Next Article in Journal
Electrochemical Detection of Surfactant-Encapsulated Aqueous Nanodroplets in Organic Solution
Next Article in Special Issue
Formaldehyde Gas Sensors Fabricated with Polymer-Based Materials: A Review
Previous Article in Journal
Engineering Rational SERS Nanotags for Parallel Detection of Multiple Cancer Circulating Biomarkers
Previous Article in Special Issue
Applying of C8-BTBT-Based EGOFETs at Different pH Values of the Electrolyte
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

SIFT-MS: Quantifying the Volatiles You Smell…and the Toxics You Don’t

by
Vaughan S. Langford
Syft Technologies Limited, Christchurch 8011, New Zealand
Chemosensors 2023, 11(2), 111; https://doi.org/10.3390/chemosensors11020111
Submission received: 18 January 2023 / Revised: 26 January 2023 / Accepted: 31 January 2023 / Published: 3 February 2023

Abstract

:
The human olfactory system is highly attuned to detection of a wide range of volatile organic compounds (VOCs), but the sensitivity varies considerably based on chemical functionality. Whereas most humans can appreciate the sensory properties of certain foods, beverages, and fragrances, and at times be alerted to volatile chemical hazards, many VOCs are hazardous below the human odor detection threshold. Since its introduction in the mid-1990s, selected ion flow tube mass spectrometry (SIFT-MS) has been widely applied to quantitative analysis of a broad range of VOCs in applications from food products to workplace safety to environmental monitoring, and most recently to pharmaceutical testing. This review surveys the applications of SIFT-MS in odor analysis and in workplace, environmental and consumer protection, with a particular focus on the complementarity of this real-time mass spectrometry analyzer to sensor technology and conventional laboratory techniques—in particular, gas chromatography–mass spectrometry (GC/MS).

1. Introduction

Time-resolved volatile organic compound (VOC) analysis provides insight into processes of fundamental and industrial relevance that are extremely difficult to probe using conventional gold-standard chromatographic techniques (gas chromatography (GC) and liquid chromatography (LC) [1,2]). Chromatographic methods separate components in time before detecting them, making both response and sample analysis times long—typically tens of minutes to an hour. Direct-injection mass spectrometry (DIMS) analytical techniques for analysis of VOCs in air and breath were developed in the late 1980s and 1990s to address this shortcoming [3].
Three DIMS techniques dominate direct air analysis: atmospheric pressure chemical ionization-mass spectrometry (APCI-MS), proton transfer reaction-mass spectrometry (PTR-MS), and selected ion flow tube-mass spectrometry (SIFT-MS) [3,4]. SIFT-MS stands apart from APCI-MS and PTR-MS in that—despite being the last of these techniques to be commercialized—it has proportionally been adopted to a greater extent by industry than academia. For example, SIFT-MS has been adopted in frontline workplace safety and in semiconductor cleanroom applications (see Section 4.1) and has a very good fit in routine and automated VOC analysis [5]. Since chemosensors find broad industrial application, and to ensure manageable scope, this review focuses on the SIFT-MS technique. It should, however, be noted that the other DIMS techniques could be similarly considered for their complementarity to chemosensors in many of the applications discussed.
Since SIFT-MS is based on mass spectrometry (MS), it is significantly more complex than most chemosensors—as reflected in price differentials up to three orders of magnitude (in the case of the lowest cost chemosensors). All MS techniques require that analytes be ionized. In the case of SIFT-MS, very soft chemical ionization (CI) is utilized. As with other MS techniques, once ionized, the gas-phase ion products are subsequently separated by their mass-to-charge ratio (m/z) and then counted—the ion abundance being proportional to analyte concentration. The spectral “fingerprint” generated by the MS instrument is determined by the elemental composition of the analyte and the ionization energetics. (An unusual feature of SIFT-MS is that for certain reaction mechanisms (especially association, Section 2.2) the chemical ionization agent (so the so-called “reagent ion”) may be incorporated into the detected product ion [6,7].) As will be described in the next section, this confers on SIFT-MS the ability to achieve multicomponent VOC analysis with high specificity, sensitivity, and direct quantitation with low calibration demand. In contrast, the output of chemosensors is typically an electrical signal generated in response to a VOC or class of VOCs [8], without identificatory mass information. Hence, sensory arrays are commonly used, combining sensors with varying responses to different VOCs that interact with their surface [8]. Sophisticated multivariate statistical analysis (pattern recognition) is utilized on a training set to build a model that can be used to identify and quantify patterns of analytes [8]. It should be noted, however, that greater certainty in identification can be achieved when chemosensors are used as detectors for chromatography systems (e.g., Alpha MOS, www.alpha-mos.com, accessed on 25 January 2023).
Given the very different technologies underlying SIFT-MS and chemosensors, and the distinctive analytical data that they generate, while both generally provide rapid speed of analysis, this review aims to demonstrate the complementarity of the approaches. Conventionally chemosensor performance has been benchmarked against chromatographic methods, with slow response time and time-to-result. SIFT-MS provides an alternative, generating analytical data in real time with high specificity and sensitivity. Therefore, this article is structured as follows. First, the SIFT-MS technique is introduced, focusing on the fundamental aspects that complement chemosensor analysis. Since comprehensive reviews of the SIFT-MS literature exist [6,9,10], this article focusses on two broad application areas in which SIFT-MS and chemosensors both find application (aroma/odor and toxics analysis), highlighting particularly new developments. Finally, recent applications of SIFT-MS that are more “chemosensor-like” are reviewed briefly. At all points the focus is on presenting the SIFT-MS literature in a manner that makes it accessible to experts in chemosensors, avoiding review of literature with which these experts are already familiar.

2. The SIFT-MS Technique—An Overview

This section introduces the SIFT-MS technique, providing an overview of the generic layout of instrumentation before it describes key analytical features pertinent to this article: breadth of analysis, specificity, and quantitation. More technical details can be found in Refs. [6,9].

2.1. Instrument Overview

All SIFT-MS instruments have three zones, as shown in Figure 1. In the first region, reagent ions are generated continuously in a plasma created at low pressure using a microwave discharge. For the positively charged reagent ions, H3O+, NO+, and O2+, the ion source is operated at approximately 54 kPa (400 torr), while for the negatively charged reagent ions (O, O2, OH, NO2, and NO3) it is operated at approximately twice this pressure to facilitate electron attachment. Multiple reagent ions provide two significant benefits that are elaborated below: (1) their different ionization properties enable a wide range of compounds to be detected, and (2) they are foundational to achieving specific analysis. The plasma containing a mixture of potential reagent ions is then passed into a quadrupole mass filter (QMF) that, using software control, selects the appropriate, single-reagent ion type (by its m/z; e.g., H3O+ with m/z = 19) for introduction as a pure stream of reagent ion into the second region, the reaction chamber (or flow tube).
In the flow tube, reagent ions first encounter carrier gas molecules (either helium or nitrogen, depending on instrument application). Through collisions with the carrier gas, the energies of the reagent ions are reduced to approximately the temperature of the carrier gas itself (i.e., the reagent ions are “thermalized”) [7]. The sample is introduced after reagent ions have been in the flow tube for approximately 1 ms, then has a residence time of about 3 to 8 ms (configuration-dependent) for reaction with reagent ions. At the end of the flow tube, ions are sampled into the third region (detection), while the bulk of carrier and sample mixture is pumped away to exhaust. Note that the carrier gas and the bulk matrix must have ionization properties that render them essentially non-reactive with the reagent ions. Conveniently, this is the case for air, with which the reagent ions listed above either do not react or react only very slowly (nitrogen, oxygen, argon, carbon dioxide, and water).
The detection region comprises, firstly, an ion guide that improves transmission of heavier product ions, second a QMF that transmits just the ions with a given m/z, and finally a particle multiplier detector that counts ions at each m/z. Typically operation in the linear range means that less than 10% of the reagent ion signal is consumed [6] and concentration is proportional to the ratio of product ion count divided by reagent ion count (Section 2.4). This is essentially an auto-normalization feature, correcting for any drift of ion signal, should it occur, and supports stable long-term operation.
SIFT-MS instruments are completely computer controlled and can operate autonomously or via remote control. The stability of ionization means that—when required—re-calibration is an infrequent task (annual for many compounds due to drift less than 10%), and software supports the generation of laboratory-grade analytical results for non-technical operators for properly developed analytical methods.

