ATMOSPHERIC PRESSURE ION SOURCES Thomas R. Covey,*,{ Bruce A. Thomson,{ and Bradley B. Schneider{ MDS Analytical Technologies, Sciex, Concord, Ontario, Canada L4K 4V8 Received 15 October 2008; received (revised) 15 January 2009; accepted 15 January 2009 Published online 22 July 2009 in Wiley InterScience (www.interscience.wiley.com) DOI 10.1002/mas.20246 This review of atmospheric pressure ion sources discusses major developments that have occurred since 1991. Advances in the instrumentation and understanding of the key physical principles are the primary focus. Developments with electrospray and atmospheric pressure chemical ionization and variations encompassing adaptations for surface analysis, ambient air analysis, high throughput, and modification of the ionization mechanism are covered. An important and limiting consequence of atmospheric pressure chemical ionization, chemical noise, is discussed as is techniques being employed to ameliorate the problem. Ion transfer and transport from atmospheric pressure into deep vacuum is an area undergoing constant improvement and refinement so is given considerable consideration in this review. # 2009 Wiley Periodicals, Inc., Mass Spec Rev 28:870–897, 2009 Keywords: atmospheric pressure ion source; atmospheric pressure chemical ionization; electrospray; ion transport; ion guides; ion focusing; liquid chromatography; mass spectrometry I. INTRODUCTION This overview begins where the last one published on this topic left off. Titled ‘‘Mass Spectrometry with Ion Sources Operating at Atmospheric Pressure.’’ Bruins (1991) summarized the understanding and developments in this field at a point in time very close to the first commercial introduction of atmospheric pressure LC/MS/MS instrumentation. Figure 1 is a histomap illustrating the expansion of the utilization of atmospheric pressure ionization techniques soon after this occurred. The map also provides a perspective on the rise and fall of most LC/MS ion sources developed over this course of history, their relative importance at any point in time, and the relationship between the different research groups. Reference is given to some of the historical reviews of those LC/MS developments from which information for this histomap was derived, as well as the authors personal experience (Thomson, 1998b; Abian, 1999; Gelpi, 2002; Willoughby, Sheehan, & Mitrovich, 2002). Since 1989 the literature has exploded with applications in areas as diverse as elemental speciation, pharmaceutical drug discovery and development, and protein sequencing. It is not the intent of this overview to image this vast landscape of applications that have come to light over the past 18 years. However, several of the papers in this series of MS Reviews as well as some additional ———— { Principal Scientist. Research Scientist. *Correspondence to: Thomas R. Covey, MDS Analytical Technologies, Sciex, 71 Four Valley Drive, Concord, Ontario, Canada L4K 4V8. E-mail: tom.covey@sciex.com { Mass Spectrometry Reviews, 2009, 28, 870– 897 # 2009 by Wiley Periodicals, Inc. books and reviews could serve as a starting point for the interested reader (Snyder, 1995; Cole, 1997; Lee, 2002, 2005; Pramanik, Ganguly, & Gross, 2002; Hopfgartner & Bourgogne, 2003; Boyd, Basic, & Bethem, 2008). It is the intent to focus on the important instrumentation developments and understanding of the physical principles that have occurred since the Bruins review 18 years ago. Atmospheric pressure ionization (API) instrumentation has evolved since 1991 in conjunction with a deeper understanding of the physical processes involved with ionization and ion transport. Never the less it is fair to say that the foundation principles as summarized by Bruins have not changed substantially. Many variations to atmospheric pressure chemical ionization and electrospray ionization have been spawned from the original concepts, resulting in improvements to better meet the needs of specific applications. Improvements in ionization efficiencies under certain conditions, as well as more effective means for transferring and transporting gas phase ions from the atmospheric region into the deep vacuum of the mass analyzer, have occurred. However, radically new modes of ionization have not yet been discovered and some of the key limitations to ion transfer and transport remain as obstinate barriers. The vast majority of the developments in atmospheric pressure ion sources that have achieved wide acceptance and commercial importance have been focused on systems where the samples are introduced in a liquid stream with the primary motivation being the coupling to liquid chromatographs. Improvements to liquid introduction atmospheric pressure chemical ionization (APCI) and electrospray ionization (ESI) sources have dominated the developments over the past 2 decades and will be the first topics considered before delving into issues involving the transport of ions into the vacuum chamber and transfer into the mass analyzer. Also reviewed are the plethora of ion source developments that have occurred during this period, many of which deal with surface analysis, solid sample introduction, or adaptations to achieve high throughput. Fundamentally these concepts can be viewed as variations on the original APCI and ESI themes. II. ELECTROSPRAY ION SOURCES The most widely used ionization technique in mass spectrometry today is ESI wherein ions are created by electrically charging a flowing stream of liquid at atmospheric pressure, resulting in the emission of molecular ions from the droplets in the subsequent spray. Since the works attempting to adapt electrostatic spray techniques to mass spectrometry in the early 1980s, the field has diverged in two general directions, high and low flow rate liquid introduction systems. The foundation principles of ion evaporation (Thomson & Iribarne, 1978, 1979) and electrospray ATMOSPHERIC PRESSURE ION SOURCES & FIGURE 1. A history of the development of LC/MS instrumentation as viewed in histomap format. The ‘‘y’’ axis is year and the ‘‘x’’ axis is the relative adoption of the technique at that point in time estimated by the extent of commercial adaptation as well as publications and activity in academic research institutions. The concept of a histomap was borrowed from a Rand McNally publication (Sparks, 1952). The year 1989 marks the time of the first commercial API ESI mass spectrometer (Sciex API 3). Names and dates refer to specific citations in the Reference section. (Whitehouse et al., 1985) set the stage for the diverging paths. The electrospray work served as the nucleus for developments in the submicroliter per minute flow regime referred to as nanoelectrospray (nano-ESI) (Wilm & Mann, 1994). The ion evaporation work pointed the way forward into the milliliter per minute regime which eventually gave way to the now ubiquitous pneumatically assisted electrospray or ion spray (Bruins, Covey, & Henion, 1987) which is a nomenclature resulting from a condensation of the ion evaporation and electrospray terms. The historical timelines of the development and relationships between these techniques are captured in Figure 1. The motivation for operating in a pure electrospray mode at low liquid flow rates is to achieve as high absolute sensitivity as possible (Wilm & Mann, 1994). At low liquid flow rate the efficiency of ion creation and transfer into the vacuum system of a mass spectrometer is optimal because small droplets are easier to evaporate than large droplets, it is easier to create and charge small droplets from low liquid flows than from high liquid flows, and it is easier to direct the resulting ions and solvent vapors into the vacuum system of a mass spectrometer. The charging of the liquid and the process of droplet generation are accomplished in a single step with the electrospray process. This involves the formation of a Taylor cone by high electric fields that emit Mass Spectrometry Reviews DOI 10.1002/mas droplets from the apex where the field is high enough to overcome the liquid surface tension forces. The droplets initially produced can be readily evaporated with gentle heat over short distances at atmospheric pressure. The entire plume of ions generated from the evaporating spray can be inhaled by the mass spectrometer with the viscous drag forces of the vacuum system so that few ions or clusters are lost to collisions with the walls in the source or vacuum interface. Under carefully controlled conditions, at flows in the tens to hundreds of nanoliters per minute range, the transfer of ions from the solution phase to the first vacuum stage has been observed to be greater than 50% but can vary over two orders of magnitude depending on the liquid emitter geometry, flow rates, solvent surface tension, and physical positioning of the sprayer. With all parameters properly balanced, sensitivities can be very high and this approach can serve as an interface to LC under the narrowly defined stability conditions required at these flow rates. The motivation for exploring the pneumatic variants of electrospray or ion spray was born out of a desire to develop a general purpose interface to liquid chromatography where operation over a liquid flow range of tens to hundreds of microliters per minute, and a tolerance to changes in mobile phase composition, are important. Most commercial HPLC instrumentation is geared toward this flow regime because of the simplification of construction, fluid flow control, and user 871 & COVEY, THOMSON, AND SCHNEIDER operation compared to instrumentation designed for nanoliters/ min flows. Given that, in the majority of cases sample concentration detection limits are the most important analytical metric and sample volumes are not limited, HPLC systems with large sample injection capacity are desirable. Since sample injection capacity scales with the square of the column diameter, the absolute sensitivity losses obtained at higher flows can be compensated for by injecting more sample, often resulting in lower sample concentration detection limits with high flow systems. To minimize the ion losses and stabilize operation over a wide range of mobile phase conditions the formation of droplets by the combination of gas shear and electrical forces was borrowed from the ion evaporation work. The process of ion emission from cloud droplets was the original incentive to study ion evaporation where droplets are charged by statistical or shear forces. Nebulizers were implemented to create an experimental platform mimicking natural droplet creation such as in cloud formations or in the cascade of a waterfall. The natural statistical charging of droplets was enhanced with an induction electrode which did not make physical contact with the liquid. During the course of these studies the observation of labile organic ions, some multiply charged, indicated that the technique had potential as an analytical tool (Thomson, Iribarne, & Dziedzic, 1982). However, the relatively low sensitivity demonstrated from this configuration stifled interest from the general community. Nonetheless it provided the basis for the concept of separating the charging from the droplet generation process which is a key element of the high flow sources. Incorporating direct electrical charging of the liquid with the proper voltage isolation resulted in significant sensitivity improvements over induction based approaches. This would form the basis of the ion spray interface where the initial pneumatically driven droplet generation step is decoupled from the charging step and the careful balance between voltages, flow rates, mobile phase compositions, aperture dimensions, and physical position becomes relaxed. What both the electrospray and ion evaporation techniques demonstrated was that gas phase ions of very labile organic compounds could be created from charged droplets at atmospheric pressure. The appearance of these atmospheric gas phase ions, as indicated by the presence of multiple charges (Covey et al., 1988; Fenn et al., 1989), reflect their solution state chemistry at the time of emission. What they both have in common is a requirement to create droplets that have a net charge, and the diameter of those droplets must be rapidly reduced to the Rayleigh limit where coulomb explosions will lead to the final cascade of droplet diameter reduction and ultimate release of ions and clusters. The process of going from macroscopic droplets to fully desolvated ions is not understood in all details. The main theories that attempt to explain the creation of single ions from droplets that contain hundreds to thousands of analyte ions have been well described and debated in the literature (Kebarle & Tang, 1993; Analytica Chimica Acta, 2000, entire volume; Gamero-Castano & Fernandez de la Mora, 2000). The main theories (the ion evaporation theory and the charged residue theory) end with ions in the gas phase that presumably contain a number of attached solvent molecules. The important point is that the high and low flow techniques share the same mechanism of ionization, share the same means of droplet charging, but they differ in the means by which the droplets are initially created. 872 Some of the directions that these two approaches have taken over nearly the past 2 decades will be described. As we will see, significant effort has been directed at evaporating as much of the spray as possible, and then sampling as many of the ions as possible into the vacuum chamber. A. Nanoelectrospray: Low Flow Rates (1–1,000 nL/min) The starting point of nano-ESI developments can be traced to the seminal works of Wilm and Mann (1996) in the mid 1990s, who described the physical principles underlying electrospray operating at nanoliters/min flows and demonstrated the practicality of the technique for protein identification after 2-D gel electrophoresis separations (Wilm et al., 1996). This original work used a commercially available triple-quadrupole instrument (API 3) with an unmodified API source. The interface on this instrument was unique by today’s standards in that the transition from atmosphere to deep analyzer vacuum was conducted in a single stage through a 125 mm pinhole aperture. This required pumping speeds on the order of 100,000 L/sec which could only be achieved by compressed helium cryogenic pumps that froze the incoming gases on cryopanels surrounding the ion optics, and thus required periodic thawing and recycling. Their work on this system indicated that there were two critical parameters to control which are as important today on modern multistage turbomolecular pumped instruments as they were back then. The first is regarding sensitivity and the full utilization of vacuum drag forces to inhale most if not all of the ions produced at atmospheric pressure. This aspect will be covered in Desolvation and Declustering Section. The second critical parameter they identified was that the nano-ESI emitter geometry is critical for achieving stable ion currents at low liquid flow rates. An explosion of research in academic and industrial institutions occurred around emitter geometries, materials, and fabrication techniques. Today this field has largely matured such that nanoESI emitter design and fabrication has advanced into the realm of commercial mass production resulting in reproducible and high performance sprayers. Short of an exhaustive review of all emitter geometries reported, the key principles that provided the foundations for current popularized designs will be discussed. 1. Relationship Between Optimal Flow Rate and Emitter Geometry The essence of the nanoflow methodology is to reduce the flow rate of the sprayed sample to the sub microliter per minute regime so that the entire vaporized spray can be inhaled into the mass spectrometer. The efficiency of the ionization and ion transfer process improves approximately proportional to the flow rate reduction. The flow at which maximum efficiency is reached depends on the characteristics of the atmosphere to vacuum interface and the sprayer geometry. The former will be discussed in the ion transport section and the latter in this section. Factors affecting the formation and stabilization of the Taylor Cone are of paramount concern and are greatly influenced by the relationship between the electrospray tip geometry and liquid flow rate as well as other factors such as solvent viscosity and conductivity. Mass Spectrometry Reviews DOI 10.1002/mas ATMOSPHERIC PRESSURE ION SOURCES The theoretical derivations by Wilm and Mann for nanoESI, as well as the studies published in the general physics literature on Taylor Cone properties (Ganan-Calvo, Davila, & Barrero, 1997; Ku et al., 2001), showed a relationship between the liquid flow rate and the diameter of the droplets emitting from a Taylor Cone. At low nanoliters/min flows, these droplets were calculated to be in the low to sub micron range, small enough to result in ion emission with very little time or distance required for further desolvation. This insight into the effects of low liquid flow rates led them to construct electrospray needles with 1–3 mm exit apertures that delivered flows on the order of 10–50 nL/min, and could thus be positioned in close proximity to the entrance aperture of the mass spectrometer maximizing the collection of ions, and posing no danger of overwhelming the vacuum system with liquid vapors. However, the theory provided little guidance as to what these electrospray needles should look like. Inspired intuition directed this and other groups working on the same problem, to experiment with glass and fused silica capillaries drawn to different dimensions and methods for delivering the electrical charge. Fortunately a large body of literature already existed describing techniques for making similar electrodes for use in measuring electrical potentials across cell membranes. For an in-depth treatise on the science and history behind microelectrode technology see Brown and Flaming (1992). A large number of publications began to emerge rapidly, optimizing tip configurations for specific applications such as infusion, capillary electrophoresis/MS (CE/MS), and capillary HPLC/MS. A common theme among them was the investigation of the various aspects of the physical architecture of these nanoflow electrodes and the effect that it has on their operational behavior, of primary interest being the lowest flow rate a particular size and shape electrode could support without compromising stability of the ion beam. The picture has developed into a complex interrelationship between tip inner diameter, outer diameter, channel taper to the tip, and solvent composition, all exerting some influence on the lowest attainable flow with a particular nanoflow electrode. The various researchers in this area derived their preferred combination of these parameters to achieve the operational results they desired for their application. Some guidelines will be put forth below, based on these authors’ experience and the collective considerations of the literature studies regarding one crucially important feature of electrode geometry as it relates to flow rate, the diameter of the emitter at the exit. A comprehensive list of the many emitter geometries and configurations adapted for various applications and flow rates is provided here (Emmett & Caprioli, 1994; Wahl, Gale, & Smith, 1994; Valaskovic & McLafferty, 1995, 1996a,b; Kriger, Cook, & Ramsey, 1995; Davis, Stahl, & Lee, 1995; Figeys et al., 1996; Valaskovic, Kelleher, & McLafferty, 1996; Bateman, White, & Thibault, 1997; Cao & Moini, 1997; Davis & Lee, 1997; Kelly, Ramaley, & Thibault, 1997; Vanhoutte et al., 1997; Gatlin et al., 1998; Geromanos et al., 1998; Hannis & Muddimann, 1998; Vanhoutte, Van Dongen, & Esmans, 1998; Wang & Hackett, 1998; Fong & Chan, 1999; Juraschek, Dulcks, & Karas,1999; Barnidge, Nilsson, & Markides, 1999; Barnidge et al., 1999; Barroso & de Jong, 1999; Hsich et al., 1999; Feng & Smith, 2000; Gatlin et al., 2000; Geromanos, Freckleton, & Tempst, 2000; Covey & Pinto, 2002; Schmidt, Karas, & Dulicks, 2003). Mass Spectrometry Reviews DOI 10.1002/mas & It has been generally observed that, as both the inner diameter (i.d.) and outer diameter (o.d.) of the tip at the exit get smaller, stable sprays can be maintained at lower flows. Amongst the lowest flows published were in the 0.5–5 nL/min range using 0.8 mm i.d. apertures (Geromanos et al., 1998). Figure 2 shows some results from a flow rate versus aperture i.d. experiment that depicts what is generally observed. A strong relationship between tip inner diameter and lowest sustainable flow is seen (Covey & Pinto, 2002). The inflection point in the graphs is referred to as the ‘‘optimum flow.’’ This optimum flow is the highest flow where ionization efficiency remains at its peak for that electrode. At flows below this the ion current begins to decrease with flow exhibiting a mass-flux sensitive response (i.e., the ion abundance varies with the rate at which sample molecules are introduced into the ion source). Here the ionization efficiency remains optimal but the mass flux into the sprayer decreases resulting in a proportionally decreasing signal. This linear response range occurs over a very narrow range of flows and is soon disrupted when the spray becomes unstable at flows that are too low. At flows above the optimal flow the ionization efficiency decreases but the signal remains relatively constant as the mass flux into the sprayer increases. In this range the response appears to be similar to a concentration sensitive detector, such as a UV detector, but this is not a true concentration sensitive response. It is an anomaly resulting from the changing ionization efficiency over this part of the flow range and is better described as a pseudo concentration sensitive response. Several possibilities can be proposed to explain this observed relationship between aperture size and optimum flow rate. Central to this understanding is a consideration of all factors affecting the size, shape, and stability of the Taylor cone, which has been studied in great detail (De la Mora & Loscertales, 1994; Cloupeau & Prunet-Foch, 1990). An explanation of this specific case, which reflects general observations from the field of nanoESI, is that optimal flow is the minimum flow where the base of the Taylor cone just bridges the emitter i.d. and maintains stability firmly anchored to the walls of the electrode. At flows much below this the cone becomes unstable as it begins to collapse inside the lumen of the emitter. This is the mass sensitive region which occurs over a very small flow range before instability takes over. At flows above the optimum the Taylor cone can remain stable over a considerable flow range as the base of the cone spreads over the outside and along the length of the electrode. With higher flows the growing Taylor cone emits larger droplets that decrease the ionization efficiency. This is the pseudo concentration sensitive response region. With increasing flow in the overflow state pneumatic nebulization is eventually required to sustain stable droplet formation. All high flow rate ESI sources operate in this regime. In this regime heat transfer and evaporation rates become increasingly important as will be discussed in more detail in the following section. Figure 2 is a reasonable portrayal of the general trends to expect with this relationship between i.d., flow rate, and ionization efficiency. However, it should be kept in mind that there are other electrode architectural features involved in establishing the optimum flow rate. Among these is the outer diameter of the tip at the exit of the electrode, the taper of the channel leading up to the exit aperture, and the method for delivering the voltage. Superimposed on electrode geometry are 873 & COVEY, THOMSON, AND SCHNEIDER FIGURE 2. Relationship between flow rate and emitter diameter. The ‘‘y’’ axis is intensity of the peptide ion signal (monitored by selected ion monitoring) expressed in relative terms and normalized for each flow rate. Emitters were made from a 1 cm length of 150 mm o.d. fused silica having six different i.d.’s. Samples were infused with a syringe pump to control the flow and the voltages were varied to maintain maximum sensitivity and stability. The solvent composition and sample concentration remained the same throughout. The arrows point to the ‘‘optimal flow’’ for each electrode. Reprinted with permission from Covey and Pinto (2002), copyright 2002, Marcel Dekker and Thomson (2007b), copyright 2007, Elsevier. solvent viscosity and conductivity, all of which will have an affect on determining the precise optimal flow for a particular emitter. Nevertheless the data in Figure 2 are a reasonable first approximation for most electrodes used today. 2. Electrospray Emitters: Three Basic Configurations The construction of electrospray emitters is guided by the desired application of the device. They fall into three general categories; those designed for bulk flow sample introduction, those designed for coupling to continuous flow sources such as liquid chromatographs, and those constructed using micromachining principles with applications for both bulk and continuous flow sample introduction. Electrodes in all three categories are commercially available. Bulk and continuous flow emitters are available from companies such as New Objective, Woburn, MA; Proxeon, Odense, Denmark; and Phoenix S&T, Chester, PA. Various configurations of micromachined devices are available from Advion Biosciences, Ithaca, NY; Agilent Technologies, Santa Clara, CA; and Waters Corporation, Milford, MA. a. Bulk flow electrodes. Bulk flow or discrete sample introduction electrodes (i.e., those not involving chromatographic separations) are typically drawn from a borosilicate glass tube of approximately 1 mm o.d 0.8 mm i.d. with apertures ranging from 1 to 5 mm. Figure 3 is a series of photos taken with various magnification of such an electrode made by New Objective with dimensional control of the aperture by an HF etching procedure. 874 In this case electrical contact is made through vapor deposited metal layers stabilized with overcoatings of SiOx. Another means of securing the vapor deposited metallic coating uses a bifunctional organosilane undercoating (3-mercaptopropyl) (Kriger, Cook, & Ramsey, 1995). Yet another approach for establishing a permanent voltage contact is called the ‘‘fairy dust’’ technique (Barnidge, Nilsson, & Markides, 1999) created by gluing gold particles to the electrode. For applications where the emitters are FIGURE 3. Photograph of a bulk flow emitter at four different magnifications. Reprinted with permission from Covey and Pinto (2002), copyright 2002, Marcel Dekker, and New Objective, Inc. Mass Spectrometry Reviews DOI 10.1002/mas ATMOSPHERIC PRESSURE ION SOURCES disposable the fragile metal layer is sometimes not secured. A non-metalized electrode has also been reported using a tungsten wire inserted into the back of the capillary (Van Berkel, Asano, & Schnier, 2001), and this was primarily used to study the electrochemical properties of ESI (Van Berkel and Kertesz, 2007b). These studies showed that unwanted analyte oxidation and reduction reactions can occur if there is an excessive residence time in the vicinity of the high voltage electrode surface. b. Continuous flow electrodes. The use of fused silica tubing available with a wide variety of precise inner bores has become popular for coupling to, or integrating with, capillary HPLC or CE columns. Exit i.d.’s are tailored to a desired flow rate range by drawing the fused silica with a laser-heated pulling apparatus. Electrical contact is made with similar metalizing techniques as used for the bulk flow electrodes or through the use of a conductive union. Figure 4A is a picture of an integrated single piece column/emitter system. Alternatively, specialized zero dead volume unions where voltage can be applied at the point of column/emitter junction have been developed commercially. One example of this is shown in Figure 4B. Continuous flow emitters have also been implemented with pneumatic nebulization to extend the flow range capability to the low microliters/min regime (Schneider et al., 2005). c. Micromachined electrodes. Microfluidic systems, generally defined as those where channels and other structures are etched in bulk substrates have demonstrated utility for nano-ESI applications. Arrays of emitter nozzles have been created in silicon FIGURE 4. Continuous flow emitters for nano-LC coupling. A: Pulled fused silica emitter with integrated LC stationary phase and frit. B: Zero dead volume fitting to couple a 75 mm i.d. capillary LC column and electrospray emitter. Voltage can be applied at the junction with an embedded wire (wire not shown). Reprinted with permission from New Objective, Inc. [Color figure can be viewed in the online issue, which is available at www.interscience.wiley.com.] Mass Spectrometry Reviews DOI 10.1002/mas & substrates with reactive plasma etching techniques to enable automation of the emitter tip replacement process. Figure 5A is a photo of such a device made by Advion using reactive plasma etching in a silicon substrate. This technique is used to produce arrays of nozzles in chip format to enable rapid tip replacement (Schultz et al., 2000). These emitters have been used for both bulk flow sample introduction or for coupling to chromatographic systems. Chip based integrated column/sprayer devices have also been developed commercially by Agilent and Waters. B. Pneumatic Electrospray: High Flow Rates (1–1,000 m L/min) Since the 1991 Bruins review on atmospheric pressure sources the ion spray interface has grown widely in popularity and dominates the application field of LC/MS. Pneumatic nebulization of a charged liquid stream will generate a stable ion current over a broad flow rate range and provide a response dynamic range of 103 –104 (Kostiainen & Bruins, 1994) while remaining relatively insensitive to changes in voltage, mobile phase composition, sprayer position, and electrode dimensions. Other mechanical means for droplet creation, such as ultrasonic nebulization, have also been demonstrated (Banks, Quinn, & Whitehouse, 1994) but the pneumatic approach remains the most widely used because of its simplicity and performance. It has also found use as a supercritical fluid chromatography interface (Baker & Pinkston, 1998). 1. Nebulizers For some insight into the mechanism of initial droplet production with a pneumatic nebulizer for electrospray devices, a first approximation of the physical force available to disperse the liquid is in order. The majority of the nebulizers used today for this purpose achieve gas velocities from 100 to 300 m/sec (Mach 1) over distances of less than 3 mm from the exit of the gas nozzle. It is worth noting that to accelerate from zero to 125 m/sec in this distance implies an average acceleration of 340,000 g. The corresponding turbulent aerodynamic force at the liquid gas interface in the first few millimeters of distance is more than sufficient to shatter any liquid at these flows into populations of charged drops ranging in dimensions from submicron to tens of microns in diameter. A photograph of this highly energetic region in the first few millimeters from the exit of a pneumatic nebulizer is shown in Figure 6. After the mechanical disruption, columbic forces take over the droplet size reduction process for the smaller droplets and rapid heating is required to assist this process for the larger ones. Further details of this nebulization process and its operating characteristics have been described elsewhere (Covey, 2007). This decoupling of the charging and initial droplet generation process with common pneumatic nebulizers leads to a usable voltage range that is quite broad. It can be so broad that at high liquid flow rates (1 mL/min) and maximum gas flows essentially no voltage needs to be applied, a condition reminiscent of the ion evaporation device. The high gas shear velocities create a situation where statistical charging of the liquid provides sufficient charge. A low voltage, on the order 875 & COVEY, THOMSON, AND SCHNEIDER FIGURE 5. Micromachined electrospray nozzles. A: Electron micrograph of a nozzle in a silicon substrate with a 10 mm inner diameter aperture. Reprinted with permission from Advion and reference Thomson (2007b), copyright 2007, Elsevier. B, C, D: Successive magnifications nozzles. The inner diameter of the nozzle in D is 10 microns. [Color figure can be viewed in the online issue, which is available at www.interscience.wiley.com.] of 50–100 V instead of several kV, is useful to establish an ion drift potential toward the inlet aperture but does little to effect charging. This operational regime has been observed for several years on systems with heated pneumatically nebulized electrosprays and employed for some drug assays (Bi et al., 2000) but is often overlooked because it provides no particular sensitivity or operational advantage over the use of high voltage for these types of applications. Recently it has been observed that operating a nebulizer in this fashion leads to reduction in protein charge states and will produce atmospheric gas phase ions of both polarities simultaneously which has interesting utility in some areas (Hirabayashi et al., 1994; Takats et al., 2004). Similar phenomena appear to be operating with other low voltage, high shear systems (Cristoni et al., 2005). 2. Jet Dynamics and Heat Transfer While the early focus on these systems was on the importance of high velocity gases, later attention during the mid 1990s began to be focused on understanding the processes involved in the slowing down of these jets and how to take advantage of these phenomena. Laser Doppler anemometry and analytical modeling were the primary investigative tools. The peak FIGURE 6. Schlieren photograph of a jet of air issuing from a nebulizer nozzle at sonic velocity, showing the Mach structure extending a few mm from the nozzle. Reproduced with permission from Covey (2007), copyright 2007, Elsevier, and from Issac (1994), copyright 1994, with permission of the American Institute of Aeoronautics and Astronautics. 876 velocities mentioned above are achieved within the first few millimeters of the expansion nozzle. After this the jet continuously decelerates to speeds of approximately 10–30 m/ sec 2–3 cm from the nozzle. The velocity profile across the jet is Gaussian, with higher speeds and larger droplet diameters near the center line and a large population of small micron and submicron sized droplets in the periphery (Covey, 2007). The region close to the jet boundaries is of particular importance. This is where droplet size is smallest and density of these small droplets is the highest. It is also at this point that the momentum transfer from the surrounding air to the jet occurs, the overall system conserving momentum. As the jet slows down there is an increase in the mass of external air being drawn in or entrained at a lower velocity. The trajectory of the gas flow streamlines are approximated in Figure 7 from the results of the analytical models. This phenomenon is commonly encountered in a shower stall with a flexible plastic curtain. When the shower spray is turned on, the curtain is annoyingly drawn inward toward the bather. Analytical models of the gas dynamics of these jets, verified by measurements of the velocities of the gas at the jet boundary, indicate that the entrainment ratios are greater than 20:1. For every liter per minute of nebulizer gas consumption, >20 L/min of external air is drawn into the body of the jet. This presents an excellent opportunity to transfer a large amount of heat to the droplets for efficient and rapid desolvation and flow rate extension. With 10–20 L/min of 8008C entrainment air injected into the primary jet, 1 mL/min of water will completely desolvate within a few cm. Injection of clean air will also minimize the inclusion of background contaminants, eliminate recirculation effects that occur in closed chambers, and reduce the radial expansion of the plume. At first consideration temperatures of 8008C seem alarming, with the assumption that thermal degradation will occur. However this is not observed to be excessive for good reason. Analytical models have been developed that predict that the temperature of the liquid in the droplet under these conditions does not rise much above ambient conditions (French, Etkin, & Jong, 1994). Evaporative cooling balances the rate of heat transfer into the drop. In a typical calculation a droplet in 7508C Mass Spectrometry Reviews DOI 10.1002/mas ATMOSPHERIC PRESSURE ION SOURCES & FIGURE 7. Entrainment gas streamlines with a fully developed, unconfined nebulizer jet. Reprinted with permission from Covey (2007), copyright 2007, Elsevier. air will reach a steady state temperature of 448C. Upon reaching emission diameters, typically considered to be around 10 nm or less, the ion will leave the droplet as a cluster or with a small solvation sphere. It will reside in the hot gas for a short period of time before being drawn into the vacuum system, typically a few tens to hundreds of microseconds. Thermally labile compounds, such as nucleotide phosphates, generate predominantly molecular ions under these circumstances. In a heated nebulizer APCI source, where vaporization occurs as a result of heat transfer during droplet impact with a hot surface at similar temperatures, these same compounds are completely thermally degraded highlighting the radically different mechanisms involved in the desolvation and analyte liberation processes with these two different high temperature techniques. 