knapp, Anal. Chem. 76 (2004)<br />

3599.<br />

[64] M.T. Davis, D.C. Stahl, S.A. Hefta, T.D. Lee, Anal. Chem. 67<br />

(1995) 4549.<br />

[65] J.N.I. Alexander, G.A. Schultz, J.B. Poly, Rapid Commun. Mass<br />

Spectrom. 12 (1998) 1187.<br />

[66] A.J. Oosterkamp, E. Gelpi, J. Abian, J. Mass Spectrom. 33 (1998)<br />

976.<br />

[67] K.B. Tomer, M.A. Moseley, L.J. Deterding, C.E. Parker, Mass<br />

Spectrom. Rev. 13 (1994) 431.<br />

[68] L.J. Licklider, C.C. Thoreen, J. Peng, S.P. Gygi, Anal. Chem. 74<br />

(2002) 3076.<br />

[69] T. Natsume, Y. Yamauchi, H. Nakayama, T. Shinkawa, M.<br />

Yanagida, N. Takahashi, T. Isobe, Anal. Chem. 74 (2002) 4725.<br />

[70] I.I. Stewart, L. Zhao, T. Le Bihan, B. Larsen, S. Scozzaro,<br />

D. Figeys, G.D. Mao, O. Ornatsky, M. Dharsee, C. Orsi, R. Ewing,<br />

T. Goh, Rapid Commun. Mass Spectrom. 18 (2004) 1697.<br />

[71] C.L. Gatlin, G.R. Kleemann, L.G. Hays, A.J. Link, J.R. Yates, Anal.<br />

Biochem. 263 (1998) 93.<br />

[72] T. Le Bihan, H.S. Duewel, D. Figeys, J. Am. Soc. Mass Spectrom.<br />

14 (2003) 713.<br />

[73] A. Amirkhani, M. Wetterhall, S. Nilsson, R. Danielsson,<br />

J. Bergquist, J. Chromatogr. A 1033 (2004) 257.<br />

[74] R.D. Voyksner, H. Lee, Anal. Chem. 71 (1999) 1441.<br />

[75] J.L. Bertsch, S.M. Fisher, K.D. Henry, E.M. Wong, Hewlett-<br />

Packard Company, US Patent 5838003, (1998).<br />

[76] B.B. Schneider, D.J. Douglas, D.D.Y. Chen, J. Am. Soc. Mass<br />

Spectrom. 13 (2002) 906.<br />

[77] X. Xu, J. Zhai, W. Shui, G. Xu, P. Yang, Anal. Lett. 37 (2004)<br />

2711.<br />

[78] W.J. Thompson, J.W. Eschelbach, R.T. Wilburn, J.W. Jorgenson,<br />

J. Am. Soc. Mass Spectrom. 16 (2005) 312.<br />

[79] T. Bantan, R. Milacic, B. Mitrovic, B. Pihlar, J. Anal. At.<br />

Spectrom. 14 (1999) 1743.<br />

[80] M.E. McComb, L.J. Donald, H. Perreault, Can. J. Chem. 77 (1999)<br />

1752.<br />

[81] M. Holcapek, K. Volna, P. Jandera, L. Kolarova, K. Lemr,<br />

M. Exner, A. Cirkva, J. Mass Spectrom. 39 (2004) 43.<br />

[82] P.K.S. Blay, S. Brombacher, D.A. Volmer, Rapid Commun. Mass<br />

Spectrom. 17 (2003) 2153.<br />

[83] C.Y. Yang, Y. Wang, M. Splendore, R.A. Thakur, ASMS<br />

Conference, Montreal, Canada, 2003.<br />

[84] M.P. Balogh, LC.GC Europe 17 (2004) 24.<br />

[85] L.R.R. Ogorzalek, H.R. Udseth, R.D. Smith, J. Am. Soc. Mass<br />

Spectrom. 3 (1992) 695.<br />

[86] E.R. Badman, P.A. Chrisman, S.A. McLuckey, Anal. Chem. 74<br />

(2002) 6237.<br />

[87] J.M. Wells, P.A. Chrisman, S.A. McLuckey, J. Am. Soc. Mass<br />

Spectrom. 13 (2002).<br />

[88] B.A. Andrien, C.M. Whitehouse, S. Shen, M.A. Sansone,<br />

Analytica of Bran<strong>for</strong>d, Inc., Int. PCT 13492, (1999).<br />

[89] P.H. Cormia, S.M. Fischer, D.L. Gourley, G.F. Ingle, Proc. 52 nd<br />

ASMS Conf. Mass Spectrom. Allied Topics, 23–27 May (2004),<br />

Nashville, TN, USA.<br />

[90] A.J. Rulison, R.C. Flagan, Rev. Sci. Instrum. 64 (1993) 683.<br />

[91] R. Kostiainen, A.P. Bruins, Rapid Commun. Mass Spectrom. 8<br />

(1994) 549.<br />

[92] L. Zeng, D.B. Kassell, Anal. Chem. 70 (1998) 4380.<br />

[93] A.I. Nepomuceno, D.C. Muddiman, R.H. Bergen III,<br />

J.R. Craighead, M.J. Burke, P.E. Caskey, J.A. Allan, Anal. Chem.<br />

75 (2003) 3411.<br />

[94] M. Moini, L. Jiang, Anal. Chem. 72 (2000) 20.<br />

[95] J.W. Flora, A.P. Null, D.C. Muddiman, Anal. Bioanal. Chem. 373<br />

(2002) 538.<br />

[96] F. Zhou, W. Shui, Y. Lu, P. Yang, Y. Guo, Rapid Commun. Mass<br />

Spectrom. 16 (2002) 505.<br />

[97] T. Wang, L. Zeng, J. Cohen, D.B. Kassell, Comb. Chem. High<br />