2.2. Breadth of Analysis

Although, as noted in the previous subsection, the SIFT-MS technique is ‘blind’ to the bulk components of air due to the low ionization energies of its reagent ions, it has remarkable breadth of analysis for VOCs and trace inorganic gases. This is due to the multiple reagent ions that ionize compounds via a wide range of ion–molecule reaction mechanisms (Table 1). These mechanisms have been reviewed in detail elsewhere [6,7]. Essentially, if the ionization properties of the target compound are a match for SIFT-MS, then it will be ionized and detected. Note, however, that unlike the electron ionization (EI) method used commonly with GC or photoionization detectors (PIDs), the different mechanisms yield a wider range of product ions, potentially increasing specificity for isobaric compounds.
Table 2 illustrates the breadth of SIFT-MS analysis by giving examples of selected compounds that are practicably detectable in air using the indicated reagent ions. Assuming that specificity is achievable using SIFT-MS (see the next subsection), virtually any combination of compounds can be detected in a single analysis. This is possible with SIFT-MS due to the diversity of the ion chemistry and the broad-spectrum ion detection provided by mass spectrometry, coupled with removal of complications imposed by the use of chromatographic columns in conventional laboratory techniques. The subsequent sections describe applications where this multicomponent analysis of gas-phase compounds with diverse ionization properties is very beneficial.
Since a substantial breadth of literature describes volatile compounds that are detectable using SIFT-MS (see, for example, very convenient review articles [6,9,10]), several examples of important compounds that are not analyzable, or have significant constraints, are discussed here.
First, consider methane which is of significant concern for climate due to its high global warming potential [27]. In positive ion mode, the O2+• reagent ion reacts with methane, as shown in Equation (1), whereas H3O+ and NO+ are too inefficient to be useful [28,29]. However, the sensitivity of SIFT-MS is lower than for most volatiles due to the slow reaction rate coefficient (k)—over 200-fold lower than typical k for VOCs. Nevertheless, methane analysis has been demonstrated for breath [29] and results for landfill and biogas analysis were reported to commercial customers by the present author over a 14-year period. Due to the low sensitivity, great care needs to be taken to ensure that the methane product ion does not suffer from interference.
O2+• + CH4 → CH2O2H+ + H k = 5 − 5.5 × 10−12 cm3 molecule−1 s−1
Interestingly, O−• reacts with methane somewhat more rapidly (Equation (2)) [30] via the hydrogen atom transfer mechanism (Table 1). However, the product ion is also formed from reaction of O−• with water [30] and other low-molecular-weight hydrocarbons [30], so it is not of practical use for specific detection of methane.
O−• + CH4 → OH + CHsk = 0.8 − 1.1 × 10−10 cm3 molecule−1 s−1
Second, carbon monoxide is toxic [31] and exposure is of significant concern in confined spaces and in the urban environment. Unfortunately, of the standard SIFT-MS reagent ions only O reacts with CO [30] and this solely via the associative detachment mechanism (Table 1). As shown in Equation (3), no useful ion is obtained since free electrons (e) are not detectable in SIFT-MS instruments.
O−• + CO → CO2 + e  k = 5.5 − 7.3 × 10−10 cm3 molecule−1 s−1
Finally, nitrous oxide (N2O)—another important greenhouse gas [27]—reacts with O as shown in Equation (4) [30]. In principle, this reaction rate coefficient provides adequate sensitivity for trace gas analysis. However, in air the NO product ion reacts with O2 to form O2 (Equation (5)) [30], which cannot be distinguished from O2 formed at trace levels from other sources. N2O is hence detectable, but it is only practicable to do so in oxygen-free matrices (e.g., in specialty gas applications).
O + N2O → NO + NO k = 1.95 − 2.5 × 10−10 cm3 molecule−1 s−1
NO + O2 → O2 + NO k = 5 × 10−10 cm3 molecule−1 s−1
Despite these limitations, SIFT-MS has found broad applications [6,9,10]. Later sections will review applications that are more closely aligned with chemosensor technologies.

2.3. Specificity of Analysis

Achieving specific analysis in moderately complex matrixes can be challenging for low-cost chemosensors. Eminent flavor chemist, Gary Reineccius wrote of the application of electronic noses (sensor arrays) in food flavor analysis that “one has no clear idea of what the instrument is responding to in making a judgement” [32].
In contrast, SIFT-MS achieves specific analysis in real-time by the combination of highly controlled, ultra-soft CI and MS, producing signals that are attributable to specific analytes. Table 1 illustrates the variety of common reaction mechanisms that occur for the standard SIFT-MS reagent ions and indicates the relative shifts from parent product ion m/z, or standard product ion m/z, for these mechanisms. Combining this behavior with the observation that different functional groups typically react in different ways [6,7], at least one unique ion for a given analyte is usually realized.
It is important to reemphasize that when using the phrase “real-time specificity” in the context of SIFT-MS, a user is not limited to a single reagent ion in each analysis. By means of the first QMF (Figure 1), combined with software control, reagent ions generated in a given ion source setting (e.g., positive ions with H3O+, NO+, and O2+) are switchable in tens of milliseconds. This means that target compounds can be analyzed specifically across different reagent ions, if necessary. This contrasts with the other DIMS techniques [3]. Note that where an ion source setting change is required—e.g., from positive ions to either of the negative ion settings—this is still near real time at approximately 10 s. Figure 2 summarizes the elements that contribute to specific analysis in SIFT-MS. Addition of a time-of-flight (TOF) mass spectrometer to SIFT-MS instruments (as has been done successfully for PTR-MS [3]) could further improve specificity, but at additional cost. However, regardless of the way that mass filtering is accomplished in SIFT-MS or PTR-MS, it is not reasonable to expect that the specificity of DIMS techniques will approach that of GC/MS with its temporal separation of analytes. GC/MS techniques—especially those utilizing high-resolution mass spectrometry—will remain the technique of choice for compound identification and analysis of very complex matrices.
Since detailed examples of SIFT-MS specificity are given in Refs. [33,34,35,36], one simple example suffices here. Consider two structural isomers of molecular formula C2H6O and molar mass 58.0791 g mol−1: acetone (propan-2-one, (CH3)2CO) and propanal (propionaldehyde, CH3CH2CHO). Table 3 shows the primary product ions formed when acetone and propanal react with the three positively charged SIFT-MS reagent ions [14]. Use of the NO+ reagent ion provides the most convenient resolution of these isomers. The H3O+ reagent ion cannot resolve the compounds, and matrix effects are encountered frequently for the unique O2+• product ions.
Table 3 ignores the so-called secondary chemistry of acetone and propanal—only the primary ions formed because of reagent ion reaction with the analytes are shown. These primary product ions can react with matrix species—most commonly water, though carbon dioxide is common for negative ions—to form secondary product ions [6,7]. Secondary reactions are commonly association reactions (Table 1), requiring that a third molecule (a so-called “third body”) is present in the reactive collision because it carries away sufficient kinetic energy for adduct formation to occur. The association mechanism is hence sensitive to the temperature and pressure in the SIFT-MS instrument flow tube [6,7]. The secondary reactions pertinent to these analytes occur for the H3O+ product (the most common situation):
( CH 3 ) 2 CO . H + + H 2 O   M   ( CH 3 ) 2 CO . H + . H 2 O
CH 3 CH 2 CHO . H + + H 2 O   M   CH 3 CH 2 CHO . H + . H 2 O
In these equations, M represents the necessary third body and the secondary product ions are located +18 units higher than the primary product ion (i.e., at m/z 77). When they occur, it is necessary to measure the signals of secondary product ions for correct quantitation (see the next subsection). Method development should also ensure that they are not interfering with other analytes nor interfered with by other analytes.