3. Sampling Efficiencies One means of quantifying the sensitivity of an ion source is to measure the sampling efficiency (El-Faramawy, Siu, & Thomson, 2005; Schneider, Javaheri, & Covey, 2006). The sampling efficiency is the fraction of analyte molecules introduced in solution into the ion source that are captured as ions by the initial ion optics in the deep vacuum of the mass spectrometer. It is the product of the ionization efficiency (efficiency of ion creation at atmospheric pressure) and transfer efficiency (efficiency of transferring ions from atmosphere into vacuum). Figure 8 pictorially depicts this concept for the high and low flow electrospray systems described. When this measurement of sampling efficiency was applied across the flow range on a high sensitivity commercially available triple quadrupole mass spectrometer, the results in Figure 9 were obtained (Covey et al., 2009). The difference in sampling Mass Spectrometry Reviews DOI 10.1002/mas efficiency between highest and lowest flows is approximately 100-fold. Referring back to the earlier comments on absolute versus sample concentration detection limits, in situations where the absolute amount of material to be analyzed is limited, such as sequencing peptides extracted from 2-D gels, nano-ESI has a large advantage. In cases of trace analysis where the sample volume or mass is not limited, such as drug quantitation in blood plasma, the advantage lies with the high flow system where the injection capacity of the HPLC system compensates for the loss in efficiency and more routine operation is afforded. Another important observation to be made from these data is that sensitivity improvements with nano-ESI systems are approaching their theoretical limits. As mentioned in the Transport of Ions From Atmosphere to Vacuum Section, with slight modifications to commercial instrumentation efficiencies as high as 80% have been achieved and reproduced. Significant gains remain to be harnessed with high flow systems, however the physics of this region of the mass spectrometer is highly complex, dominated by multiple high velocity gas jets, space charge, and a medium (atmospheric pressure) where electrostatic focusing is difficult (see Focusing Ions at Atmospheric Pressure Section). The underlying predictions of the behavior of multiple intersecting fluids and gases in the turbulent regime lie in the solutions to the Navier–Stokes equations for fluid flow, which is one of the seven Millennium Problems in mathematics resisting a general solution over the years despite its massive practical importance to many fields (Girvan, 2003). Computational fluid dynamics working in close cohort with empirical experimentation may bring important advances. Experiments with an air amplifier on focusing low flow rate electrospray plumes (Zhou et al., 2003) have shown some initial indications that further improvements in focusing droplets at atmospheric pressure with gas flow may be possible. 877 & COVEY, THOMSON, AND SCHNEIDER FIGURE 8. Sampling efficiency comparison. The sampling efficiency is the ratio of the number of analyte molecules present in solution in the sprayer to the number of ions that enter the RF multipole ion guide in the vacuum. It is the product of the ionization and transfer efficiency. A: Corona discharge needle with APCI. B: Low flow ESI system. FIGURE 9. Sampling efficiency versus flow rate. The data in the upper dashed trace (black), read from the right Y axis, is an expansion of the data in the lower solid trace (red) read from the left Y axis. For each flow the optimum source and interface configuration was used for those conditions. For high flow rates above 1 mL/min a heated TurboIonSpray1 source with a conical pinhole aperture interface (standard interface) on an API 5000TM triple quadrupole instrument. For points below 1 mL/min a nanospray source with a 15 mm aperture fused silica capillary and an interface optimized for nanoflow introduction (PDI interface, See Ion Transport Section, Declustering, PDI interface) was used. 878 Mass Spectrometry Reviews DOI 10.1002/mas ATMOSPHERIC PRESSURE ION SOURCES III. ATMOSPHERIC PRESSURE CHEMICAL IONIZATION SOURCE APCI mass spectrometry was commercially developed many years before electrospray ionization (ESI), but remained a niche technique until the explosion of interest in LC/MS for biological analysis. APCI/MS was used primarily for air analysis because of the high sensitivity afforded toward polar volatile and semivolatile species (French et al., 1984). The analysis of airborne substances and thermally desorbed surfaces and solids with hot gases was the primary application of atmospheric pressure chemical ionization sources in the 1980s. A common demonstration of this method of sample introduction was the direct detection of volatiles emanating from solid substances as shown in Figure 10. A carrier gas sweeps over the sample bringing ambient volatiles into the source. The instrument is a triple-quadrupole called the trace atmospheric gas analyzer (TAGA), some of which were installed in vans as mobile laboratories for real-time monitoring and tracking of industrial pollutants (see Lane & Thomson, 1981 for a report on an emergency-response application using an earlier single-quadrupole version). This technology also was the foundation for the AROMIC system, a real-time cargo examination system developed in the mid-1980s that used APCI with MS/MS for contraband detection. While these systems had some commercial success, there was no single large application for APCI to carry it into the mainstream of analytical mass spectrometry which was largely GC/MS based By the time of the Bruins review in 1991, the interest in atmospheric pressure ionization had shifted largely toward liquid introduction systems, driven by electrospray rather than APCI as the ionization method. Since that time, however, interest in APCI and direct analysis has re-emerged with a focus on the rapid analysis of substances thermally desorbed from solid surfaces. This will be discussed further in the section of this article describing variations on the APCI theme. & A. Ionization The foundation principles of APCI can be found in the early literature on low pressure chemical ionization and are well understood (Harrison, 1992; Lias & Bartness, 2008). APCI is a gas phase ionization process as opposed to the liquid phase ionization process of ESI. Ionization of the volatilized neutral analyte will only occur if it has sufficient gas phase basicity or acidity to extract or donate a proton from the reagent ion population present in great excess (French et al., 1984; Bruins, 1991). In the case of LC/MS the solvents are the dominant reagent ions but they can be carefully selected to enhance either positive or negative ionization (Schaefer & Dixon, 1996) Adduct ion formation is also a possible channel for ionization with ion molecule reactions. Other processes such as charge transfer and electron capture can be forced to occur given the correct analyte and reagent ion population conditions to yield very high sensitivities (Singh et al., 2000). One glaring difference between liquid phase and gas phase ionization is the absence of multiply charged ions. It is both thermodynamically and kinetically unfavorable for two or more reagent ions to react with the same neutral molecule within a short period of time. Both charge repulsion and collision probabilities prevent this from occurring. One possible exception would be chemical ionization of massive molecules such as proteins where this becomes a possibility. The occasional observation of minor doubly or triply charged protein ions in a MALDI spectrum may be the result of a multiple-event chemical ionization reaction. Ionization by APCI is an indirect process where an initial source of ions is used to create ions from the background gas. At atmospheric pressure the mean free path between collisions is short so a large number of gas phase reactions occur very quickly and come to an equilibrium state where the most stable species predominate and serve as the reagent ions. The high density of reagent ions results in very high ionization efficiencies, nearly FIGURE 10. APCI ambient air monitoring interface. A carrier gas sweeps over the sample in the glass inlet transporting the volatiles into the APCI source of a TAGA 6000 triple quad instrument. This photograph was taken in 1984. [Color figure can be viewed in the online issue, which is available at www. interscience.wiley.com.] Mass Spectrometry Reviews DOI 10.1002/mas 879 & COVEY, THOMSON, AND SCHNEIDER all analyte molecules are ionized if the reaction chemistry is favorable. These final reagent ions are typically protonated (positive ions) or deprotonated (negative ions) resulting in (M þ H)þ or (M H) analyte molecular ions or the formation of adducts. If great care is taken to exclude all proton donating or abstracting species (ubiquitous water being the main source of these), electron donating or abstracting reagent ions can be used resulting in radical cations or anions, similar to those that would be observed from a direct ionization source such as a low pressure electron ionization source. On occasion analyte ions will be directly ionized by the primary source of ions, such as the corona discharge itself, but at atmospheric pressure those ions are lost in subsequent reactions so the final products are dominated by the rules of gas phase ion chemistry. It is very difficult to achieve and maintain direct ionization at atmospheric pressure and at the same time have high efficiency. The most widely used source for the initial ionization event is the direct current corona discharge needle because of its simplicity, robustness, and high efficiency. Many other ionizing devices have been used for different reasons, such as the 63Ni beta emitter commonly used today for field portable instruments (Horning et al., 1974), photons from vacuum UV lamps which operate without high electric fields (Robb, Covey, & Bruins, 2000; Syage, Hanning-Lee, & Hanold, 2000), RF plasma’s (Ratcliffe et al., 2007), crystal discharges for miniaturization (Neidholdt & Beauchamp, 2007), microwave plasmas (Shen & Satzger, 1991; Moini et al., 1998), and lasers (Constapel et al., 2005). The most important sample inlet device for APCI is the heated nebulizer which vaporizes the LC solvent and creates a reagent ion population from this gas with a high voltage corona discharge (Thomson, 2007a). The nebulizer is designed to cause pneumatically generated droplets to collide with hot surfaces to transition the solvent and analyte molecules to neutral gas phase species. The basic principles have not changed radically over the last 20 years, but improvements have been made to enhance performance. These improvements have been primarily centered on improving droplet trajectories, impact dynamics, and desolvation rates. B. Desolvation: Nucleate Boiling The primary problem is one of achieving complete evaporation of the liquid and vaporizing the analyte in as short a time frame as possible, <one millisecond considering spray is emitting from a near-sonic nozzle and traversing only a few centimeters. To achieve this, nebulized droplets are driven to impact hot surfaces. The droplets are evaporated rapidly by conductive heat transfer whereby the droplet spreads on a heated surface, and the analyte is flash vaporized by a process referred to as nucleate boiling (Bernardin, Stebbins, & Mudawari, 1997). This is an efficient means of heat transfer but can lead to thermal degradation as a result of a temporarily dried analyte residing on a hot surface. It can also lead to sample memory effects as a result of high boiling point materials slowly desorbing. The heated nebulizers produced during the 1990’s were of this vintage. C. Nebulization and Desolvation: Vapor Film Boiling More advanced designs have evolved that take advantage of nebulizer jet properties to drive the droplets toward the heated walls of the desolvation chamber. Figure 7 shows the fully developed unconfined nebulizer jet of a heated pneumatic electrospray nebulizer where as much entrainment gas is supplied as the nebulizer demands, allowing an unperturbed forward velocity spray to result. With optimized APCI interfaces the nebulizers are choked by enclosing them in a small finite space, so the only source of gas to supply the entrainment originates from the gas within the space and from the expanding gas jet. The average streamlines then take the form of a closed loop or recirculation eddy. In this region highly turbulent flow FIGURE 11. Drawing of a heated nebulizer desolvation chamber showing the vapor film boiling process and the zoned heaters. Reprinted with permission from Thomson (2007a), copyright 2007, Elsevier. [Color figure can be viewed in the online issue, which is available at www.interscience.wiley.com.] 880 Mass Spectrometry Reviews DOI 10.1002/mas ATMOSPHERIC PRESSURE ION SOURCES predominates and the droplets are driven to the surface of the chamber in a short region downstream of the tip as diagrammatically portrayed in Figure 11. Improvements to the vaporization step have also been implemented by utilizing a vapor film boiling process, instead of a nucleate boiling process, which takes advantage of the Leidenfrost effect (Chandra & Avedisian, 1991; Hatta et al., 1995). Heated nebulizers have evolved to maximize this effect, minimizing direct sample contact with the hot surface. As the droplet approaches a surface of sufficiently high temperature and surface polish a vapor film will grow between them. The vapor pressure will eventually exceed the droplet momentum toward the surface and rebound the droplet before contact, in the process losing approximately 50% of its volume. Depending on the surface roughness, the droplet impact energy, and the impact angle, the droplet will rebound and break up into smaller droplets. Below this temperature the droplet and sample will stick and spread, undergoing nucleate boiling. The design concept of heated nebulizers taking advantage of this effect is also shown in Figure 11. Photographic images of droplets undergoing nucleate and vapor film boiling are shown in Figure 12A,B. & After several repetitive rebounds the initial droplet volume will reduce by several orders of magnitude, resulting in a dry vapor. By maximizing this effect vaporization can occur without sample directly depositing on the surface. Polished ceramic tubes with embedded zoned heaters have been widely employed to serve this purpose. The flexibility to zone or gradient the heat with such heaters allows for temperatures approaching 8008C to be located in the region of initial droplet impact and cooler temperatures during the final stages of vapor drying where the fine droplets have lost their momentum, and are no longer impacting the heater surfaces. Excessive convective heat transfer during this final stage of drying will induce thermal degradation. The Leidenfrost effect can be observed in the kitchen when water droplets are sprinkled on a hot pan (dancing drops). In the case of a near-sonic nebulized spray the Weber numbers (measure of droplet impact energy) are much higher than for droplets impinging by gravitational acceleration alone. Thus heated nebulizer surface temperatures need to be much higher. Despite these efforts, thermal degradation occurs to a significantly greater extent with these interfaces than with thermally assisted electrospray. This is because the neutral FIGURE 12. Time course photographs of nucleate and vapor film boiling starting with initial droplet contact with a surface at time 0.2 msec with final frame at 8 msec. Reprinted with permission from Chandra and Avedisian (1991), copyright 1991, Royal Society. Mass Spectrometry Reviews DOI 10.1002/mas 881 & COVEY, THOMSON, AND SCHNEIDER molecules are driven from the liquid phase to the gas phase entirely by thermal processes. With electrospray the final phase transfer is field induced which is a more mild process. The energetics of the chemical ionization event may also contribute. However this approach using very rapid heating of samples dissolved in droplets produces much less thermal degradation then techniques that thermally desorb solids or samples dried on surfaces. D. APCI and Liquid Flow Rates One of the main attributes of APCI interfaces is the excellent stability of the ion beam that results. This makes it popular for designing targeted quantitative assays for chemical species that are known not to suffer thermal degradation effects. Heated nebulizers readily accommodate conventional LC flow rates in the milliliters per minute range, which enable large injection volumes to be used resulting in low sample concentration detection limits. Matrix suppression of ionization also tends to be more easily controlled than for electrospray based assays (King et al., 2000). For these reasons APCI has maintained a very high adoption rate over the years, as depicted in Figure 1 with the first demonstration of high speed quantitative drug analyses to appear in 1986 (Covey, Lee, & Henion, 1986). It is not uncommon, in laboratories dedicated to quantifying drugs in biological samples, to conduct 20–30% of their analyses by APCI with the remainder by high flow ESI. Sampling efficiencies in the milliliters per minute flow range are very similar to those obtained for the high flow ESI interfaces at the same flows as the data shown in Figure 9. The reason for the ion losses at these high flow rates is similar to the high flow rate electrospray situation. Ions are created with very high efficiency but are distributed throughout the API source and not subject to the efficient focusing provided by the vacuum drag into the mass spectrometer. APCI has not been adapted to nano flows largely because there has been no impetus from a methodology point of view for reasons cited above. However it is possible that low volumes of liquid and samples can be ionized by APCI directly in front of the vacuum aperture, and thus achieve similar sampling efficiencies as nano-ESI, by taking advantage of the vacuum drag at that point. Corona discharge sources are not ideal for this because the space charge that they induce will repel ions out of this critical vacuum drag region. However other field-free means of ionizing the reagent gas can be used such as 63Ni beta emitters or photons. Some steps have been made to construct micro-heated nebulizers to accommodate the nanoflow regime where all of the ions and solvent vapors could be inhaled by the MS (Ostman et al., 2004). IV. OTHER ATMOSPHERIC PRESSURE ION SOURCES: VARIATIONS ON THE ORIGINAL THEME A myriad of other atmospheric pressure ionization sources have developed over the past decade. Some of them are based on liquid sample introduction systems and multiplexed versions of these, others have moved back to the earlier concepts of solid sample introduction reminiscent of API in the 1980s. So far they all share 882 in common the final method of sample ionization, either ESI or APCI, or some slight variation on these themes. A variety of different methods for transferring the sample from the solid to the gas phase have emerged often for the explicit purpose of surface analysis or chemical imaging (Van Berkel, Pasilis, & Ovchinnikova, 2008). Below we will briefly mention some of the more popular techniques that have emerged to date, many of which have not yet reached their full potential. A. Variants of ESI 1. Electrochemical Cells Avery large body of work studying the electrochemical processes involved in the electrospray process has been conducted by Van Berkel (2007a) and Van Berkel and Kertesz (2007b). During these studies it was observed that one can take advantage of these processes to drive oxidation and reduction reactions to extend the ionization capabilities of ESI, as well as to conveniently create by-product species that mimic biological metabolism. Specialized porous graphite flow-through electrochemical cells were integrated into the ESI emitters to maximize this additional ionization capability (Van Berkel & Kertesz, 2005). 