Throughput Screen. 2 (1999) 327.<br />

[98] V. de Biasi, N. Haskins, A. Organ, R. Bateman, K. Giles, S. Jarvis,<br />

Rapid Commun. Mass Spectrom. 13 (1999) 1165.<br />

[99] B. Feng, A.H. Patel, P.M. Keller, R.J. Slemmon, Rapid Commun.<br />

Mass Spectrom. 15 (2001) 821.<br />

[100] B. Feng, S. McQueney, T.M. Mezzasalma, R.J. Slemmon, Anal.<br />

Chem. 73 (2001) 5691.<br />

[101] D. Morrison, A.E. Davies, A.P. Watt, Anal. Chem. 74 (2002)<br />

1896.<br />

[102] P.R. Tiller, L.A. Romanyshyn, U.D. Neue, Anal. Bioanal. Chem.<br />

377 (2003) 788.<br />

[103] L. Yang, T.D. Mann, D. Little, N. Wu, R.P. Clement,<br />

P.J. Rudewicz, Anal. Chem. 73 (2001) 1740.<br />

[104] B.B. Schneider, D.J. Douglas, D.D.Y. Chen, Rapid Commun. Mass<br />

Spectrom. 16 (2002) 1982.<br />

[105] B.B. Schneider, D.J. Douglas, D.D.Y. Chen, Rapid Commun. Mass<br />

Spectrom. 15 (2001) 2168.<br />

[106] L. Zhou, B. Yue, D.V. Dearden, E.D. Lee, A.L. Rockwood,<br />

M.L. Lee, Anal. Chem. 75 (2003) 5978.<br />

[107] B.B. Schneider, V.I. Baranov, H. Javaheri, T.R. Covey, J. Am. Soc.<br />

Mass Spectrom. 14 (2003) 1236.<br />

[108] B.B. Schneider, X. Guo, L.M. Fell, T.R. Covey, J. Am. Soc. Mass<br />

Spectrom. 16 (2005) 1545.<br />

[109] Z. Takats, J.M. Wiseman, B. Golagan, R.G. Cooks, Science<br />

(Washington, DC) 306 (2004) 471.<br />

[110] G.J. Van Berkel, M.J. Ford, M.A. Deibel, Anal. Chem. 77 (2005)<br />

1207.<br />

[111] N. Lion, J.-O. Gellon, H. Jensen, H.H. Girault, J. Chromatogr. A<br />

1003 (2003) 11.<br />

[112] Y. Yang, J. Kameoka, T. Wachs, J.D. Henion, H.G. Craighead,<br />

Anal. Chem. 76 (2004) 2568.<br />

[113] I.M. Lazar, L. Li, Y. Yang, B.L. Karger, Electrophoresis 24 (2003)<br />

3655.<br />

[114] A. Manz, N. Graber, H.M. Widmer, Sens. Actuators B 1 (1990)<br />

244.<br />

[115] J. Khandurina, A. Guttman, J. Chromatogr. A 943 (2002)<br />

159.<br />

[116] P.-A. Auroux, D. Iossifidis, D.R. Reyes, A. Manz, Anal. Chem. 74<br />

(2002) 2637.<br />

[117] K. Huikko, R. Kostiainen, T. Kotiaho, Eur. J. Pharm. Sci. 20<br />

(2003) 149.<br />

[118] Y. Liu, C.D. Garcia, C.S. Henry, Analyst (Cambridge, UK) 128<br />

(2003) 1002.<br />

[119] S.J. Lee, S.Y. Lee, Appl. Microbiol. Biotechnol. 64 (2004) 289.<br />

[120] U. Bilitewski, M. Genrich, S. Kadow, G. Mersal, Anal. Bioanal.<br />

Chem. 377 (2003) 556.<br />

[121] D. Figeys, D. Pinto, Electrophoresis 22 (2001) 208.<br />

[122] A.E. Guber, M. Heckele, D. Hermann, A. Muslija, V. Saile,<br />

L. Eichhorn, T. Gietzelt, W. Hoffmann, P.C. Hauser,<br />

J. Tanyanyiwa, A. Gerlach, N. Gottschlich, G. Knebel, Chem.<br />

Eng. J. 101 (2004) 447.<br />

[123] S. Arscott, S. Le Gac, C. Druon, P. Tabourier, C. Rolando, Sens.<br />

Actuators B 98 (2004) 140.<br />

[124] M. Schilling, W. Nigge, A. Rudzinski, A. Neyer, R. Hergenroder,<br />

Lab Chip 4 (2004) 220.<br />

[125] H. Yin, K. Killeen, R. Brennen, D. Sobek, M. Werlich,<br />

T. van de Goor, Anal. Chem. 77 (2005) 527.<br />

http://www.elsevier.com/locate/trac 255


Trends Trends in Analytical Chemistry, Vol. 25, No. 3, 2006<br />

[126] I.M. Lazar, B.L. Karger, Anal. Chem. 74 (2002) 6259.<br />

[127] T.A. van de Goor, PharmaGenomics (2004) 20.<br />

[128] G.A. Schultz, T.N. Corso, S.J. Prosser, S. Zhang, Anal. Chem. 72<br />

(2000) 4058.<br />

256 http://www.elsevier.com/locate/trac<br />

[129] S. Zhang, C.K. Van Pelt, D.B. Wilson, Anal. Chem. 75 (2003)<br />

3010.<br />

[130] S. Zhang, C.K. Van Pelt, J.D. Henion, Electrophoresis 24 (2003)<br />

3620.

Hooray! Your file is uploaded and ready to be published.

Saved successfully!

Ooh no, something went wrong!