2.4. Quantitation in SIFT-MS

The preceding subsections have described how SIFT-MS can analyze a wide range of compounds and generally do so with high specificity. It remains, then, to outline how quantitative results are obtained in SIFT-MS. Since the details of the gas-phase kinetic theory have been described elsewhere [6,37], and the detailed calculation of concentration presented [38], only a brief sketch will be given here. A future article will describe the calculation, as currently adapted in commercial SIFT-MS instruments, in more detail. In this article, the use of non-SI pressure units (torr) follows that of the SIFT-MS literature [38].
The master equation used in commercial instruments to determine the concentration of analyte A in air ([A]i in units of parts-per-billion by volume, ppbV) for each primary product ion (i) of A utilized in the analytical method is:
[ A ] i = 1.035 × 10 10 T F T P F T   ( φ c a r r φ s a m p + 1 )   P i + .   T F P r o d ( i ) + a l l   k i P k i +   T F P r o d ( k i ) t r   b r i   a l l   j k j   R j +   T F R e a g ( j )  
Here:
  • TFT is the temperature in the flow tube in Kelvin (K)
  • PFT is the pressure in the flow tube in torr
  • kj is the rate coefficient for reaction of reagent ion Rj+ with the analyte (in cm3 molecule−1 s−1)
  • tr is the reaction time (in seconds)
  • φcarr is the carrier gas flow in torr L s−1
  • φsamp is the sample flow in torr L s−1
  • Pi+ is the primary product ion signal (in counts per second, cps) for primary product ion i counted by the particle multiplier detector
  • TFProd(i) is the transmission factor for the primary product ion Pi+ (dimensionless)
  • Pki+ is the secondary product ion signal (in cps) for secondary product ion k derived from primary product ion i
  • TFProd(ki) is the transmission factor for the secondary product ion Pki+ (dimensionless)
  • Rj+ is the reagent ion signal (in cps) for the injected reagent ion (j = 0) and its water cluster ions (if appropriate; j = 1, 2, 3)
  • TFReag(j) is the transmission factor for the reagent ion Rj+ (dimensionless)
  • bri is the branching ratio for primary product ion i (0 < i ≤ 1 for calculation purposes, but ordinarily tabulated as a percentage; see, e.g., Table 3).
The coefficient (1.035 × 10−10) in Equation (8) is derived from several physical constants and unit conversions and has units of Torr cm3 molecule−1 K−1. Note that the concentration [A]i is directly proportional to the ratio of measured product ion signal to measured reagent ion signal, which overcomes any ion source drift, should that occur, and confers measurement stability compared to MS detection coupled with chromatographic techniques. Clearly for correct calculation of [A]i the appropriate secondary ions need to be included in the analytical method (e.g., Equations (6) and (7)), otherwise the concentration will underreport.
The parameters in Equation (8) are obtained from various sources:
  • Instrument operating parameters: TFT, PFT, φcarr, and φsamp
  • The instrument’s automated performance check on a certified gas standard: tr, TFProd(i), TFProd(ki), and TFReag(j)
  • Software library: kj and bri (clearly, together with the m/z at which the relevant reagent and product ions will be located)
  • Measurement of sample: Pi+, Pki+, and Rj+.
When more than one concentration determination is made for analyte A (i.e., several primary product ions are used in the method), then instrument software supports automated interference rejection through comparison of the [A]i values (within a user-chosen tolerance; by default, 20%) [39]. The most common failure mode for this approach is when secondary chemistry is omitted from the analytical method.
Several brief comments must be made to conclude the brief overview of SIFT-MS. First, SIFT-MS is quantifying compounds based on unique reagent ion-product ion pairs, not on pattern recognition from a sensor array—the analyst knows which compound is being detected using SIFT-MS, in contrast to sensors [32]. This confidence is evident in the extensive SIFT-MS literature reviewed elsewhere [7,9,10]. Second, several publications have compared SIFT-MS with gold-standard chromatography-based techniques and have reported favorable results, for example, Refs. [40,41,42]. Third, SIFT-MS analysis can be validated according to industry standard procedures (e.g., [43,44]), with only minor adaptations required despite the significant change in analytical approach from chromatographic separation to DIMS. Finally, the CI approach underpinning the SIFT-MS technique provides a wide linear range [6,45]. Ensuring that the instrument remains in the linear range is important for reliable concentration determination [6], although calibration and related approaches can circumvent this restriction when used with care [5]. Generally, however, SIFT-MS measurements do not drift with changes in matrix as chemosensors may. Hence SIFT-MS might provide an alternative to chromatographic methods for benchmarking chemosensors since it provides high-quality analysis in real-time.

3. Quantification of the Volatiles You Smell

Chemosensors—and, particularly, arrays of sensors—have been applied widely in odor assessment, as comprehensively reviewed by Guthrie [8]. Machine olfaction, however, does not currently provide an adequate representation of human olfactory response [8]. Braithwaite [46], however, has posited that SIFT-MS in combination with gas chromatography-olfactometry and the odor profile method may provide an effective complete odor solution—at least for environmental monitoring. Although correlation of SIFT-MS measurements with human sensory measurements is currently in its infancy, the technique has been valuably applied to quantitation of chemical odorants from shortly after it was introduced [6]. Hence the scope of this section is wider than sensory-related studies, encompassing chemical analysis of odorous compounds even when odor was not the main concern. It demonstrates the breadth of odorant analysis achievable with SIFT-MS. Where material has been reviewed in depth previously, that article is cited, and its conclusions summarized. Table 4 provides a convenient summary of the SIFT-MS applications below and lists the complementary chemosensor usage for each.