2. Sonic Spray The origins of sonic spray hearken back to the original research in ion evaporation, the creation of charged droplets by frictional forces rather than electrochemical means driven by an external power supply. It occurs when a solution in a capillary is sprayed with a high velocity gas flowing co-axially to the capillary, and approaches optimal conditions when the gas flow approaches the speed of sound (Hirabayashi et al., 1994). A similar technique is Electrosonic Spray Ionization, however in this case a bias voltage is used similar to the ion spray technique as described earlier (Bi et al., 2000). Protein ions tend to have fewer charges than those produced by nano-ESI, suggesting that protein folding is conserved during ionization (Takats et al., 2004). Another variant of this is Cold Spray ionization which sprays into a liquid nitrogen cooled chamber to try to preserve non-covalent complexes (Yamaguchi, 2003). 3. DESI Desorption Electrospray Ionization (Cooks et al., 2006) is a surface ionization technique that operates by impacting highvelocity charged droplets with the surface to be analyzed, extracting the surface chemicals by the liquid that makes momentary contact (Costa & Cooks, 2007). The sensitivity is primarily dictated by the extraction efficiency that occurs during the impact event. The nebulizers that are used are similar to the pneumatically assisted electrospray and sonic spray devices described earlier. A similar approach, referred to as DeSSI, utilizes a sonic spray nozzle without the high voltage (Haddad, Sparrapan, & Eberlin, 2006). 4. Surface Sampling Probe The surface sampling probe (SSP) is a direct liquid extraction device for surface analysis and imaging (Van Berkel et al., 2008). Mass Spectrometry Reviews DOI 10.1002/mas ATMOSPHERIC PRESSURE ION SOURCES It is coupled to the ESI source utilizing the self aspirating capabilities of pneumatic nebulizers to deliver and draw sampling liquid from a surface. Because of the high extraction efficiency it has very high sensitivity for use as a surface analysis device and can be adapted to chromatographic separations if needed (Kertesz, Ford, & Van Berkel, 2005). It is also readily adapted to APCI sources (Asano et al., 2005). 5. Fused Droplet ESI Work to ionize the eluent from a gas chromatograph (Lee & Shiea, 1998) or flow pyrolyzer (Hong, Tsai, & Shiea, 2000), through intersection with charged electrosprays, eventually evolved into a new variation of electrospray. Early fused droplet electrospray sources involved intersection of an ultrasonically nebulized neutral liquid stream containing the analytes with multiple electrospray plumes to ionize various components of the stream (Shiea et al., 2001). The use of a simple nebulizer and single electrospray probe, oriented to ensure charged plume and neutral droplet intersection (Shieh, Lee, & Shiea, 2005), are common between this and a similar technique referred to as extractive electrospray ionization (Chen, Venter, & Cooks, 2006). 6. ELDI, MALDESI, and LAESI Electrospray-assisted laser desorption/ionization mass spectrometry (ELDI) is another surface analysis technique whereby a nitrogen laser is used to directly desorb neutral analyte molecules by thermal processes. The neutral vapor is intersected with electrosprayed droplets that incorporate the gas phase molecules which then ionize by ESI (Shiea et al., 2005). With matrix assisted laser desorption ESI (MALDESI) the sample is indirectly desorbed via a transfer of energy through a MALDI matrix to the sample using a similar laser to ELDI (Sampson, Hawkridge, & Muddiman, 2006). The laser ablation electrospray ionization technique (LAESI) uses a mid-IR laser to directly desorb the sample and has the ability to accommodate aqueous samples (Nemes & Vertes, 2007). Both MALDIESI and LAESI also utilize electrosprayed droplets intersecting the vaporized sample for the final ionization event. & eventually results in transfer of charge to the analyte ions through gas phase reactions (Robb, Covey, & Bruins, 2000). In many instances APPI has shown a sensitivity advantage over the corona discharge source likely because of the field-free nature of the photon source. Space charge repulsion of the high intensity corona can lead to some scattering of ions. The technique has found further application as a component of field deployable and air analysis instruments (Syage, Hanning-Lee, & Hanold, 2000) and as a component of multipurpose sources combined with APCI (Syage et al., 2004). 2. DART The direct analysis in real time device (DART) was developed primarily as a surface ionization source (Cody & Laramee, 2005). Initial ions are created with a discharge in a flowing stream of a hot noble gas such as argon or helium, forming metastable excited state atoms by Penning ionization which transfer charge to the background gas in an API source by chemical ionization processes. Analyte molecules are desorbed from surfaces by the heated gas. 3. ASAP Similar to DART the atmospheric pressure solids analysis probe (ASAP) was developed as a rapid and convenient device to analyze solid sample surfaces and liquids (McEwen, McKay, & Larsen, 2005). Vaporization of the sample is achieved with a hot nitrogen gas stream and ionization occurs by APCI initiated with a corona discharge. 4. PADI Plasma assisted desorption/ionization (PADI) is a surface analysis technique which involves the direct contact of a sample on a surface with a plasma generated from a radio-frequency discharge needle. The ionization in the plasma is thought to include a mixture of metastable Penning ionization, electron ionization, and ion molecule reactions. Since the resulting spectra are dominated by products of ion molecule reactions (proton transfer and proton abstraction) chemical ionization appears to dominate (Ratcliffe et al., 2007). B. Variants of APCI 1. Photoionization 5. Laser Desorption Techniques: LDTD, AP-LD/CI, LIAD The motivation for the research in atmospheric pressure photoionization (APPI), as with many of the others cited below, was to try to find a means of direct ionization of the analyte molecules so that substances that could not be ionized efficiently with chemical ionization processes could be accessed with an API source. In general this goal has met with limited success to date with any of the approaches described in this review, but other benefits have accrued from these developments. With APPI the sample is vaporized with a conventional heated nebulizer and the initial ionization event is with photons from vacuum ultraviolet lamps having variable photon energies around 10 eV. Often dopants such as toluene are used to create the first ionization event, which Several laser based techniques are used to desorb samples from surfaces and subsequently ionize by APCI. One technique indirectly desorbs ions using conductive heat transfer of samples through a metal surface using IR lasers. This technique is referred to as laser diode thermal desorption or LDTD and was presented at the 54th ASMS conference in 2005 by P. Picard. Atmospheric pressure laser desorption/ chemical ionization (AP-LD/CI) uses direct laser heating of the sample followed by chemical ionization (Coon, McHale, & Harrison, 2002). Laser-Induced Acoustic Desorption (LIAD) uses a Nd:YAG laser to generate acoustic waves in a metal foil to desorb the sample (Shea et al., 2007) followed by the chemical ionization process. Mass Spectrometry Reviews DOI 10.1002/mas 883 & COVEY, THOMSON, AND SCHNEIDER 6. Laser Ionization Techniques a. APLI. Atmospheric pressure laser ionization (APLI) is a laser based technique but is different from the others in that the sample is not desorbed by the laser but rather ionized by it. It is a spray technique where a heated nebulizer is used to vaporize a solvent containing the sample creating a dry, droplet free vapor. Ionization is performed by selective resonance enhanced multiphoton ionization (REMPI) where the laser energy appears to be directly ionizing the sample molecules as indicated by formation of the radical cation molecular ions from aromatic hydrocarbons (Constapel et al., 2005). Similar spectra have been observed with photoionization and aprotic dopants as a result of electron transfer processes from the reagent ions, and it is yet to be established to what degree APLI avoids ion molecule reaction chemistry. b. AP MALDI. While MALDI has been widely used in vacuum for many years, only recently has it been applied to API sources (Wolfender et al., 1999; Laiko, Baldwin, & Burlingame, 2000). Ion sampling considerations are much the same as they are with ESI based sources. Although there are no liquids to desolvate per se, clusters and particles are formed in the atmospheric pressure ablation plume that require significant quantities of heat to liberate the ions. The heated transport tubes or chambers typically used to help decluster ions for ESI tend to require higher temperatures for AP/MALDI to achieve maximum ion transmission. This would seem reasonable given the lower vapor pressure of ablated solid matrix crystals and clusters versus liquid solvents. High ion transport efficiencies, similar to those with nano-ESI, can be obtained by taking advantage of the vacuum drag effect (Wang et al., 2006). The mechanism of the desorption and ionization event is a topic of intense study, with a large component of the process appearing to be the formation of matrix/analyte clusters and the subsequent charge separation of these species particularly with protein and peptide like molecules (Karas & Kruger, 2003). The behavior of low molecular weight drug like substances (<1,200 Da) with MALDI empirically bears a striking resemblance to the thermal desorption APCI process. A study of low molecular weight compounds (120–700 Da), covering a broad range of chemical space including strong and weak acids and bases, neutrals, quaternary ammonium compounds, and amphoteric compounds, comparing MALDI (alpha-cyano matrix), APCI, and high flow ESI on a triple-quadrupole instrument, showed a strong correlation between MALDI and APCI with very little correlation of either of them with ESI (Covey et al., 2009). Compounds that thermally degraded with flow injection APCI and showed low molecular ion sensitivity also did so with MALDI. ESI showed strong molecular ion response on these types of labile molecules (drug conjugates for example) and, from this comprehensive compound set, roughly 35% of the substances fell into this category, that is, strong molecular ion response with ESI but little response with APCI or MALDI. Both mono- and bi-functional pre-formed ions (quaternary ammonium compounds such as succinyl choline) showed strong response with ESI for the solution state ion, (M)2þ in the case of bifunctional quaternary ammonium compounds, supporting a direct desorption mechanism, but no substantial response (at least 884 1,000-fold lower) from either MALDI or APCI indicating that thermal degradation or gas phase neutralization processes dominate. Neutral molecules behaved for all three sources according to the expectations from gas phase ion/molecule reaction thermodynamics which correlates closely with solution state polarity. In our experience a good first indication of whether or not a compound will efficiently produce molecular ions by MALDI is to flow inject the analyte into a heated nebulizer APCI source. Recently new matrices designed to optimize gas phase proton transfer reactions have been developed which improves the MALDI ionization of low proton affinity substances compared to the alpha-cyano matrix (Jaskolla, Lehmann, & Karas, 2008). Different controlled biological matrices that suppressed ionization were also compared in this same study with MALDI, APCI, and ESI, again showing a close similarity between APCI and MALDI (Covey et al., 2009). Biological extracts that were high in lipids and surfactants were devastating to electrospray ionization efficiency, consistent with a desorption mechanism from a droplet surface. These same matrices were not nearly as suppressing for APCI or MALDI, indicating a gas phase ionization mechanism although they were not entirely immune from all suppression effects. High gas phase proton affinity substances, which tend to be low molecular weight and polar in biological samples, strongly suppressed APCI and MALDI which supports a gas phase reagent ion depletion process. These same substances (urea for example), because of their polarity, tend to elute in the void volume of reversed phase LC separations, creating a situation where APCI in practical analytical scenarios demonstrates fewer examples of ion suppression problems than ESI since lipids elute during the chromatographic process sometimes at the same time as the analytes. It is to be expected that LC/MALDI would demonstrate the same advantage as APCI in this regard. C. Multisprayer Sources Because neither vacuum consideration nor physical space are limitations with API sources, they are particularly convenient for mounting multiple ESI and APCI inlets, and sometimes combinations of the two. This has found application in such areas as accurate mass measurement, ion-ion reactions, sensitivity improvement at higher liquid flow rates, and multiplexing to achieve high throughput analysis. Combinations of any of the various atmospheric pressure ion sources are possible; however, multiple sprayer electrospray ion sources and dual sources based upon ESI, APCI, and APPI are the dominant approaches in the literature. A recent publication reviewed the general and patent literature relating to ion sources containing multiple electrospray probes (Schneider & Covey, 2007). In general, these types of sources can be categorized by the presence or lack of indexing, which is the means by which the mass spectrometer rapidly turns on and off the signal coming from each of the multiple inlets and assigns the resulting signal to a particular inlet channel. The literature contains a number of examples of non-indexed multiple sprayer approaches for purposes such as mass calibration (Jiang & Moini, 2000) ion/ion reactions (Ogorzalek Loo, Udseth, & Smith, 1991, 1992), ion/molecule reactions (Ogorzalek Loo et al., 1992), and throughput enhancement (Hiller et al., 2000). Mass Spectrometry Reviews DOI 10.1002/mas ATMOSPHERIC PRESSURE ION SOURCES Indexed multiple sprayer systems require a means for enabling and disabling each of the sprayers, as well as the ability to assign a signal to a particular sprayer. The various means can be grouped into four main categories: fluid selectors (A), spray controllers (B), spray blockers (C), and ion beam selectors (D), as shown in Figure 13. Fluid selectors operate by enabling and disabling the liquid flow to a sprayer with valves (Schneider & Covey, 2007). Spray controllers operate by turning on and off the strong electric field at the electrospray tip (Schneider, Douglas, & Chen, 2002). Spray blockers interrupt the spray between the emitter and vacuum entrance by mechanical means such as a blocking plate (Wang et al., 1999) or by moving the sprayers relative to the entrance aperture (Chan et al., 2002; Park, Kim, & Kim, 2004). Multiple API vacuum apertures, each with an independent sprayer, can be used to create independent ion beams in the vacuum system that can be deflected with lenses. Although this presents the fastest means of switching, it would require higher vacuum pumping speeds that may not be practical. Efforts to improve compound coverage have led to the combination of ESI and APCI inlets. Agilent and Waters have accomplished this by electrospraying into a region containing chemical ionization reagent ions. The use of APCI and APPI has been combined into a single ion source in an attempt to extend ionization capabilities on Thermo instruments (Syage et al., 2004). Alternatively, a switching valve may be used to direct a single flow stream sequentially to separate APCI and ESI probes to generate mass spectra based upon each ionization method as commercialized by AB/Sciex. FIGURE 13. Four different means to index a multiple sprayer ion source to a mass spectrometer. A: Fluid selectors gate the signal by interrupting the fluid on the way to the sprayer. B: Spray controllers gate the signal by controlling the electric field at the sprayer tip. C: Spray blockers interrupts the spray between the sprayer tip and the