3.1. Breath Analysis

Abnormal odors in exhaled breath have long been associated with potential ill health [47]. This prompted Linus Pauling and co-workers to pioneer analysis of breath VOCs using gas chromatography in the early 1970s [48]. The SIFT method was applied to breath analysis in one of the earliest articles describing its realization as an analytical technique [49,50], and since that time has found wide application in breath research as described in several reviews [6,9,10,51,52]. Detailed protocols for SIFT-MS analysis of breath are given in various articles [53,54,55,56].
Cross-over between direct-injection mass spectroscopy techniques and SIFT-MS is perhaps greatest for breath malodor, where SIFT-MS is used as the gold-standard technique for validation of lower-cost sensor technologies that are not only more affordable but can also be used as a point-of-care tools for rapid and early-stage diagnosis of disease in clinical practice [57]. SIFT-MS is used to understand the microbial generation of volatile organic compounds and volatile sulfur compounds resulting from microbial metabolism of oral biofilms [58] in parallel with sensory analysis for the development of improved oral care formulations that combat oral malodor [59]. Simple, direct, real-time, and quantitative analysis of malodorous ammonia, hydrogen sulfide, methyl mercaptan, amines, volatile fatty acids, and indoles on breath using SIFT-MS are important contributors to its success [60,61]. It has enabled evidence-based claims to be produced for different malodor products and to provide credibility and validity for all sensor technologies tested to date accelerating the route to market in the field of healthcare [57].

3.2. Food: Flavor and Fragrance

A comprehensive review in 2019 described food flavor applications of SIFT-MS amongst food safety and packaging applications [33], with newer literature summarized in recent articles [10,62,63]. The most relevant area for comparison with chemosensor arrays is human sensory analysis. A substantial subsection of the earlier review was devoted to these applications across a wide range of food products [33]. Although no studies to date have made direct comparison of SIFT-MS and chemosensors, the SIFT-MS studies have consistently demonstrated high correlation with human sensory panels. Recently, automated-SIFT-MS analysis has been adopted by industry for high-throughput screening of commercial dairy products, reducing the workload of a sensory panel [64].

3.3. Materials and Packaging

SIFT-MS is being applied increasingly to the analysis of materials, including packaging (see Section 4.3 and Section 5), but largely remains unexplored from a SIFT-MS perspective, despite the potential indicated by food flavor research [33]. The sole exception is provided by a recent comprehensive study of recycled paperboard packaging, in which automated headspace-SIFT-MS analysis and human sensory analysis were compared [65]. The study concluded that SIFT-MS has potential for sensory classification of paper samples, both for odor intensity ratings and identification of sensory odor notes. Hence SIFT-MS may be able to address the expensive human sensory testing (due to the use of multiple highly trained panelists and low daily sample throughput) by predicting odor intensity rating and odor note.

3.4. Environment

Odors in the environmental context can arise from natural and industrial sources—and in the case of natural sources, these can be amplified through anthropogenic interventions, such as intense farming operations. Odors are usually complex and although chemosensors have been applied with some success [46], they do not qualify as equivalents to the human nose [66] and understanding odor composition at a chemical level is not possible [32]. This is where SIFT-MS can complement sensor data, since it is able to uniquely detect and quantify most odorants, including volatile sulfur compounds [15,16], ammonia and the amines [11,12], aldehydes and ketones [14,67,68], and volatile fatty acids [69,70]. Most compounds in these classes can be quantified by SIFT-MS to below human odor detection thresholds [71]. However, it should be noted that geosmin and 2-methylisoborneol, the well-known “earthy” compounds in drinking water, currently present a challenge for SIFT-MS in complex matrices such as wastewater headspace [72].
SIFT-MS has been applied to various environmental odor applications by van Langenhove and associates at Ghent University. Their early work profiled odors from facilities that intensively farmed laying hens, broiler chickens, and pigs [73,74,75]. The focus of their SIFT-MS research then turned to optimization of odor mitigation technologies, first establishing the scope [76], before determining fundamental parameters [77,78,79]. They then demonstrated that real-time analysis using SIFT-MS is very valuable to engineering of more advanced biofiltration technologies [80,81,82,83,84,85].
Odors emitted from intensively farmed poultry in southeast Queensland, Australia have also been probed—both on-site and in flux chambers—using real-time SIFT-MS [86]. The University of New England, Australia, also utilized SIFT-MS to better understand—and potentially mitigate—chicken litter odors through modified diet [87] and in vitro studies to better understand the microbial origins of increased litter odor during Clostridium perfringens infection [88].
Researchers in South Korea demonstrated the feasibility of odorant monitoring within a broader subset of VOCs in their ambient monitoring study (Section 4.2) [89].
In New Zealand, two industrial-scale odor studies have been performed. The first investigated odor at a gelatin factory, profiling odor sources (i.e., identifying specific odorants) and then assessing the effectiveness of ultraviolet odor neutralization controls on several odor sources [90]. In the second, a substantial study of odor sources was conducted at a large wastewater treatment plant (WWTP) [72]. Identification and quantitation of a wide variety of odorants (aldehydes, amines, organosulfur compounds, volatile fatty acids, etc.) was demonstrated and the results of SIFT-MS analysis were evaluated versus gold-standard olfactometry by a trained sensory panel. From this study, SIFT-MS may have potential as an improved tool for instrument-based sensory analysis compared to chemosensors, though further research and wider application is required. Of particular significance is the ability of SIFT-MS to easily distinguish hydrogen sulfide and methyl mercaptan which have very different odor characters and relative abundances as shown by the lack of correlation between these species in Figure 3.

4. Protection from the Toxic Volatiles You Do Not Smell

A significant proportion of VOCs are toxic. For example, the United States Environmental Protection Agency’s (US EPA) lists 197 hazardous air pollutants (HAPs), of which approximately half are VOCs [91]. In Chapter 467, the United States Pharmacopeia [92], following the International Council for Harmonisation [93], lists five residual solvents that are very toxic or hazardous to the environment (Class 1) and 29 solvents for which there are exposure concerns (Class 2). For such scenarios, if the matrix is consistent and only a handful of volatiles are present, relatively low-cost chemosensors may be appropriate for screening purposes. However, where the matrix is more complex and/or variable and quantitation of multiple components to low concentrations is required, then SIFT-MS likely provides a better fit. SIFT-MS applications of this type are reviewed in this section and summarized in Table 5 (together with the complementary chemosensor applications).

4.1. Occupational Safety and Health, and Indoor Air Quality

VOCs are important contributors to workplace exposure and indoor air quality (IAQ). Because people can be exposed to VOCs for long periods each day in these environments—and hence exposure can be chronic—these areas are treated together in this review.
Smith, Španěl, and coworkers pioneered quantitative analysis of toxic volatiles, and validation of the methodology, only a short time after the selected ion flow tube was adapted to gas analysis [94,95,96]. They focused on the potential of the emerging technique as a tool for evaluating chemical exposure. McEwan’s research group, with clinical co-workers explored several additional applications, including breath alcohol [97] and exposure to common industrial solvents [98,99].
Detection of fumigants and toxic industrial chemicals (TICs) in shipping containers [10] is a significantly more challenging application of SIFT-MS—in terms of both the gaseous chemicals that must be detected and the intended instrument users. The chemical diversity is illustrated in Figure 4 based on their polarity and ionization energy properties. This means that multiple GC analyses are required to report the full suite of volatiles, which is impractical for frontline, non-technical staff. Prior to adoption of SIFT-MS, colorimetric tubes have been utilized by some organization—with very high consumable costs and low user confidence—while handheld photoionization detectors (PIDs), metal oxide-based sensors, and portable infrared spectrometers had unacceptable false positive rates due to poor specificity. Since its adoption in 2006, SIFT-MS instrumentation has proved very effective in this application [100,101,102,103], addressing both the need for broad-spectrum analysis and ease-of-use in an industrial environment. By using both positive and negative ionization, while having a wide range of ionization mechanisms available (not just election transfer; Table 1) even hydrogen cyanide, with its high polarity and ionization energy, is readily detected [104]. Samples are analyzed for all compounds within 2 min. by operational staff.
Other applications of SIFT-MS in workplace exposure assessment include speciating and quantifying airborne solvents in the collision repair industry [105] and identification and quantification of leakage from closed system drug-transfer devices [106]. SIFT-MS has also been used to monitor of a new dosing system for preparation of reference gases at the German Institute for Occupational Safety and Health (IFA) [42]. This study involved full validation versus GC and high-performance LC (HPLC) for routine VOCs and aldehydes, respectively.
Indoor air quality analysis is an expanding area for SIFT-MS, especially through adoption by the semiconductor industry, where it protects both employees and product [10]. SIFT-MS provides broad-spectrum detection of compounds of greatest concern, which cover VOCs [7], acidic gases (HF, HCl, SO2 [25]), and ammonia [12]. With most compounds requiring sub-ppbV detection limits, this application is beyond the capabilities of most chemosensors, but within those of SIFT-MS [45,107].
SIFT-MS is emerging as a powerful tool for the quantification of volatile emissions from diverse materials into residential indoor air. Emissions from personal care products have been reported for aerosol-based hairsprays (in which benzene was detected) [108], non-aerosol products [109], and facial moisturizers [110]. Cleaning [111,112] and cooking [113,114,115] emissions are a very recent focus of research. SIFT-MS has also been applied to analysis of hazardous semi-volatile organophosphate esters used as flame retardants [116], and—by using negatively charged reagent ions—ozone [117]. Exposure to volatile emissions in automobile interiors has also been demonstrated [118], with simultaneous analysis of VOCs and aldehydes, rather than separate GC and HPLC analyses, respectively.

4.2. Environment

This subsection highlights the contributions that SIFT-MS is increasingly making to environmental protection, covering air quality monitoring, water analysis, and an emerging application: microplastics.

4.2.1. Air Quality Monitoring

In 2010, Prince et al. [45] demonstrated high-sensitivity continuous ambient monitoring for the first time, targeting benzene, 1,3-butadiene, ethanol, ethene, and toluene. This work demonstrated low part-per-trillion by volume (pptV) detection limits for the first time. Side-by-side comparison with gold-standard GC/MS [40] using a subset of the US EPA TO-15 method [119], which is based on canister sampling, gave excellent comparability for benzene, toluene, ethylbenzene, and the xylene isomers (BTEX) across diverse field samples. Field deployment of an instrument operating on the newly released nitrogen carrier gas was trialed in mid-2015 [120]. Combined, these studies provided a platform for wider adoption in field monitoring. In Korea, Son et al. monitored 60 volatiles with a cycle time of 3.2 min [89], achieving good linearity over their calibration range (0.174 to 100 ppbV), good repeatability (10% relative standard deviation, %RSD), and sub-ppbV LODs across a wide range of chemical functionalities. This work was followed by a mobile monitoring campaign [121].
The Wolfson Atmospheric Chemistry Laboratory (WACL) at the University of York (UK) has extended the application of SIFT-MS into atmospheric research. In 2016 and 2017 they participated in a campaign in Beijing, China [122]. The campaign comprised two six-week deployment periods (winter and summer) and included automated calibration on a three-day cycle [123]. Air was sampled at a height of 102 m and delivered to the instrument located in a hut at ground level. Figure 5 shows comparative concentration data for SIFT-MS and dual-channel gas chromatography–flame ionization detector (DC-GC-FID) over a 10-day period in the winter campaign [123]. Agreement is excellent for at least one reagent-product ion pair in each case (despite the GC sampling from only 8 m), illustrating the benefit of rapidly switchable reagent ions in SIFT-MS for providing specific analysis and interference rejection in real time. Effective, real-time analysis of nitrous acid (HONO) was also demonstrated during the campaign [124].
More recent campaigns have demonstrated the benefit of SIFT-MS analysis for (i) both organic and inorganic compounds in the field [125,126], (ii) accurate, high-sensitivity analysis of oxygenated VOCs [127], and (iii) monitoring in a moving laboratory [128] with sub-ppbV detection limits.

4.2.2. Water

Automated SIFT-MS may have benefits for water analysis as a screening tool because of its high-sensitivity direct analysis that is robustness to water content. Laboratory analysis of a range of VOCs in water has demonstrated ongoing improvements in quantitation limits [36,43,129]. The latest of these [36] has achieved a 100 ppt limit of quantitation for benzene in water, exceeding European Union and United States maximum contaminant levels by 10- and 50-fold, respectively. Lee et al. [130] demonstrated that SIFT-MS has potential for online analysis of industrial wastewater, but their prototype automated system did not achieve the detection limits required for drinking water.

4.2.3. Soil

SIFT-MS has significant potential for soil analysis, but relatively little research has been conducted to date. Langford et al. [40] analyzed soil gas samples as part of their evaluation of SIFT-MS compared to GC/MS and found that BTEX compared very well. In complex soil-gas samples (e.g., those sampled near a gas station), low-molecular-weight chlorinated hydrocarbons were prone to interference from saturated hydrocarbon residues. Perkins [131] has demonstrated that methanolic extraction [132] can be utilized with SIFT-MS, although only the NO+ reagent ion can be utilized. Good results were obtained for BTEX, with a sample throughput threefold higher than GC/MS.

4.2.4. Microplastics

Researchers at the University of Pisa, Italy, have demonstrated that SIFT-MS is well-suited to analysis of volatile residues in environmental microplastics—pollutants of increasing concern [133,134]. Rapid SIFT-MS analysis is presented as a valuable part of a protocol for comprehensive microplastics identification, since it can rapidly quantify toxic volatile degradation products [134].

4.3. Pharmaceuticals, Personal Care Products, and Packaging

Potential for providing rapid screening of pharmaceutical and personal care products, and their packaging, for toxic chemicals was only realized after automation was seamlessly integrated with SIFT-MS [135]. Automation enables the power of direct sample analysis to be applied repeatably in routine analysis [5], even for compounds that are chromatographically challenging and require special sample prep or instrument configurations [136]. This subsection highlights this emerging application.