mass spectrometer entrance. D: Ion beam selectors gate the signal inside the vacuum system with a lens where the beams originate from multiple atmosphere-tovacuum apertures. Reprinted with permission from Schneider, 2007, copyright 2007, Elsevier. Mass Spectrometry Reviews DOI 10.1002/mas & Operation of arrays of nano-ESI emitters has been studied for several years as a means of increasing the ion sampling efficiency at liquid flow rates beyond the nano-ESI regime (Rulison & Flagan, 1993; Deng et al., 2006). Multiple emitters in front of multiple atmosphere to vacuum sampling apertures have shown gains over a single emitter/aperture arrangement at 1 mL/ min flow (Kelly et al., 2008). While promising, this approach introduces a high degree of complexity and simpler single emitter systems have achieved 80% sampling efficiency at 500 nL/min (Schneider, Javahari, & Covey, 2006) leaving little room left for improvement. There are, as mentioned earlier, significant opportunities for improvements in the high flow regime and it is here where the multiple sprayer approach might make gains in the future. V. CHEMICAL NOISE: AN OUTCOME OF ATMOSPHERIC PRESSURE IONIZATION TECHNIQUES Chemical noise is a special problem of the atmospheric pressure ionization techniques which is often responsible for raising detection limits in both quantitative and qualitative analyses. Many substances are ionized by API techniques such that trace contaminants in solvents, apparatus, and air, in combination with clusters, generate ions across the m/z range. Compounds such as phthalates, silicones, phenyl phosphates, sebacates, adipates, and clusters of these ions with HPLC solvents and additives, are ubiquitous and unavoidable. Despite dramatic efforts to improve ion source hardware with pure materials such as stainless steel and ceramics, judicious use of heat, positive source pressure to exclude ambient air contaminants, off-axis sprays, curtain gases, declustering fields, off-axis and curved entrance ion optics, and software noise filtration techniques, this problem has remained persistent. Meticulous laboratory practices utilizing the purest available solvents and gases have also not solved the problem in a general sense. At low mass ions exist at nearly every nominal mass at different intensities, many of which have been identified (Guo, Bruins, & Covey, 2006; Keller et al., 2008). With the improved sensitivity of API sources chemical noise has appeared even under tandem MS conditions, confounding spectra interpretation and increasing baseline noise under multiple reaction monitoring conditions. Since chemical noise is a consequence of atmospheric pressure ion sources, approaches for dealing with it have been emerging in recent years and this line of development will undoubtedly continue on into the future with increasing emphasis. For this reason some of the recently investigated techniques will be briefly mentioned here whether or not they are an integral part of the API source. They tend to fall into two general categories. In the first case extrinsic properties of molecules and their interactions with other species in their immediate environment are leveraged, that is, chemical properties. In the second case we are dealing with intrinsic properties of molecules that are independent of their immediate surroundings such as nuclear composition (mass) or internal bond energies (molecular fragmentation). One successful example of a chemical means to remove background ions is the dynamic reaction cell developed for 885 & COVEY, THOMSON, AND SCHNEIDER inductively coupled plasma mass spectrometry, an atmospheric pressure ion source designed for elemental analysis. Gas phase ion/molecule reactions in cohort with high and low mass band pass filters can be used to effectively separate isobaric elemental ions. The separation of argon and calcium ions with ammonia is an immediate example (Tanner, Baranov, & Bandura, 2002). This principle has been applied to electrospray instruments and organic analyses, utilizing dimethyl disulfide to selectively separate clusters and sodiated ions from protonated molecules (Guo, Bruins, & Covey, 2007). Another approach that presents promise for development is the use of ion/ion reactions to induce charge inversion of targeted analyte ions (Min, Emory, & McLuckey, 2005). This borrows from concepts in accelerator mass spectrometry but accomplishes the charge inversion in a much different way. A third approach is the use of mobility separations prior to mass analysis. Of particular interest is the implementation of FAIMS (Purves & Guevremont, 1998) or Differential Mobility spectrometry (Levin et al., 2006) which separate ions on the basis of the difference between their high and low field mobilities. The reason these mobility approaches are classified as chemical techniques is because the selectivity obtained by such devices is highly influenced by the type of carrier gases that are used and can be adjusted by altering this parameter. This behavior bears analogy with the mobile phase in HPLC. The second category of approaches to noise reduction is an extension of mass spectrometry development that has been undergoing continued activity since the earliest days. These include means of obtaining higher resolution and serial fragmentation of molecules with collisions and multiple stages of mass analysis. The field is currently focused on trapping type analyzers such as FTMS and Orbitraps as well as the pulsed beam approach of time-of-flight. They all can achieve resolution in the tens of thousands or greater, and each has their own particular attributes such as relative speed and dynamic range. The important question is how much resolution is required to eliminate all possible scenarios where chemical noise is limiting. Some early empirical studies presented by Bateman at the 2008 ASMS in Denver with an Orbitrap mass spectrometer indicated that a resolution of 30,000 may provide sufficient selectivity for routine drug discovery assays, but these are early days and the number of potential scenarios where noise is limiting is almost infinite. Multiple stages of MS/MS is another way to gain selectivity from intrinsic properties of molecules. Three-dimensional ion traps have demonstrated the capability to perform MSn but suffer from duty cycle and ion current loss limitations. Combinations of linear ion traps with other mass analyzers such as Orbitraps, quadrupoles, and TOFs are beginning to show promise in this area, where the particular noise reduction application will be for targeted quantitative analysis (Leuthold et al., 2004; Collings, 2007). What is emerging from this field of active research is that no one technique will solve all chemical noise problems. Techniques that rely on intrinsic properties of molecules will likely be more generally applicable and easier to control, requiring less understanding of chemical selectivity. However, they will never be able to distinguish between substances that have the same exact mass which is well within the realm of the techniques that use chemical interactions and reactions (Eberlin, 2006). 886 VI. ION TRANSPORT A. Transport of Ions From Atmosphere to Vacuum 1. Desolvation and Declustering Declustering is the term used to describe the process of removing the last layer of solvent molecules from gas-phase ions, while desolvation is a broader term that can include the process of evaporating macroscopic droplets. All atmospheric pressure ion sources and interfaces must accommodate means for both desolvating and declustering, and these techniques can overlap. The problem is similar for the high and low flow sources, differing only in the magnitude of the amount of solvent that has to be vaporized. Declustering must be carried out before the ion enters the mass analyzer. The general techniques used for declustering (Bruins, 1991) have not significantly changed in the last 15 years. However there have been some modifications tailored to the nanoflow regime that will be discussed below. On the atmospheric pressure side of the aperture declustering can be carried out by drying the ions with either heat or dry gas, or a combination of both. On the vacuum side of the aperture an electric field can be used to decluster by means of collisioninduced dissociation. A counterflow of dry curtain gas in front of the orifice acts to shift the cluster equilibrium to smaller clusters (and eventually the bare ion), and also keeps solvent from entering the free jet expansion where it can re-cluster with the ions. Heating the gas or the region in front of the aperture assists further in declustering. One way to achieve this is to use a heated capillary tube to sample the ions from atmosphere into vacuum. The ions become declustered in their passage down the heated tube, which may be 0.3–0.5 mm diameter and up to 10 cm long. Today all methods described above are used widely, and often together. Special efforts to increase the residence time of droplets in hot zones in front of the aperture have been undertaken to improve nano-ESI operation and attempt to maintain nano-ESI efficiencies at flow rates as high as possible, extending into the low microliters per minute range. Since the first description of nano-ESI it was apparent that operation in the nanoflow regime offered the possibility for very high sampling efficiency. A number of research groups demonstrated nano-ESI sampling efficiencies on the order of approximately 1–10% for electrospray infusion at solvent flows ranging from 2 nL/min up to tens of nanoliters per minute (Wilm & Mann, 1994; Geromanos, Freckleton, & Tempst, 2000; El-Faramawy, Siu, & Thomson, 2005). The desire to improve nano-ESI efficiency further, and increase the flow rate to couple it more readily with liquid chromatography approaches, has led to the development of new types of interfaces capable of high performance operation at flow rates in the hundreds of nanoliters per minute. As the flow rate increases from the traditional nano-ESI regime (<100 nL/min) up to the nano-LC regime (100–1,000 nL/min), maintaining high efficiency is challenging due to the increased desolvation requirements as well as the increased tendency for the spray to diverge over a larger area. High sampling efficiency requires capturing the majority of the electrospray plume and providing sufficient desolvation means for efficient ion liberation. Mass Spectrometry Reviews DOI 10.1002/mas ATMOSPHERIC PRESSURE ION SOURCES Figure 8B shows an illustration of a particle discriminator interface (PDI) specifically optimized to provide high efficiency for electrospray operation up to approximately 1 mL/min (Schneider et al., 2003, 2005). It comprises a gas conductance limiting orifice separating the atmospheric pressure source region from the first vacuum chamber of the mass spectrometer. Sealed onto the orifice is a short laminar flow chamber of larger diameter than the orifice that can be heated to temperatures on the order of 3008C to ensure complete desolvation over a variety of flows. The interface configuration also includes a counter-current gas flow to aid in removal of solvent molecules prior to entrance into the vacuum system. Since the orifice controls the rate of gas flow into the vacuum system of the mass spectrometer, the diameter of the heated laminar flow chamber can be varied substantially with no effect on the downstream pressures. The vacuum draw into the orifice establishes laminar flow conditions through the heated chamber, and generates a suction effect at the entrance. With suitable sizing of the laminar flow chamber diameter, electrospray plumes for solvent flow rates up to approximately 1 mL/min can be entirely consumed into the inlet. The charged droplets and ions within the laminar flow chamber are entrained in the net gas flow to the orifice as shown in Figure 14. In general, the laminar flow chamber is designed such that the ratio of volume (dependent upon radius and length) to gas flow rate provides approximately 1 msec residence time for ions. Increasing the residence time further can improve desolvation efficiency for a given temperature setting, but diffusional ion losses can become significant. A fully optimized version of this interface has been shown to maintain electrospray sampling efficiency for a favorable sample in the range of 70–80% at flows up to approximately 500 nL/min, making it amenable to nanoflow LC/MS (Schneider, Javaheri, & Covey, 2006). The necessary temperature to achieve these high efficiencies was directly related to both the solvent flow rate and the gas flow rate (dictated by the size of the orifice). Similar experiments conducted using a traditional heated capillary inlet demonstrated similar plume consumption phenomena with >90% of the total ion current consumed in to the inlet at flows up to approximately 1 mL/min. However, diffusion losses and insufficient heat for desolvation limited the observed sampling efficiency to approximately 20% (Page et al., 2007). These experiments highlight the critical importance of both capturing the electrospray plume and achieving sufficient desolvation as the main drivers for very high sampling efficiency in electrospray ionization. Complete declustering is not usually achieved on the atmospheric pressure side of the aperture. In addition, neutral solvent molecules that enter the vacuum system will re-cluster with molecular ions in the supersonic expansion, requiring the use of additional declustering on the vacuum side of the aperture. In the most common sampling configuration, a skimmer is located with the tip 2–10 mm downstream of the aperture. The background pressure in the vacuum region between the aperture and skimmer is of the order of 1–3 Torr (Bruins, 1991), the skimmer tip can either be within the Mach disc of the supersonic expansion, or it can be downstream of the Mach disc. More details of the gas flow and pressure patterns associated with the supersonic expansion are provided later, but here it is sufficient to note that the gas density between the orifice and the Mass Spectrometry Reviews DOI 10.1002/mas & skimmer is highly inhomogeneous, changing by several orders of magnitude in a few mm. A voltage difference between the aperture and the skimmer of 10–200 V accelerates the ions to collide with the background gas (air or nitrogen). The voltage is usually optimized to produce declustering without fragmentation (although the ability to induce fragmentation by increasing the voltage is an important feature of this geometry). A second and equally important function of the voltage difference is to focus the ions, as described later. The declustering voltage can be compound dependent (due to different tendencies to fragment), m/z dependent and solvent dependent. The actual voltages used depend on the distance and the pressure (or gas density) between the orifice and skimmer, which varies among different instruments. The voltage is normally considered to be a user-adjustable parameter, to be optimized as needed. B. Transport of Ions From the Ion Source Into the Vacuum Chamber 1. Focusing Ions at Atmospheric Pressure Ions created in the ion source must be transported from atmospheric pressure into high vacuum where the pressure is typically 3 105 Torr or lower, requiring small apertures to separate the ion source from the vacuum. This suggests immediately that focusing the ions in the source toward the orifice should improve the transmission efficiency and sensitivity. However, at the most basic level, it can be shown that any attempt to increase the concentration of ions by using DC focusing fields at atmospheric pressure where ion motion is governed by ion mobility is frustrated by Maxwell’s equations (specifically Gauss’ Law). Considering the issue in more detail, the first aperture separating the source from the ion optics is typically 0.2–1 mm in diameter, most commonly 0.3–0.7 mm. In some cases the aperture is a tube with a length-to-diameter (L/D) ratio of >20, in other cases a thin plate with L/D ratio of approximately 1. Ultimately it is gas drag that carries ions from atmospheric pressure into the first vacuum stage, which is normally pumped by a mechanical vacuum pump to a pressure of 0.5–10 Torr. Gas drag (or suction) in the region close to the orifice pulls ions from an area that is larger than the actual area of the hole. The effective size of the aperture due to the gas drag (i.e., the area from which ions are pulled through into vacuum), relative to the physical aperture size, depends on the electric field strength on the high pressure side of the aperture. Strong focusing fields can and usually are applied to direct ions toward the orifice. At atmospheric pressure ion velocity is directly proportional to the field strength (with minor higher order terms as the field strength becomes very high), and ions effectively follow the lines of electric flux. Flux lines can be made to converge toward a small region (the orifice), focusing ions toward the orifice. As suggested above, it is often not appreciated that the volume concentration of ions does not increase, due to Gauss’ law: as the flux lines converge toward a point, the magnitude of the electric field increases, as does the ion velocity. Therefore the ions converge toward a smaller area in two dimensions, but their velocity in the 3rd dimension (toward the hole) increases so that 887 & COVEY, THOMSON, AND SCHNEIDER FIGURE 14. Computational fluid dynamics (CFD) models of gas entrainment streamlines with PDI interface with a 1 mm i.d. laminar flow chamber attached to a 0.25 mm orifice with a 1 mm spacer of wider i.d. FIGURE 16. Computational fluid dynamics (CFD) model of the free jet expansion, and calculated ion trajectories. Upper half of figure shows gas flow (purple lines) and ion trajectories for m/z 228 (black lines). Lower half of figure shows ion trajectories for mass 16,950 Da (9þ charge state at m/z 1,883.4). Color key shows gas velocity. Upper half shows a velocity range from 0 to 3,150 m/sec; lower figure half shows an expanded range of 0–720 m/sec. 888 Mass Spectrometry Reviews DOI 10.1002/mas ATMOSPHERIC PRESSURE ION SOURCES the ion concentration is unchanged. Because the gas volume flowing through the hole is constant, the ion flux through the hole cannot be increased if the ion volume concentration remains unchanged. Of course, if the ions could be focused toward a small hole and carried through the hole into the ion optics, and if the gas could be pumped away while the ions are captured by the optics, then the