4.3.1. Pharmaceutical Products (Including Active Pharmaceutical Ingredients (APIs) and Excipients)

Although few pharmaceutical products are volatile, VOCs can be used as reagents, as solvents in synthesis and/or purification, and may form as byproducts in synthesis or through degradation during storage. In the pharmaceutical industry, GC-based techniques are primarily utilized for these analyses.
The procedure detailed in chapter 467 of the United States Pharmacopeia (USP), “Residual Solvents” [92], is perhaps the most widely employed procedure for the analysis of residual solvents. As promulgated, it utilizes headspace-GC (with flame ionization detection, FID, or mass spectrometric detection), but alternative procedures can also be validated [137]. Recently, a feasibility study was conducted across Class 1 and 2 solvents for water-soluble drug products using headspace-SIFT-MS, with full validation of the Class 2A and 2B solvents according to USP 1467 [44]. The validation study met USP 1467′s acceptance criteria for all solvents except pyridine and hexane (recoveries failed in selected drug products). Work is ongoing because of the throughput benefit (up to 16-fold increase compared to GC), and potential for online monitoring.
SIFT-MS can also be applied to target volatile impurities. Since volatile, carcinogenic nitrosamines were discovered in sartan-class drug products in 2018, widespread nitrosamine testing has been implemented [138,139]. They have since been found in other drug products, but all sources of nitrosamines are yet to be identified [138,139,140]. Discussions about the extent of drug product testing required, and the high-sensitivity analytical methods to do so, are ongoing [140]. Nitrosamines are readily detected using SIFT-MS [141] and recent work has demonstrated a 2-ng g−1 quantitation limit for N-nitrosodimethylamine (NDMA) in 500 mg of pharmaceutical product [142,143], well below the acceptable intake of 96 ng day−1 [138,139].
Ethylene oxide—another toxic impurity in drug products—is also readily analyzed using SIFT-MS. It is a widely used feedstock in the chemical industry [144], including in the manufacture of polyethylene glycols (PEGs) used as surfactants and emulsifiers perfumes, cosmetics, and pharmaceuticals. The United States Pharmacopeia (USP) monograph Polysorbate 80 [145] describes the conventional gas chromatography-flame ionization detection (GC-FID) method for ethylene oxide analysis. This approach is slow, both from the perspective of sample preparation [146] and sample analysis. The rate-limiting step in sample preparation is the 6-hr purification of the PEG matrix for matching to the sample. Furthermore, the GC-FID run time is 38 min. per sample (Figure 6). In contrast, the higher sensitivity of SIFT-MS removes the requirement for matrix matching through 40-fold higher dilution of the sample in water, enabling aqueous ethylene oxide standards to be utilized [147]. Through direct, chromatography-free analysis, SIFT-MS also provides much faster time to result and higher sample throughputs (Figure 6).
Formaldehyde is another compound of concern because it is a known human carcinogen [148,149]. It can arise in polyethylene glycol (PEG) ester surfactants, which are widely used as excipients [150], through autoxidation and thermal degradation. Formaldehyde is, however, challenging to analyze using conventional chromatographic techniques due to its high polarity and reactivity, so typically involves derivatization (and, for high-sensitivity analysis, sample preconcentration). SIFT-MS simplifies formaldehyde analysis via direct analysis from air and headspace at toxicologically relevant concentrations [14,151]. A recent multiple headspace extraction (MHE) study has described quantitation of formaldehyde at low μg g−1 concentration in Gelucire 44/14, a PEG ester used as an excipient in drug products [152].

4.3.2. Personal Care Products

Benzene has recently been detected as a contaminant in diverse commercial personal care products (PCPs), including sunscreens, body sprays and dry shampoos [153]. Chronic exposure to benzene is linked to a significantly increased risk of certain cancers, including leukemia. As noted above, benzene has been quantified to sub-ppb levels in water using headspace-SIFT-MS [36]. Recently, by using the method of standard additions, benzene concentrations in the ppb range have been measured in various suntan lotion matrices using headspace-SIFT-MS analysis [154]. A recent citizen’s petition to the United States Food and Drug Administration (FDA) utilized—in part—SIFT-MS real-time analysis of benzene from dry shampoo products as they are sprayed [108].
Mentioned previously in the context of IAQ, the work of Yeoman and co-workers [109,110] demonstrates that personal exposure to volatiles is significant during use of cosmetic products.

4.3.3. Packaging

Volatile compounds in packaging materials, adhesives, inks, etc., can transfer to pharmaceutical and consumer products—even passing through other packaging barriers—and present a health risk. Automated SIFT-MS has been successfully applied to detection of volatiles in polymeric packaging (e.g., polystyrene [155], high-density polyethylene (HDPE) [156], polyoxymethylene (POM) [157]) and paper/cardboard [65]. A preliminary investigation of detection of mineral oil saturated hydrocarbons and mineral oil aromatic hydrocarbons (so-called “MOSH and MOAH” [158]) on a variety of polymer and paper products holds promise for a rapid screening test using SIFT-MS [159]. Particularly significant, given the recent publicity on personal care product contamination [108,153], was the detection of benzene in most products (Figure 7).
To date the most thorough SIFT-MS study focused on packaging safety targeted formaldehyde in drug delivery devices [160]. The method was validated successfully and approved in a regulatory submission. Direct injection of whole-air samples into the SIFT-MS instrument greatly simplified and accelerated method development and product testing compared to the conventional approach involving derivatization with 2,4-dinitrophenylhydrazine (DNPH), solvent elution, and analysis using HPLC with ultraviolet detection (HPLC-UV).

5. Untargeted SIFT-MS Analysis: Parallels with Sensor Arrays

SIFT-MS instruments can be utilized in a manner analogous to the so-called electronic noses, or enoses (sensor arrays), by acquiring sample data in full-scan mode followed by post-processing using multivariate statistical analysis. The basic parallels are illustrated schematically in Figure 8. This approach has emerged in recent years, with applications including rapid origin authentication of Moroccan Argan [161] and Mediterranean olive [162] oils, detection of trace adulterants in Moroccan Argan oil [163], strawberry flavor mixes [164], classification of parmesan cheese [62], rapid discrimination of Vitis vinifera berries [165,166], determination of human diseased states via breath [167,168], and virgin and recycled HDPE [156]. The main advantage of this approach is that it is of broader scope than targeted analysis, detecting a wider range of compounds, including potentially important species that may lie outside the scope of a targeted analysis, including contaminants.
Although the above studies utilize the SIFT-MS technique essentially as a “black box” there is always the ability to retrospectively interrogate the full-scan data to identify the compounds whose spectral features are responsible for classifying different types of samples (Figure 8). This approach has been utilized recently by Langford and coworkers for strawberry flavors [164] and HDPE [156], and by Geffroy and co-workers for wine grapes [165,166]. This capability is lacking for sensor arrays, as noted by Reineccius (Section 2.3 above) [32], but is possible where chemosensors are coupled to gas chromatographs. With SIFT-MS, however, it can be used to create optimized targeted methods through identification of the significant discriminatory variables.
It must be noted, however, that for untargeted SIFT-MS analysis extra care is required in method development and routine analysis, since the normalization of ion signals and robustness to humidity, are no longer handled automatically—in contrast with concentration calculation in targeted analysis [6,38]. Hence it is of particular importance that: (1) reagent ions and their 18O isotopologues should not be included, (2) samples should be well within the linear range (for matrix volatiles as well as target analytes) [6], and (3) humidity should be kept constant. Failure to do so can result in the statistical analysis generating spurious results through assigning importance to variables that are merely indicators of humidity changes and/or volatiles present at high concentrations undergoing secondary reactions [6,7]. These factors will be addressed in detail in a forthcoming article.