intended increase in ion concentration and hence ion sampling efficiency would be achieved. To first order, the real gain in ion transmission due to atmospheric pressure focusing with static (DC) fields is given by the ratio of the electric field strength in the vacuum region downstream of the aperture to the electric field strength on the atmospheric pressure side of the aperture (assuming that all ions could be captured and focused by the optics). However, the high electric field strength that focuses the ions toward the orifice cannot be carried through into vacuum because of electrical breakdown. The weaker field strength that must be maintained on the vacuum side causes the ion trajectories to diverge, and the focusing gain is lost as ion trajectories diverge and ions are lost to the wall of the aperture as they enter the vacuum system. In spite of this rather negative view of the benefit of focusing ions toward the hole, some successes can be and have been achieved. For example, a small conical-shaped orifice was employed on the API III triple quadrupole system. This conical shape produced some focusing of the ions toward the tip of the cone, and produced a twofold improvement in sensitivity compared to a flat disc with the same orifice diameter. A saving factor is that as the ions come close to the hole, the gas drag begins to dominate over the electric field. As the electric field starts to weaken close to and through the plane of the orifice, and the ion flow starts to diverge, the gas drag force overcomes the field strength and ions are pulled through. This dominance of the gas drag over the electric field near the aperture is the vacuum suction effect. In a typical source with field strengths of 1,000– 3,000 V/cm on the upstream side (and lower fields on the vacuum side) the effective area of the orifice can be three to four times larger than the actual area. This applies to thin apertures as well as tubes (where the field strength inside the tube drops to zero). In principle, a weaker field on the upstream (high pressure) side results in a larger effective sampling area. In the limit, if ions could be generated in a large field free volume, and the only exit was suction through the vacuum orifice, then all ions could be sampled (limited by diffusional losses to the walls which are not considered here). This issue has been recognized by Willoughby and Sheehan and presented at the 2008 ASMS conference in Denver. Methods were proposed for reducing electric fields in the high pressure ion source, as well as for using laminated apertures and tubes to prevent the field divergence into vacuum. However, practical implementation in an electrospray source has not been fully demonstrated. In reality, most API sources use significant gas flow to nebulize and dry ions, and most of the gas, droplets and charge exit through a waste port. Any attempt to focus the ions from this flow toward the orifice must overcome the gas flow velocity in the source, extracting ions from the flow and moving them toward the orifice. The large field required for this purpose limits the ability to focus ions through the hole for the reasons described above. This catch-22 situation frustrates the simple focusing schemes that have been proposed and tried in the past. Some early attempts Mass Spectrometry Reviews DOI 10.1002/mas & to focus ions from a corona discharge point (by reconverging the field lines from the discharge point toward an orifice in a ‘‘mirror’’ configuration for the electric field lines) were able to achieve almost complete focusing of the 5 microamp current on the region close to the orifice (or on a small collector plate approximately the size of the orifice), but required field strengths approaching the breakdown limit to do so (the mirror image of the corona discharge itself). Even staying below the breakdown limit, attempts to get the highly focused beam through the hole were not very successful. Compared to the normal, rather unfocused geometry, a factor of three increase in ion current was the best that could be achieved (unpublished data). In recent years, ion focusing at atmospheric pressure has been accomplished using RF fields in specific geometries. The most successful configuration has probably been achieved with a FAIMS source. While FAIMS has been developed as an ion separation method at atmospheric pressure, it can also act to focus or concentrate the ions into a smaller region of space (Guevremont & Purves, 1999). The method has been used to both focus (while separating ions) and to trap ions in a small region of space in front of the orifice. It is important to note that the technique does not focus all ions together, but rather focuses ions of a specific mobility and configuration, while other ions are defocused or lost, so it is not a general solution to the high-pressure focusing problem. It has also been proposed to use RF ion guides to focus ions at atmospheric pressure (Willoughby & Sheehan, 1984). RF ion guides are often used at low or intermediate pressure to focus ions (described later, and also in the review article by Douglas, this volume). While the same technique can in principle be applied at atmospheric pressure, the focusing effect is very slow (in time) at high pressure, so either a small device is required (which makes it unsuitable for collecting and focusing ions from a large-volume source) or a very long device is required to give the ions sufficient time to focus to the center. C. Sampling Ions Into High Vacuum 1. The Orifice-Skimmer Geometry The vacuum aperture consists of either a thin plate with an orifice or a tube that can be several centimetres long. The orifice produces a well-defined supersonic expansion through the hole, with ions and gas accelerated to supersonic velocities in a jet structure that has been thoroughly studied and characterized in the field of rarified gas dynamics (Campargue, 1984). Figure 15 shows the basic structure of the barrel-shaped expanding jet, terminating in a shock wave (often referred to as the Mach disc) where the supersonic gas velocity becomes sub-sonic through a thin re-compression region. Ions carried by the expanding gas are usually sampled into the next vacuum stage through a skimmer. The skimmer can be located with the tip upstream of the Mach disc (point A in Fig. 1) so that ions and gas are sampled inside the so-called ‘‘zone of silence,’’ or downstream of the Mach disc (point B of Fig. 15), a position that allows ions to pass through the shock wave before being sampled. There are several considerations when deciding where to sample with the skimmer. Upstream of the Mach disc the 889 & COVEY, THOMSON, AND SCHNEIDER FIGURE 15. Structure of the supersonic free jet. (1) Orifice plate. (2) Barrel shock defining the radial boundary of the free jet expansion. (3) Mach disc, defining the axial boundary of the supersonic portion of the free jet expansion. (4) Skimmer. A: Position upstream of the Mach disc. B: Position downstream of the Mach disc. velocity of ions and gas is high and primarily directed parallel to the axis. A skimmer at this location samples a higher gas and ion flow than a skimmer located further downstream. Additionally, the gas density is lower within the zone of silence, so that ions can be more readily focused toward the skimmer, using for example a small ring electrode as shown in Figure 15. Outside of the Mach disc, the pressure is higher, but the gas velocity is lower. Because of the larger distance from the orifice, the gas flux into the next chamber is significantly reduced, which is beneficial when considering the size of vacuum pump required. However, the ion flux is also lower and it is more difficult to refocus ions efficiently through the skimmer at the higher pressure. These considerations have led to different approaches. In Sciex mass spectrometer systems the skimmer has always been located upstream of the Mach disc, providing maximum benefit for ion focusing and declustering (both leading to higher sensitivity). Other systems use a skimmer located outside of the Mach disc and in some cases the skimmer is located slightly offaxis to prevent large droplets from being sampled directly into the ion optics and subsequently to the detector. The search for a deeper understanding of the limitations of ion sampling efficiency in the orifice-skimmer region led to the application of computational fluid dynamics (CFD) to the ion sampling region. A sophisticated gas-dynamic model was developed and combined with an electric field and ion trajectory calculator to study the neutral gas and ion motion between the orifice and the skimmer (Jugroot et al., 2004a,b, 2008). This modeling effort provided clear insight into the structure of the expansion which was numerically modeled in detail for the 890 first time on this small scale. It also provided information on the ion motion due to the electric fields that are applied to both focus and decluster the ions. The simplistic view has been represented in the past as an axi-symmetric expansion with a Mach disc that is penetrated by the skimmer (as shown in the schematic of Fig. 15). The nature of the flow in the region downstream of the skimmer (into the next vacuum chamber) has never been clear. The numerical model provided details that showed that the skimmer disrupts the Mach disc so that it is never visible, and the flow persists through the skimmer and into the next chamber as shown in Figure 16. Figure 16 shows the computed ion trajectories (black lines) and neutral gas flow (purple lines) for a 0.25 mm orifice followed by a 2.3 mm diameter skimmer located 2 mm downstream. The upper part of the figure shows the trajectories of ions of m/z 228 (typical of a small singly charge ion) and the lower part shows ion trajectories for mass 16,951 amu with a charge state of 9þ (m/z 1,859), representative of a large biomolecule. The focusing effect is provided by a potential difference of 80 V between the orifice and skimmer. For this case the focusing ring shown in Figure 15 was not included. It is clear that the smaller ions are more efficiently focused through the skimmer, but some are lost to the inner wall of the skimmer due to the field divergence. The larger ions are less affected by the field and are less well focused. The model predicted sampling efficiencies of 68% and 38% for these two extremes of m/z, values that agree roughly with our measurements of transmission efficiency (unpublished). The model solves the Navier–Stokes equation for the gas flow, and the Laplace equation for the electric field, and then uses a drag model for ion motion through the gas (with measured or estimated ion-neutral cross-sections). The model also allows an assessment of ion temperature by calculating the heating of the ion due to its drag through the gas. While probably only qualitatively correct, the results clearly show that the regions of strong heating are concentrated near the edge of the skimmer, with much less heating in the center of the gas flow. It is this heating, controlled by the voltage difference between the orifice and the skimmer, that provides for declustering and even fragmentation in this region. The obvious inhomogeneity in the ion temperature shown by the model suggests that orificeskimmer fragmentation is not a very precise technique for controlling the degree of ion heating and fragmentation. 2. The High Pressure RF Ion Guide To increase the sampling efficiency the atmosphere-to-vacuum orifice diameter can be increased. This either requires using a larger mechanical pump to maintain the interface pressure at about 1 Torr, or else it requires focusing ions at a higher pressure. Focusing ions into the skimmer becomes more difficult when the gas density is higher, so the efficiency decreases when the pressure increases. Additionally, increasing the aperture diameter causes the gas flow through the skimmer to become higher, resulting in a pressure rise in the next chamber. Consideration of these issues led to the development of an rf-only quadrupole (the QJet1) to focus ions from the orifice into the next vacuum stage. An orifice diameter of 0.62 mm results in a vacuum pressure of 3.4 Torr for a reasonable sized mechanical Mass Spectrometry Reviews DOI 10.1002/mas ATMOSPHERIC PRESSURE ION SOURCES pump with a pumping speed of 11 L/sec. The high pressure requires that RF voltages between the rods be kept at a reasonably low value (<300 V) to avoid the possibility of electrical breakdown. The CFD gas flow and ion trajectory model described above was used to show that ions are largely entrained in the free jet and therefore confined within the boundaries of the barrel shock during the initial expansion. The diameter of the rfonly quadrupole was therefore selected to be slightly larger than the diameter of the barrel shock. Ions expanding into the volume of the ion guide are captured by the rf field after the initial pressure decrease in the jet expansion, while the neutral gas is allowed to escape between the rods. The ions are eventually collisionally focused by the action of the gas pressure and rf field (see review by Douglas in this volume for a description of collisional focusing). This decreases the diameter of the ion beam so that it efficiently passes through a small aperture into the next vacuum chamber. Figure 17a shows an example of the neutral gas flow expanding from the orifice into the chamber with an rf-only ion & guide at a pressure of 2.3 Torr. Although the rods disturb the development of the barrel shock and Mach disc, the supersonic portion of the jet can be observed by the red color in the figure. The yellow and green colors indicate sub-sonic gas velocity along the axis until the end of the rods where most of the flow diverges, with some flow continuing into the next (lowerpressure) chamber. Not visible in this view is the structure of the gas flow between the rods (to the pumping port located on the side wall). Figure 17b shows the ion trajectories calculated for a singly charged ion of m/z 609 under the action of the gas flow and rf field with 150 V peak-to-peak applied between the rods that are 4 mm apart. This view shows the plane between rods, where the gas can exit toward the pumping port. While some ions from the edge of the expansion are lost by being pulled by the gas flow in between the rods, it is clear that a large proportion of ions are collisionally focused toward the axis by the rf field. We have measured efficiencies of 50% for the ion transport though a 1.5 mm aperture at the end of the ion guide, comparing the ion current entering through the orifice to the ion current measured in the following chamber. This geometry therefore provides significant advantage over the orifice/skimmer combination when the gas flow and pressure in the first vacuum chamber are high. Some additional focusing and declustering can usually be achieved by applying a voltage difference of 100–200 V (dc) between the rods and the orifice, providing a similar function as the voltage difference between the orifice and skimmer in previous geometries. This voltage is empirically optimized (either manually or automatically under computer control) to provide the best sensitivity and/or signal to noise ratio. 3. The Ion Funnel and Other Ring Guides FIGURE 17. a: CFD model of 5 cm-long QJet1 with free jet expanding through 0.6 mm orifice into a pressure of 2.3 Torr. The red color indicates the region of fastest flow. b: Calculated trajectories of ions of m/z 609 through the QJet1 with an RF amplitude of 150 V on the rods. Mass Spectrometry Reviews DOI 10.1002/mas Focusing and confinement of ions in rf-only devices is not limited to the use of a quadrupole. Hexapoles and octapoles are also widely used as both collision cells and ion guides (see Douglas review in this volume). Compared to a quadrupole, these higher order multipoles provide stronger containment fields for the same voltage and frequency applied to the rods, but a broader pseudopotential well that results in a less tightly focused beam. Therefore they can be advantageous for confining ions that have high radial energies due either to the energy distribution of the incoming beam or to scattering from collisions inside the multipole. The principle of rf-focusing can be broadly described (Gerlich, 1992) as due to the action of a spatially inhomogeneous oscillating field. It can be shown mathematically that the rf field produces a net force on ions that is directed toward regions where the rf-fields are weakest (the force being identical in direction for both positive and negative ions). This force can be thought of as due to the action of a pseudo-potential created by the gradient of the rf-field. In the case of a multipole the rf-field strength is highest at large radial distances (but within the boundaries of the rod set) and weakest on the axis (where the field is always zero in a perfect quadrupole field). To first order the pseudo-potential has the form of a harmonic potential that causes ions to oscillate in the x- and y-directions as they move along the axis of the guide, with the combined x-and y-motions acting to produce rotational motion. Superimposed on this secular motion is the micromotion 891 & COVEY, THOMSON, AND SCHNEIDER at the rf frequency (Thomson, 1998a). The general principle, however, allows the possibility of a variety of electrode structures that can function as rf ion guides. Gerlich (1992) pioneered the use of rf-ring guides as containment devices with a pseudopotential that is very steep near the electrodes but very flat in the center. The ring guide consists of a series of plates with apertures that form the walls of the ring guide, with opposite rf phases applied to alternate rings. The device can be thought of as a tunnel with corrugated walls. If the rings are widely spaced, small trapping regions are formed between the rings. With the correct ring spacing, the walls appear relatively smooth and the trapping regions are very small. The rf-ring guide was adapted by the Smith group at Pacific Northwest National Laboratories to form a high-pressure ion focusing device by arranging decreasing ring aperture diameters along the axis, forming a funnel structure (Schaffer et al., 1998) as shown in Figure 18. Termed an ion funnel, this structure has been used in a variety of applications to accept a widely dispersed ion beam and focus it to a smaller diameter, usually to introduce the ions through a small aperture into the next vacuum chamber. Because reasonably small rf voltages are required when the rings are spaced closely together, the device has been successfully used to efficiently focus ions at pressures up to at least 5 Torr (Kim et al., 2000). The structure is ideally arranged to allow an axial electric field to be applied, simply by providing a resistive divider network between the rings and a single voltage source to provide decreasing DC voltages on each ring. For a fixed rf voltage, a wide mass range of ions is stably confined. The device does have the disadvantage of being more mechanically and electrically complex than a multipole, but its advantages (primarily being more efficient at focusing ions at higher pressures) have been significant enough to see adoption in some instrument configurations. For example, it has been used to effect by the Smith group to refocus ions at the end of an ion mobility drift tube that operates at a few torr pressure, after the ion beam has expanded due to diffusion (Tang et al., 2005). Multiple ion funnels have been used between stages of tandem ion mobility cells, and ultimately to focus ions into the aperture of the mass spectrometer vacuum chamber. Another adaptation of the ring guide has been developed and commercialized by Waters (Pringle et al., 2006). They used a straight ring guide, but applied a traveling wave to the plates to move ions through the ring guide. Instead of a dc axial field, the traveling wave uses a moving potential front to push ions along. The ring guide prevents ions from escaping radially. The traveling wave allows ions to ‘‘surf,’’ moving on the wave front or falling into the next trough to be pushed by the next wave front. This novel design has found use as a low-resolution mobility separation device between the quadrupole and TOF section of a QTOF, providing additional separation that is a function of charge state and collision cross-section. Other rf-only ion guide structures have been proposed and tested, but all provide the same basic function of confining ions radially while allowing them to move axially (or pushing them axially with axial electric fields), and they have all proven to be efficient at pressures of several torr. This makes them useful as devices to efficiently transfer ions from a high pressure ion source into vacuum. Many instruments employ multiple rf ion guides in series, usually at different pressures. Segmented rf multipoles have been used as collision cells, ion guides, ion mobility drift cells (Dodonov et al., 1997; Javahery & Thomson, 1997). Mass spectrometers with an atmospheric pressure ion source almost always use one or more rf ion guides to transfer ions through the higher pressure vacuum stages. FIGURE 18. The ion funnel configuration in cross-section. Each plate is a disc with a central hole as shown. Two phases of rf voltage are applied to adjacent rings. 892 Mass Spectrometry Reviews DOI 10.1002/mas ATMOSPHERIC PRESSURE ION SOURCES VII. CONCLUSION The number of topics addressed in this review attests to the growth in interest in atmospheric pressure ionization techniques since 1991, and the increase in the number of academic and industrial researchers that are now actively working to further the development and application of atmospheric pressure ionization mass spectrometry instrumentation and methods. The techniques of electron and chemical ionization that were still dominant in the early 1990s have become routine tools and commodity items with little growth or active development. The early predictions about the utility of API due to the seeming simplicity of having the ‘‘ion source in the laboratory’’ are now true, and the growth is continuing. Of course, having the ion source at atmospheric pressure where the samples originate solves some problems but raises others that we have touched on here—issues to do with the extreme sensitivity toward ubiquitous impurities that contribute to background chemical noise and raise detection limits; issues of how to efficiently transport ions into the vacuum system; issues of how best to ionize the enormous diversity of compounds that are of interest to the biological, environmental and industrial communities. These challenges will continue to drive the development of new and novel instruments and techniques in the foreseeable future, hopefully with a ‘‘change-the-game’’ outcome that will rival that experienced in the last 20 years. REFERENCES Abian J. 1999. The coupling of gas and liquid chromatography with mass spectrometry. J Mass Spectrom 34:157–161. Alexandrov ML, Gall LN, Krasnov NV, Nikolaev VI, Pavlenko VA, Shkurov VA. 1984. Extraction of ions from solution under atmospheric pressure as a method for mass spectrometric analysis of bioorganic compounds. Dolk Adak Nauk SSSR. 277, 379–383. English translation re-published in Rapid Commun Mass Spectrom 2008 22: 267–270. Analytica Chimica Acta. 2000. (entire issue). 406(2) 119–336. Arpino P, Baldwin MA, McLafferty FW. 1974. Liquid chromatography mass spectrometry-continuous monitoring. Biomed Mass Spectrom 1:80– 82. Asano KG, Ford MJ, Tomkins BA, Van Berkel GJ. 2005. Self aspirating atmospheric pressure chemical ionization source for direct sampling of analytes on surfaces and in liquid solutions. Rapid Commun Mass Spectrom 19:2305–2312. Baker TR, Pinkston JD. 1998. Development and application of packedcolumn supercritical fluid chromatography/pneumatically assisted electrospray mass spectrometry. J Am Soc Mass Spectrom 9:498–509. Baldwin MS, McLafferty FW. 1973. The direct introduction of liquid solutions into a chemical ionization mass spectrometer. Org Mass Spectrom 7:1111–1112. Banks JF, Quinn JP, Whitehouse CM. 1994. LC/ESI-MS determination of proteins using conventional liquid chromatography and ultrasonically assisted electrospray. Anal Chem 66:3688–3695. Barber M, Bordoli RS, Elliott GJ, Sedgwick RD, Tyler AN. 1982. Fast atom bombardment mass spectrometry. Anal Chem 54(4):645A–657A. Barnidge DR, Nilsson S, Markides KE. 1999. A design for low-flow sheathless electrospray emitters. Anal Chem 71:4115–4118. Barnidge DR, Nilsson S, Markides KE, Rapp H, Hjort K. 1999. Metallized sheathless electrospray emitters for use in capillary electrophoresis orthogonal time-of-flight mass spectrometry. Rapid Commun Mass Spectrom 13:994–1002. Mass Spectrometry Reviews DOI 10.1002/mas & Barroso MB, de Jong AP. 1999. Sheathless preconcentration-capillary zone electrophoresis-mass spectrometry applied to peptide analysis. J Am Soc Mass Spectrom 10:1271–1278. Bateman KB, White RL, Thibault P. 1997. Disposable emitters for on-line capillary zone electrophoresis/nanoelectrospray mass spectrometry. Rapid Commun Mass Spectrom 11:307–315. Bernardin JD, Stebbins CJ, Mudawari I. 1997. Effects of surface roughness on water droplet impact history and heat transfer regimes. Int J Heat Mass Transfer 40(1):73–88. Bi H, Hoffman KL, Pace G, Rossi DT. 2000. Mixed-mechanism ionization to enhance sensitivity in atmospheric pressure ionization LC/MS. J Pharm Biomed Analysis 22(5):861–867. Blakley CR, Carmody JJ, Vestal ML. 1980. Liquid-chromatograph-mass spectrometer for analysis of nonvolatile samples. Anal Chem 52:1636. Boyd RK, Basic C, Betham RA. 2008. Trace quantitative analysis by mass spectrometry. West Sussex, England: John Wiley & Sons Ltd. p. 748. Brown KT, Flaming DG. 1992. Advanced micropipette techniques for cell physiology. New York: John Wiley & Sons. p. 296. Bruins AP. 1991. Mass spectrometry with ion sources operating at atmospheric pressure. Mass Spectrom Rev 10:53–77. Bruins AP, Covey TR, Henion JD. 1987. Ion spray interface for combined liquid chromatography/atmospheric pressure ionization mass spectrometry. Anal Chem 59:2642–2646. Campargue R. 1984. Progress in overexpanded supersonic jets and skimmed molecular beams in free-jet zones of silence. J Phys Chem 88:4466– 4474. Cao P, Moini P. 1997. A novel sheathless interface for capillary electrophoresis/electrospray ionization mass spectrometry using an incapillary electrode. J Am Soc Mass Spectrom 8:561–564. Caproli RM, Fan T, Cottrell JS. 1986. Continuous-flow sample probe for fast atom bombardment mass spectrometry. Anal Chem 58:2949–2954. Carroll CI, Dzidic RN, Stillwell RN, Horning MG, Horning EC. 1974. Subpicogram detection system for gas phase analysis based in atmospheric pressure ionization mass spectrometry. Anal Chem 46(6):706. Chan NWC, Lewis D, Hewko S, Hindsgaul O, Schriemer DC. 2002. Screening of mixtures by frontal affinity chromatography. Comb Chem High Throughput Screen 5(5):395–406. Chandra S, Avedisian CT. 1991. On the collision of a droplet with a solid surface. Proc R Soc Lond A 432:13–41. Chen H, Venter A, Cooks RG. 2006. Extractive electrospray ionization for direct analysis of undiluted urine, milk and other complex mixtures without sample preparation. Chem Commun 2042–2044. Chowdhury SD. 1991. Method for the electrospray ionization of highly conductive aqueous solutions. Anal Chem 63:1660–1662. Cloupeau M, Prunet-Foch B. 1990. Electrostatic spraying of liquids. Main functioning modes. J Electrostatics 23:165–184. Cody RB, Laramee JA. 2005. Versatile new ion source for the analysis of materials in open air under ambient conditions. Anal Chem 77:2297– 2302. Cole RB. 1997. Electrospray ionization mass spectrometry: Fundamentals, instrumentation, and applications. New York: John Wiley & Sons. p. 577. Collings BA. 2007. Fragmentation of ions in a low pressure linear ion trap. J Am Soc Mass Spectrom 18:1459–1466. Constapel M, Schellentrager M, Schmitz OJ, Gab S, Brockmann KJ, Giese R, Benter TH. 2005. Atmospheric-pressure laser ionization: A novel ionization method for liquid chromatography/mass spectrometry. Rapid Commun Mass Spectrom 19:326–336. Cooks RG, Ouyang Z, Takats Z, Wiseman JM. 2006. Ambient mass spectrometry. Science 311:1566–1570. Coon JJ, McHale KJ, Harrison WW. 2002. Atmospheric pressure laser desorption/chemical ionization mass spectrometry: A new ionization 893 & COVEY, THOMSON, AND SCHNEIDER method based on existing themes. Rapid Commun Mass Spectrom 16:681–685. Costa AB, Cooks RG. 2007. Simulation of atmospheric transport and droplet thin-film collisions in desorption electrospray ionization. Chem Comm 3915–3917. Covey TR. 1995. Analytical characterisitics of the electrospray ionization process. In: AP, Snyder editor. Biochemical and Biotechnological Applications of Electrospray Ionization Mass Spectrometry - ACS Symposium Series 619. Washington, DC: American Chemical Society. pp. 21–59. Covey TR. 2007. Pneumatically assisted electrospray. In: Gross M, Caprioli R, editors. Encyclopedia of mass spectrometry, Vol. 6. Ionization methods. Oxford: Elsevier Ltd. pp. 426–435. Covey TR, Pinto D. 2002. Nanoelectrospray development. In: Pramanik BN, Ganguly AK, Gross ML, editors. Applied electrospray mass spectrometry. New York: Marcel Dekker, Inc. pp. 103–148. Covey TR, Lee ED, Henion JD. 1986. High-speed liquid chromatography/ tandem mass spectrometry for the determination of drugs in biological samples. Anal Chem 58:2453–2460. Covey TR, Bonner RF, Shushan BI, Henion J. 1988. The determination of protein, oligonucleotide, and peptide molecular weights. Rapid Commun Mass Spectrom 2(11):249–256. Covey TR, Schneider BB, Javaheri H, LeBlanc JCY, Ivosev G, Corr JJ, Kovarik P. 2009. ESI, APCI, and MALDI. A comparison of the central analytical figures of merit: Sensitivity, reproducibility, and speed. In: Cole RB, editor. Electrospray and MALDI mass spectrometry: Fundamentals, instrumentation, practicalities, and biological applications. Hoboken: John Wiley & Sons, Inc. Cristoni S, Rossi-Bernardi l, Guidugli F, Tubaro M, Traldi P. 2005. The role of different phenomena in surface—Activated chemical ionization (SACI) performance. J Mass Spectrom 40:1550–1557. Davis MT, Lee TD. 1997. Variable flow liquid chromatography-tandem mass spectrometry and the comprehensive analysis of complex protein digest mixtures. J Am Soc Mass Spectrom 8:1059–1069. Davis MT, Stahl DC, Lee TC. 1995a. Low-flow high performance liquid chromatography solvent delivery system design for tandem capillary liquid chromatography-mass spectrometry. J Am Soc Mass Spectrom 6:571–577. Davis MT, Stahl DC, Hefta SA, Lee TD. 1995b. A microscale electrospray interface for on-line, capillary liquid chromatography/tandem mass spectrometry of complex protein mixtures. Anal Chem 67:4549– 4566. De la Mora JF, Loscertales IG. 1994. The current transmitted through an electrified conical meniscus. J Fluid Mech 260:155–184. Deng W, Klemic JF, Li X, Reed MA, Gomez A. 2006. Increase of electrospray throughput using multiplexed microfabricated sources for the scalable generation of monodisperse droplets. J Aerosol Sci 37:696–714. Dodonov A, Kozlovsky V, Loboda A, Raznikov V, Sulimenkov I, Tolmachev A, Kraft A, Wollnick H. 1997. A new technique for decomposition of selected ions in molecule ion reactor coupled with ortho-time-of-flight mass spectrometer. Rapid Commun Mass Spectrom 11:1649–1656. Dole M, Mack LL, Hines RL, Mobley RC, Ferguson LD, Alice MB. 1968. Molecular beams of macroions. J Phys Chem Phys 49(1):2240–2249. Eberlin MN. 2006. Structurally diagnostic ion/molecule reactions: Class and functional-group identification by mass spectrometry. J Mass Spectrom 41:141–156. El-Faramawy A, Siu M, Thomson BA. 2005. Efficiency of nano-electrospray ionization. J Am Soc Mass Spectrom 16:1702–1707. Emmett MR, Caprioli RM. 1994. Micro-electrospray mass spectrometry: Ultra-high sensitivity analysis of peptides and proteins. J Am Soc Mass Spectrom 5:605–613. Feng B, Smith RD. 2000. A simple nanoelectrospray arrangement with controllable flowrate for mass analysis of submicroliter protein samples. J Am Soc Mass Spectrom 11:94–99. 894 Fenn JB, Mann M, Meng CK, Wong SF, Whitehouse CM. 1989. Electrospray ionization for mass spectrometry of large biomolecules. J Biol Sci 246:64–70. Figeys DI, Oostveen I, Ducret A, Aebersold R. 1996. Protein identification by capillary zone electrophoresis/microelectrospray ionization-tandem mass spectrometry at the subfemptomole level. Anal Chem 68(11): 1822–1828. Fong KWY, Chan TWD. 1999. A novel nonmetallized tip for electrospray mass spectrometry at nanoliter flow rate. J Am Soc Mass Spectrom 10: 72–75. French JB, Thomson BA, Davidson WR, Reid NM, Buckley JA. 1984. Atmospheric pressure chemical ionization mass spectrometry. In: Karasek FW, Hutzinger O, Safe SH, editors. Mass spectrometry in environmental sciences. New York: Plenum. pp. 101–121. French JB, Etkin B, Jong R. 1994. Monodisperse dried microparticulate injector for analytical instrumentation. Anal Chem 66:685–691. Gale DC, Smith RD. 1993. Small volume and low flow-rate electrospray ionization mass spectrometry of aqueous samples. Rapid Commun Mass Spectrom 7:1017–1020. Gamero-Castano M, Fernandez de la Mora J. 2000. Mechanisms of electrospray ionization of singly and multiply charged salt clusters. Anal Chim Acta 406(1):93–104. Ganan-Calvo AM, Davila J, Barrero A. 1997. Current and droplet size in the electrospraying of liquids. Scaling laws. J Aerosol Sci 28(2):249–275. Gatlin CL, Kleemann GR, Hays LG, Link AJ, Yates JR. 1998. Protein identification at the low femtomole level from silver-stained gels using a new fritless electrospray interface for liquid chromatography-microspray and nanospray mass spectrometry. Anal Biochem 263:93–101. Gatlin CL, Eng JK, Cross ST, Detter JC, Yates JR. 2000. Automated identification of amino acid sequence variations in proteins by HPLC/ Microspray tandem mass spectrometry. Anal Chem 72:757–763. Gelpi E. 2002. Interfaces for coupled liquid-phase separation/mass spectrometry techniques. An update on recent developments. J Mass Spectrom 37:241–253. Gerlich D. 1992. Inhomogeneous rf fields. A versatile tool for the study of processes with slow ions. In: Ng CY, Baer M, editors. State-selected and state-to-state ion molecule reaction dynamics, Part 1: Experiment. Advances in chemical physics series. Vol. LXXXII. New York: John Wiley and Sons, Inc. pp. 1–176. Geromanos S, Philip J, Frekleton G, Tempst P. 1998. Injection adaptable fine ionization source (JaFIS) for continuous flow nano-electrospray. Rapid Commun Mass Spectrom 12:551–556. Geromanos S, Freckleton G, Tempst P. 2000. Tuning of an electrospray ionization source for maximum peptide-ion transmission into a mass spectrometer. Anal Chem 72:777–790. Girvan R. 2003. Fluid Dynamics. Scientific Computing World March/April: 55–58. Guevremont R, Purves RW. 1999. Atmospheric pressure ion focusing in a high-field asymmetric waveform ion mobility spectrometer. Rev Sci Instrum 70:1370–1383. Guo X, Bruins AP, Covey TR. 2006. Characterization of typical chemical background interferences in atmospheric pressure ionization liquid chromatography-mass spectrometry (API LC-MS). R Comm Mass Spectrom 20:3145–3150. Guo X, Bruins AP, Covey TR. 2007. Method to reduce chemical background interference in atmospheric pressure ionization liquid chromatographymass spectrometry using exclusive reactions with the chemical reagent dimethyl disulfide. Anal Chem 79:4013–4021. Haddad R, Sparrapan R, Eberlin MN. 2006. Desorption sonic spray ionization for (high) voltage-free ambient mass spectrometry. Rapid Commun Mass Spectrom 20:2901–2905. Hannis JC, Muddimann DC. 1998. Nanoelectrospray mass spectrometry using non-metalized tapers fused silica capillaries. Rapid Commun Mass Spectrom 12:443–448. Mass Spectrometry Reviews DOI 10.1002/mas ATMOSPHERIC PRESSURE ION SOURCES Harrison AG. 1992. Chemical ionization mass spectrometry. Boca Raton: CRC Press. p. 224. Hatta N, Fujimoto H, Takuda H, Kinoshita K, Takahashi O. 1995. Collision dynamics of a water droplet impinging on a rigid surface above the leidenfrost temperature. ISIJ Int 35(1):50–55. Henion JD. 1978. Drug analysis by continuously monitored liquid chromatography/mass spectrometry with a quadrupole mass spectrometer. Anal. Chem 50(12):1887. Henion JD, Thomson BA, Dawson PH. 1982. Determination of sulfa drugs in biological fluids by liquid chromatography/mass spectrometry/mass spectrometry. Anal Chem 54:451–4456. Hiller DL, Brockman AH, Goulet L, Ahmed S, Cole RO, Covey TR. 2000. Application of a non-indexed dual sprayer pneumatically assisted electrospray source to the high throughput quantitation of target compounds in biological fluids. Rapid Commun Mass Spectrom 14:2034–2038. Hirabayashi A, Sakari M, Koizumi H. 1994. Sonic spray ionization method for atmospheric pressure ionization mass spectrometry. Anal Chem 66:4557–4559. Hong CM, Tsai FC, Shiea J. 2000. A multiple channel electrospray source used to detect highly reactive ketenes from a flow pyrolyzer. Anal Chem 72:1175–1178. Hopfgartner G, Bourgogne E. 2003. Quantitative high-throughput analysis of drugs in biological matrices by mass spectrometry. Mass Spectrom Rev 22:195–214. Horning EC, Carroll DI, Dzidic I, Haegele KD, Horning MG, Stillwell RN. 1974. Liquid chromatograph-mass spectrometer-computer analytical system: Continuous flow system based on atmospheric pressure ionization mass spectrometry. J Chromatogr 99:13–21. Hsich F, Baronas E, Muir C, Martin SA. 1999. A novel nanospray capillary zone electrophoresis/mass spectrometry interface. Rapid Commun Mass Spectrom 13:67–72. Issac K. 1994. Atomization experiments in a coaxial co-flow Mach 1 nebulizer. AIAA J 32(8):1640–1646. Ito Y, Takeuchi T, Goto M, Ishii D. 1985. Direct coupling of micro highperformance liquid chromatography with fast atom bombardment mass spectrometry. J Chromatogr 346:161–166. Jaskolla TW, Lehmann W, Karas M. 2008. 