6. Conclusions

The SIFT-MS technique has found broad applications in the detection of odorous and hazardous volatile compounds during the nearly three decades since it was introduced. Its applications can generally be characterized as those that require specific analysis of multiple compounds with diverse chemical functionalities at high sensitivity in real-time (or with high sample throughput). For industrial applications, robustness and ease of use are also important attributes, enabling laboratory-grade analysis to be conducted by non-technical staff in less-controlled environments. Fumigant analysis furnishes a good example of an industrial application ideally matched to SIFT-MS. In terms of technical applications, SIFT-MS odor analysis shows promise for effective instrument-based sensory analysis because of the broad-spectrum detection of chemically diverse odorants (including hydrogen sulfide, ammonia, and low-molecular-weight aldehydes and volatile fatty acids) in a single run. In the pharmaceutical industry, the ability to rapidly switch between very different analytical methods—due to chromatography-free analysis—is highly valued (e.g., from nitrosamine analysis to ethylene oxide to residual solvents).
SIFT-MS better complements chemosensor technologies than do more conventional off-line, laboratory-based techniques such as GC and LC because it is real-time and field-portable, while results compare favorably with the “gold standard” GC and LC techniques. This means that SIFT-MS may be better suited to benchmarking chemosensor performance than GC and LC—especially for organizations manufacturing, standardizing, or utilizing sensors for diverse volatile compounds.
The broad applications of the SIFT-MS technique enable it to complement low-cost chemosensors. The very soft chemical ionization coupled with mass spectrometry enables one SIFT-MS instrument to be applied to a wide variety of matrices, providing specific, sensitive, and quantitative analysis of diverse volatile compounds. Because chemosensors usually exhibit higher susceptibility to interference they typically do not perform well in matrices that are complex or show significant variability. SIFT-MS applied in parallel may support faster development and tuning of chemosensors (for example, for oral malodor diagnosis). Chemosensors and SIFT-MS instruments have important and complementary roles to play in gas-phase volatiles analysis.

Funding

This research received no external funding.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Not applicable.

Acknowledgments

The author is grateful to many colleagues and collaborators (past and present, local and global) with whom he has had the privilege of exploring many new applications of the SIFT-MS technique.

Conflicts of Interest

The author is an employee of Syft Technologies Limited, a manufacturer of commercial SIFT-MS instruments.