4-Chloro-a-cyanocinnamic acid is an advanced, rationally designed MALDI matrix. Proc Natl Acad Sci USA 105:12200–12205. Javahery G, Thomson BA. 1997. A segmented radiofrequency-only quadrupole collision cell for measurements of ion collision cross section on a triple quadrupole mass spectrometer. J Am Soc Mass Spectrom 8:697– 702. Jiang L, Moini M. 2000. Development of multi-ESI-sprayer, multiatmospheric-pressure-inlet mass spectrometry and its application to accurate mass measurement using time-of-flight mass spectrometry. Anal Chem 72(1):20–24. Jugroot M, Groth CPT, Thomson BA, Baranov V, Collings BA. 2004a. Numerical investigation of interface region flows in mass spectrometers: Neutral gas transport. J Phys D Appl Phys 37:1289– 1300. Jugroot M, Groth CPT, Thomson BA, Baranov V, Collings BA. 2004b. Numerical investigation of interface region flows in mass spectrometers: Ion transport. J Phys D Appl Phys 37:550–559. Jugroot M, Groth CPT, Thomson BA, Baranov V, Collings BA, French JB. 2008. Coupled gas and ion transport in quadrupole interfaces. J Phys D Appl Phys 41: 025205 (10pp). Juraschek R, Dulcks T, Karas M. 1999. Nanelectrospray—more than just a minimized-flow electrospray ionization source. J Am Soc Mass Spec 10(4):300–308. Karas M, Kruger R. 2003. Ion formation in MALDI: The cluster ionization mechanism. Chem Rev 103:427–439. Mass Spectrometry Reviews DOI 10.1002/mas & Kebarle P, Tang L. 1993. From ions in solution to ions in the gas phase. Anal Chem 65:972A–986A. Keller BO, Sui J, Young AB, Whittal RM. 2008. Interferences and contaminants encountered in modern mass spectrometry. Anal Chim Acta 627(1):71–81. Kelly JF, Ramaley L, Thibault P. 1997. Capillary zone electrophoresiselectrospray mass spectrometry at submicroliter flow rates: Practical considerations and analytical performance. Anal Chem 69: 51–60. Kelly R, Page JS, Tang K, Smith R. 2008. Array of chemically etched fused silica emitters for improving the sensitivity and quantitation of electrospray ionization mass spectrometry. Anal Chem 79(11):4192– 4198. Kertesz V, Ford MJ, Van Berkel GJ. 2005. Automation of a surface sampling probe/electrospray mass spectrometry system. Anal Chem 77(22): 7183–7189. Kim T, Tolmachev AV, Harkewicz R, Prior DC, Anderson GA, Udseth HR, Smith RD, Bailey TH, Rakov S, Futrell JH. 2000. Design and implementation of a new electrodynamic ion funnel. Anal Chem 72: 2247–2255. King R, Bonfiglio R, Fernandez-Metzler C, Miller-Stein C, Olah T. 2000. Mechanistic investigation of ionization suppression in electrospray ionization. J Am Chem Soc 11:942–950. Kostiainen R, Bruins AP. 1994. A multiple sprayer study of dynamic range and flow rate limitations in electrospray and ion spray mass spectrometry. Rapid Commun Mass Spectrom 2:549–558. Kriger MS, Cook KD, Ramsey RS. 1995. Durable gold coated fused silica capillaries for use in electrospray mass spectrometry. Anal Chem 67:385–389. Ku BK, Kim SS, Kim YD, Lee SY. 2001. Direct measurement of electrospray droplets in the submicron diameter using a freezing method and a TEM image processing. J Aerosol Sci 32(12):1459– 1477. Laiko VV, Baldwin MA, Burlingame AL. 2000. Atmospheric pressure matrix-assisted laser desorption/ionization mass spectrometry. Anal Chem 72:652–657. Lane DA, Thomson BA. 1981. Monitoring a chlorine spill from a train derailment. Air Poll Control Assoc J 31:122–127. Lee MS. 2002. LC/MS applications in drug development. New York: John Wiley & Sons, Inc. p. 256. Lee MS. 2005. Integrated strategies for drug discovery using mass spectrometry. New York: John Wiley & Sons, Inc. p. 568. Lee CY, Shiea J. 1998. Gas chromatography connected to multiple channel electrospray ionization mass spectrometry for the detection of volatile organic compounds. Anal Chem 70:2757–2761. Leuthold LA, Grivet C, Allen M, Baumert M, Hopfgartner G. 2004. Simultaneous selected reaction monitoring, MS/MS and MS3 quantitation for the analysis of pharmaceutical compounds in human plasma using chip-based infusion. Rapid Commun Mass Spectrom 18:1995– 2000. Levin DS, Vouros P, Miller RA, Nazarov EG, Morris JA. 2006. Characterization of gas-phase molecular interactions on differential mobility ion behavior utilizing an electrospray ionization-differential mobility-mass spectrometer system. Anal Chem 78:96–106. Lias S, Bartness JE. 2008. Gas-Phase Ion Thermochemistry Data Base. National Institute of Standards and Technology: http://webbook.nist. gov/chemistry/ion/. McEwen CN, McKay RG, Larsen BS. 2005. The analysis of solids, liquids, and biological tissues using solids probe introduction at atmospheric pressure on commercial LC/MS instrumentation. Anal Chem 77:7826– 7831. McFadden WH, Schwartz HL, Evans S. 1976. Direct analysis of liquid chromatographic effluents. J Chromatogr 122:296–386. 895 & COVEY, THOMSON, AND SCHNEIDER Min H, Emory JF, McLuckey SA. 2005. Reagent anions for charge inversion of polypeptide/protein cations in the gas phase. Anal Chem 77:3173– 3182. Moini M, Xia M, Stewart JB, Hofmann B. 1998. Atmospheric pressure microwave induced plasma ionization source for molecular mass spectrometry. J Am Soc Mass Spectrom 9:42–49. Neidholdt EL, Beauchamp JL. 2007. Compact ambient pressure pyroelectric ion source for mass spectrometry. Anal Chem 79(10):3945–3948. Nemes P, Vertes A. 2007. Laser ablation electrospray ionization for atmospheric pressure, in vivo, and imaging mass spectrometry. Anal Chem 79:8098–8106. Ogorzalek Loo RR, Udseth HR, Smith RD. 1991. Evidence of charge inversion in the reaction of singly charged anions with multiply charged macroions. J Phys Chem 95:6412–6415. Ogorzalek Loo RR, Loo JA, Udseth HR, Fulton JL, Smith RD. 1992. Protein structural effects in gas phase ion/molecule reactions with diethyl amine. Rapid Commun Mass Spectrom 6:159–165. Ogorzalek Loo RR, Udseth HR, Smith RD. 1992. A new approach for the study of gas-phase ion-ion reactions using electrospray ionization. J Am Soc Mass Spectrom 3:695–705. Ostman P, Marttila SJ, Kotiaho T, Franssila S, Kostiainen R. 2004. Microchip atmospheric pressure chemical ionization source for mass spectrometry. Anal Chem 76:6659–6664. Page JS, Kelly RT, Tang K, Smith RD. 2007. Ionization and transmission efficiency in an electrospray ionization-mass spectrometry interface. J Am Soc Mass Spectrom 18:1582–1590. Park H, Kim K, Kim S. 2004. Effects of a guard plate on the characteristics of an electrospray in the cone-jet mode. J Aerosol Sci 35(11):1295– 1312. Pramanik BN, Ganguly AK, Gross ML, 2002. Applied electrospray mass spectrometry. New York: Marcel Dekker, Inc. Pringle SD, Giles K, Wildgoose JL, Williams JP, Slade SE, Thalassinos K, Bateman RH, Bowers MT, Scrivens JH. 2006. An investigation of the mobility separation of some peptide and protein ions using a new hybrid quadrupole/travelling wave IMS/oaToF instrument. Int J Mass Spectrom 261:1–12. Purves RW, Guevremont R. 1998. Mass spectrometric characterization of a high-field asymmetric waveform ion mobility spectrometer. Rev Sci Instrum 69(12):4094–4105. Ratcliffe LV, Rutten FJM, Barrett DA, Whitmore T, Seymour D, Greenwood C, Aranda-Gonzalvo Y, Robinson S, McCoustra M. 2007. Surface analysis under ambient conditions using plasma-assisted desorption/ionization mass spectrometry. Anal Chem 79:6094– 6101. Robb DB, Covey TR, Bruins AP. 2000. Atmospheric pressure photoionization: An ionization method for liquid chromatography/mass spectrometry. Anal Chem 72:3653–3659. Rulison A, Flagan RC. 1993. Scale-up of electrospray atomization using linear arrays of Taylor cones. Rev Sci Instrum 64(3):683–686. Sampson JS, Hawkridge AM, Muddiman DC. 2006. Generation and detection of multiply-charged peptides and proteins by matrix-assisted laser desorption electrospray ionization (MALDI ESI) FTICR MS. J Am Soc Mass Spectrom 17:1712–1716. Schaefer WH, Dixon F, Jr. 1996. Effect of HPLC mobile phase components on sensitivity in Neg APCI LCMS. J Am Soc Mass Spectrom 7:1059– 1069. Schaffer SA, Prior DC, Anderson GA, Udseth HR, Smith RD. 1998. An ion funnel interface for improved ion focusing and sensitivity using electrospray ionization mass spectrometry. Anal Chem 70:4111– 4119. Schmidt A, Karas M, Dülcks T. 2003. Effect of different solution flow rates on analyte ion signals in nano-ESI MS, or when does ESI turn into nanoESI. J Am Soc Mass Spectrom 14:492–500. 896 Schneider BB, Covey TR. 2007. Multiple sprayers in electrospray ionization. In: Gross M, Caprioli R, editors. Encyclopedia of mass spectrometry, Vol. 6. Ionization Methods. Oxford: Elsevier Ltd. pp. 456–467. Schneider BB, Douglas DJ, Chen DDY. 2002. Multiple sprayer system for high-throughput electrospray ionization mass spectrometry. Rapid Commun Mass Spectrom 16:1982–1990. Schneider BB, Baranov VI, Javaheri H, Covey TR. 2003. Particle discriminator interface for nanoflow ESI-MS. J Am Soc Mass Spectrom 14:1236–1246. Schneider BB, Guo X, Fell LM, Covey TR. 2005. Stable gradient Nanoflow LC/MS. J Am Soc Mass Spectrom 16:1545–1551. Schneider BB, Javaheri H, Covey TR. 2006. Ion sampling effects under conditions of total solvent consumption. Rapid Commun Mass Spectrom 20:1538–1544. Schultz GA, Corso TN, Prosser SJ, Zhang S. 2000. A fully integrated monolithic microchip electrospray device for mass spectrometry. Anal Chem 72:4058–4063. Shea RC, Petzold CJ, Liu J, Kenttamaa HI. 2007. Experimental investigations of the internal energy of molecules evaporated via laser/induced acoustic desorption into a FTICR MS. Anal Chem 79:1825–1832. Shen WL, Satzger RD. 1991. Development of a low power microwave atmospheric pressure molecular ionization source for mass spectrometry with direct introduction of gaseous and liquid samples. Anal Chem 63:1960–1964. Shiea J, Chang DY, Lin CH, Jiang SJ. 2001. Generating multiply charged protein ions by ultrasonic nebulization/multiple channel-electrospray ionization mass spectrometry. Anal Chem 73:4983–4987. Shiea J, Huang M, Hsu H, Lee C, Yuan C, Beech I, Sunner J. 2005. Electrospray-assisted laser desorption/ionization mass spectrometry for direct ambient analysis of solids. Rapid Commun Mass Spectrom 19: 3701–3704. Shieh I, Lee C, Shiea J. 2005. Eliminating the interferences from TRIS buffer and SDS in protein analysis by fused-droplet electrospray ionization mass spectrometry. J Proteome Res 4:606–612. Singh G, Gutierrez A, Xu K, Blair IA. 2000. Liquid chromatography/electron capture atmospheric pressure chemical ionization/mass spectrometry: Analysis of pentafluorobenzyl derivatives of biomolecules and drugs in the attomole range. Anal Chem 72:3007–3013. Snyder AP. 1995. Biochemical and biotechnological applications of electrospray ionization mass spectrometry. Washington: American Chemical Society. p. 601. Sparks JB. 1952. Histomap of world history. Chicago: Rand McNally. Syage JA, Hanning-Lee MA, Hanold KA. 2000. A man-portable photoionization time-of-flight mass spectrometer. Field Anal Chem Tech 4(4):204–215. Syage JA, Hanold KA, Lynn TC, Horner JA, Thakur RA. 2004. Atmospheric pressure photoionization: Dual source ionization. J Chromatog A 1050(2):137–149. Takats Z, Wiseman JM, Gologan B, Cooks RG. 2004. Electrosonic spray ionization. A gentle technique for generating folded proteins and protein complexes in the gas phase and for studying ion-molecule reactions at atmospheric pressure. Anal Chem 76:4050–4058. Talrose VL, Karpov GV, Gorodetskii IG, Skurat VE. 1968. Capillary system for the introduction of liquid mixtures into an analytical massspectrometer. Russ J Phys Chem 42(12):1658. Tang K, Schvartzburg AA, Lee H-N, Prior DC, Buschback MA, Li F, Tolmachev AV, Anderson GZ, Smith RD. 2005. High-sensitivity ion mobility mass spectrometry using electrodynamic ion funnel interfaces. Anal Chem 77:3330–3339. Tanner SA, Baranov VI, Bandura DR. 2002. Reaction cells and collision cells for ICP-MS: A tutorial review. Spectrochim Acta Part B 57:1361–1452. Thomson BA. 1998a. Radio frequency quadrupole ion guides in modern mass spectrometry. Can J Chem 76:499–505. Mass Spectrometry Reviews DOI 10.1002/mas ATMOSPHERIC PRESSURE ION SOURCES Thomson B. 1998b. Atmospheric pressure ionization LC/MS—Together at last. J Am Soc Mass Spectrom 9:187–193. Thomson BA. 2007a. The heated nebulizer LC-MS interface. In: Gross M, Caprioli R, editors. Encyclopedia of mass spectrometry, Vol. 6. Ionization methods. Oxford: Elsevier Ltd. pp. 366–370. Thomson BA. 2007b. Micro and nano-electrospray ionization techniques. In: Gross M, Caprioli R, editors. Encyclopedia of mass spectrometry, Vol. 6. Ionization methods. Oxford: Elsevier Ltd. pp. 435–444. Thomson BA, Irabarne JV. 1979. Field induced ion evaporation from liquid surfaces at atmospheric pressure. J Chem Phys 71(11):4451–4463. Thomson BA, Iribarne JV. 1978. The fate of electrical charges in evaporating cloud droplets. Rev Geophys Space Phys 16(3):431–434. Thomson BA, Irabarne JV, Dziedzic PJ. 1982. Liquid ion evaporation/mass spectrometry/mass spectrometry for the detection of polar and labile molecules. Anal Chem 54:2219–2224. Valaskovic GA, McLafferty FW. 1995. Attomole-sensitivity electrospray source for large-molecule mass spectrometry. Anal Chem 67:3802– 3805. Valaskovic GA, McLafferty FW. 1996a. Long-lived metallized tips for nanoliter electrospray mass spectrometry. J Am Soc Mass Spectrom 7: 1270–1272. Valaskovic GA, McLafferty FW. 1996b. Sampling error in small-bore sheathless capillary electrophoresis/electrospray-ionization mass spectrometry. Rapid Commun Mass Spec 10:825–828. Valaskovic GA, Kelleher NL, McLafferty FW. 1996. Attomole protein characterization by capillary electrophoresis-mass spectrometry. Science 372:1199–1202. Van Berkel GJ. 2007a. Electrochemistry of the electrospray ionization source. In: Gross M, Caprioli R, editors. Encyclopedia of mass spectrometry, Vol. 6. Ionization methods. Oxford: Elsevier Ltd. pp. 422–426. Van Berkel GJ, Kertesz V. 2005. Expanded electrochemical capabilities of the electrospray ion source using porous flow-through electrodes as the upstream ground and emitter high voltage contact. Anal Chem 77: 8041–8049. Van Berkel GJ, Kertesz V. 2007b. Using the electrochemistry of the electrospray ion source. Anal Chem 79:5510–5520. Van Berkel GJ, Asano KG, Schnier PD. 2001. Electrochemical processes in a ‘‘wire in a capillary’’ bulk-loaded, nano-electrospray emitter. J Am Soc Mass Spectrom 12:853–862. Van Berkel GJ, Kertesz V, Koeplinger KA, Vavrek M, Kong AT. 2008. Liquid micro-junction surface sampling probe electrospray mass spectrometry for detection of drugs and metabolites in thin tissue sections. J Mass Spectrom 43:500–508. Van Berkel GJ, Pasilis SP, Ovchinnikova O. 2008. Established and emerging atmospheric pressure surface sampling/ionization techniques for mass spectrometry. J Mass Spectrom 43:1161–1180. Vanhoutte K, Van Dongen W, Hoes I, Lemiere F, Esmans E, Van Onckelen H, Eeckhout E, Van Soest REJ, Hudson AJ. 1997. Development of a nanoscale liquid chromatography/electrospray mass spectrometry Mass Spectrometry Reviews DOI 10.1002/mas & methodology for the detection and identification of DNA adducts. Anal Chem 69:3161–3168. Vanhoutte K, Van Dongen W, Esmans EL. 1998. On-line nanoscale liquid chromatography nano-electrospray mass spectrometry: Effect of the mobile phase composition and the electrospray tip design on the performance of a nanoflowTM electrospray probe. Rapid Commun Mass Spectrom 12:15–24. Wahl JH, Gale DC, Smith RD. 1994. Sheathless capillary electrophoresis— Electrospray ionization mass spectrometry using 10 mm I.D. capillaries: Analyses of tryptic digests of cytochrome c. J Chromatogr A 659:217– 222. Wang H, Hackett M. 1998. Ionization within a cylindrical capacitor: Electrospray without an externally applied high voltage. Anal Chem 70:205–212. Wang T, Zeng L, Cohen J, Kassel DB. 1999. A multiple electrospray interface for parallel mass spectrometric analyses of compound libraries. Comb Chem High Throughput Screen 2:327–334. Wang Y, Schneider BB, Covey TR, Pawliszyn J. 2006. High-performance SPME/AP MALDI system for high-throughput sampling and determination of peptides. Anal Chem 77(24):8095–8101. Whitehouse CM, Dreyer RN, Yamashita M, Fenn JB. 1985. Electrospray interface for liquid chromatographs and mass spectrometers. Anal Chem 57(3):675–679. Willoughby RC, Browner RF. 1984. Monodisperse aerosol generation interface for combining liquid chromatography with mass spectrometry. Anal Chem 56:2626–2631. Willoughby R, Sheehan E, Mitrovich S. 2002. A global view of LC/MS. Pittsburg: Global View Publishing. p. 518. Wilm M, Mann M. 1994. Electrospray and Taylor-Cone theory, Dole’s beam of macromolecules at last? Int J Mass Spectrom Ion Processes 136:167– 180. Wilm M, Mann M. 1996. Analytical properties of the nanoelectrospray ion source. Anal Chem 68:1–8. Wilm M, Shevchenko A, Houtaeve T, Breit S, Schweigerer L, Fotsis T, Mann M. 1996. Femtomole sequencing of proteins from polyacrylamide gels by nano-electrospray mass spectrometry. Nature 379:466–469. Wolfender JL, Chu F, Ball H, Wolfender F, Fainzilber M, Baldwin MA, Burlingame AL. 1999. Identification of tyrosine sulfation in Conus pennaceus conotoxins a-PnIA and a-PnIB: Further investigation of labile sulfo- and phosphopeptides by electrospray, matrix-assisted laser desorption/ionization (MALDI) and atmospheric pressure MALDI mass spectrometry. J Mass Spectrom 34:447–454. Yamaguchi K. 2003. Cold-spray ionization mass spectrometry: Principle and the applications. J Mass Spectrom 38:473. Yamashita M, Fenn JB. 1984. Electrospray ion source. Another variation on the free-jet theme. J Phys Chem 88:4451–4459. Zhou L, Yue B, Deardon DV, Lee ED, Rockwood AL, Lee ML. 2003. Incorporation of a venturi device in electrospray ionization. Anal Chem 75:5978–5983. 897