References

  1. Dettmer-Wilde, K.; Engewald, W. Practical Gas Chromatography; Springer: Heidelberg, Germany, 2014; 902p. [Google Scholar]
  2. Fanali, S.; Haddad, P.R.; Poole, C.F.; Riekkola, M.-L. Liquid Chromatography: Fundamentals and Instrumentation, 2nd ed.; Elsevier: Oxford, UK, 2017; 784p. [Google Scholar]
  3. Taylor, A.J.; Beauchamp, J.D.; Langford, V.S. Non-destructive and high-throughput–APCI-MS, PTR-MS and SIFT-MS as methods of choice for exploring flavor release. In Dynamic Flavor: Capturing Aroma Release using Real-Time Mass Spectrometry; Beauchamp, J.D., Ed.; American Chemical Society: Washington, DC, USA, 2021; pp. 1–16. [Google Scholar] [CrossRef]
  4. McEwan, M.J. Direct analysis mass spectrometry. In Ion Molecule Attachment Reactions: Mass Spectrometry; Fujii, T., Ed.; Springer: New York, NY, USA, 2015; pp. 263–317. [Google Scholar]
  5. Perkins, M.J.; Langford, V.S. Application of routine analysis procedures to a direct mass spectrometry technique: Selected ion flow tube mass spectrometry (SIFT-MS). Rev. Sep. Sci. 2021, 3, e21003. [Google Scholar] [CrossRef]
  6. Smith, D.; Španěl, P. Selected ion flow tube mass spectrometry (SIFT-MS) for on-line trace gas analysis. Mass Spectrom. Rev. 2005, 24, 661–700. [Google Scholar] [CrossRef] [PubMed]
  7. Smith, D.; McEwan, M.J.; Španěl, P. Understanding gas phase ion chemistry is the key to reliable selected ion flow tube-mass spectrometry analyses. Anal. Chem. 2020, 92, 12750–12762. [Google Scholar] [CrossRef] [PubMed]
  8. Guthrie, B. Machine olfaction. In Springer Handbook of Odor; Buettner, A., Ed.; Springer: Cham, Switzerland, 2017; pp. 459–504. [Google Scholar]
  9. Španěl, P.; Smith, D. Progress in SIFT-MS: Breath analysis and other applications. Mass Spectrom. Rev. 2011, 30, 236–267. [Google Scholar] [CrossRef] [PubMed]
  10. Smith, D.; Španěl, P.; Demarais, N.; Langford, V.S.; McEwan, M.J. Recent developments and applications of selected ion flow tube mass spectrometry, SIFT-MS. Mass Spec. Rev. 2023, e21835. [Google Scholar] [CrossRef]
  11. Španěl, P.; Smith, D. Selected ion flow tube studies of the reactions of H3O+, NO+, and O2+ with several amines and some other nitrogen-containing molecules. Int. J. Mass Spectrom. 1998, 176, 203–211. [Google Scholar] [CrossRef]
  12. Smith, D.; Bloor, R.; George, C.; Pysanenko, A.; Španěl, P. Release of toxic ammonia and volatile organic compounds by heated cannabis and their relation to tetrahydrocannabinol content. Anal. Methods 2015, 7, 4104–4110. [Google Scholar] [CrossRef]
  13. Španěl, P.; Smith, D. SIFT studies of the reactions of H3O+, NO+ and O2+ with several aromatic and aliphatic hydrocarbons. Int. J. Mass Spectrom. 1998, 181, 1–10. [Google Scholar] [CrossRef]
  14. Španěl, P.; Ji, Y.; Smith, D. SIFT studies of the reactions of H3O+, NO+ and O2+ with a series of aldehydes and ketones. Int. J. Mass Spectrom. Ion Process. 1997, 165, 25–37. [Google Scholar] [CrossRef]
  15. Williams, T.L.; Adams, N.G.; Babcock, L.M. Selected ion flow tube studies of H3O+(H2O)0.1 reactions with sulfides and thiols. Int. J. Mass Spectrom. 1998, 172, 149–159. [Google Scholar] [CrossRef]
  16. Španěl, P.; Smith, D. Selected ion flow tube studies of the reactions of H3O+, NO+, and O2+ with some organosulphur molecules. Int. J. Mass Spectrom. 1998, 176, 167–176. [Google Scholar] [CrossRef]
  17. LabSyft Software: Compound Library; Syft Technologies Limited: Christchurch, New Zealand, 2017.
  18. Španěl, P.; Smith, D. SIFT studies of the reactions of H3O+, NO+ and O2+ with a series of alcohols. Int. J. Mass Spectrom. Ion Process. 1997, 167, 375–388. [Google Scholar] [CrossRef]
  19. Guo, Y.; Grabowski, J.J. Mechanistic insight into the reactions of O with aromatic compounds and the synthesis of didehydroaromatic anions in the gas phase. Int. J. Mass Spectrom. Ion Process. 1992, 117, 299–326. [Google Scholar] [CrossRef]
  20. Španěl, P.; Smith, D. Selected ion flow tube studies of the reactions of H3O+, NO+, and O2+ with several aromatic and aliphatic monosubstituted halocarbons. Int. J. Mass Spectrom. 1999, 189, 213–223. [Google Scholar] [CrossRef]
  21. Thomas, R.; Liu, Y.; Mayhew, C.A.; Peverall, R. Selected ion flow tube studies of the gas phase reactions of O, O2 and OH with a variety of brominated compounds. Int. J. Mass Spectrom. Ion Process. 1996, 155, 163–183. [Google Scholar] [CrossRef]
  22. Fehsenfeld, F.C.; Ferguson, E.E. Laboratory studies of negative ion reactions with atmospheric trace constituents. J. Chem. Phys. 1974, 61, 3181–3193. [Google Scholar] [CrossRef]
  23. Fahey, D.W.; Böhringer, H.; Fehsenfeld, F.C.; Ferguson, E.E. Reaction rate constants for O2(H2O)n ions n = 0–4 with O3, NO, SO2 and CO2. J. Chem. Phys. 1982, 76, 1799–1805. [Google Scholar] [CrossRef]
  24. Smith, D.; Pysanenko, A.; Španěl, P. The quantification of carbon dioxide in humid air and exhaled breath by selected ion flow tube mass spectrometry. Rapid Commun. Mass Spectrom. 2009, 23, 1419–1425. [Google Scholar] [CrossRef]
  25. Hera, D.; Langford, V.S.; McEwan, M.J.; McKellar, T.I.; Milligan, D.B. Negative reagent ions for real time detection using SIFT-MS. Environments 2017, 4, 16. [Google Scholar] [CrossRef]
  26. Böhringer, H.; Fahey, D.W.; Fehsenfeld, F.C.; Ferguson, E.E. Temperature dependence of the three-body association of Cl, NO2, and NO3 with SO2. J. Chem. Phys. 1984, 81, 2696–2698. [Google Scholar] [CrossRef]
  27. United States Environmental Protection Agency. Understanding Global Warming Potentials. Available online: https://www.epa.gov/ghgemissions/understanding-global-warming-potentials (accessed on 17 January 2023).
  28. Wilson, P.F.; Freeman, C.G.; McEwan, M.J. Reactions of small hydrocarbons with H3O+, O2+ and NO+ ions. Int. J. Mass Spectrom. 2003, 229, 143–149. [Google Scholar] [CrossRef]
  29. Dryahina, K.; Smith, D.; Španěl, P. Quantification of methane in humid air and exhaled breath using selected ion flow tube mass spectrometry. Rapid Commun. Mass Spectrom. 2010, 24, 1296–1304. [Google Scholar] [CrossRef]
  30. Ikezoe, Y.; Matsuoka, S.; Takebe, M.; Viggiano, A.A. Gas Phase Reaction Rate Constants through 1986; Maruzen: Tokyo, Japan, 1987. [Google Scholar]
  31. Kao, L.W.; Kristine, A.; Nañagas, K.A. Toxicity associated with carbon monoxide. Clin. Lab. Med. 2006, 26, 99–125. [Google Scholar] [CrossRef] [PubMed]
  32. Reineccius, G. Flavor Chemistry and Technology; CRC Press: Boca Raton, FL, USA, 2006; p. 63. [Google Scholar]
  33. Langford, V.S.; Padayachee, D.; McEwan, M.J.; Barringer, S.A. Comprehensive odorant analysis for on-line applications using selected ion flow tube mass spectrometry (SIFT-MS). Flavour Fragr. J. 2019, 34, 393–410. [Google Scholar] [CrossRef]
  34. Lehnert, A.-S.; Perreca, E.; Gershenzon, J.; Pohnert, G.; Trumbore, S.E. Simultaneous real-time measurement of isoprene and 2-methyl-3-buten-2-ol emissions from trees using SIFT-MS. Front. Plant Sci. 2020, 11, 578204. [Google Scholar] [CrossRef]
  35. Allpress, C.; Crittenden, D.; Ma, J.; McEwan, M.; Robinson, S.; Wilson, P.; Wu, M. Real-time differentiation of ethylbenzene and the xylenes using selected ion flow tube mass spectrometry. Rapid Commun. Mass Spectrom. 2019, 33, 1844–1849. [Google Scholar] [CrossRef]
  36. Perkins, M.J.; Langford, V.S. Application of headspace-SIFT-MS to direct analysis of hazardous volatiles in drinking water. Environments 2022, 9, 124. [Google Scholar] [CrossRef]
  37. Španěl, P.; Smith, D. Influence of water vapour on SIFT-MS analyses of trace gases in humid air and breath. Rapid Commun. Mass Spectrom. 2000, 14, 1898–1906. [Google Scholar] [CrossRef]
  38. Španěl, P.; Dryahina, K.; Smith, D. A general method for the calculation of absolute trace gas concentrations in air and breath from selected ion flow tube mass spectrometry data. Int. J. Mass Spectrom. 2006, 249, 230–239. [Google Scholar] [CrossRef]
  39. LabSyft Software, Version 1.0 and Above; Syft Technologies Limited: Christchurch, New Zealand. Available online: www.syft.com (accessed on 15 January 2023).
  40. Langford, V.S.; Graves, I.; McEwan, M.J. Rapid monitoring of volatile organic compounds: A comparison between gas chromatography/mass spectrometry and selected ion flow tube mass spectrometry. Rapid Commun. Mass Spectrom. 2014, 28, 10–18. [Google Scholar] [CrossRef]
  41. Hastie, C.; Thompson, A.; Perkins, M.J.; Langford, V.S.; Eddleston, M.; Homer, N. Selected ion flow tube-mass spectrometry (SIFT-MS) as an alternative to gas chromatography/mass spectrometry (GC/MS) for the analysis of cyclohexanone and cyclohexanol in plasma. ACS Omega 2021, 6, 32818–32822. [Google Scholar] [CrossRef] [PubMed]
  42. Kaus, C. Method Development for Continuous Monitoring of Selected VOC Test Gases by GC and SIFT-MS and Its Use for Verifying a New Dosing System for Test Gas Generation. Ph.D. Dissertation, University of Wuppertal, Wuppertal, Germany, 2021. [Google Scholar] [CrossRef]
  43. Perkins, M.J.; Langford, V.S. Standard validation protocol for selected ion flow tube mass spectrometry methods applied to direct headspace analysis of aqueous volatile organic compounds. Anal. Chem. 2021, 93, 8386–8392. [Google Scholar] [CrossRef] [PubMed]
  44. Biba, E.; Perkins, M.J.; Langford, V.S. Stimuli to the Revision Process: High-throughput residual solvent analysis using selected ion flow tube mass spectrometry (SIFT-MS). United States Pharmacopeia. Pharm. Forum 2021, 47, 1. Available online: https://online.usppf.com/usppf/document/GUID-2EE1BF6B-C82B-4F11-8E0B-C5520A4E8C3D_10101_en-US (accessed on 15 January 2023).
  45. Prince, B.J.; Milligan, D.B.; McEwan, M.J. Application of selected ion flow tube mass spectrometry to real-time atmospheric monitoring. Rapid Commun. Mass Spectrom. 2010, 24, 1763–1769. [Google Scholar] [CrossRef]
  46. Braithwaite, S.K. Sensory Analysis and Health Risk Assessment of Environmental Odors. Ph.D. Thesis, University of California-Los Angeles, Los Angeles, CA, USA, 2019. Available online: https://escholarship.org/uc/item/21g660d3 (accessed on 10 January 2023).
  47. Manolis, A. The diagnostic potential of breath analysis. Clin. Chem. 2003, 29, 5–15. [Google Scholar] [CrossRef]
  48. Teranish, R.; Robinson, A.B.; Cary, P.; Mon, T.R.; Pauling, L. Gas-chromatography of volatiles from breath and urine. Anal. Chem. 1972, 44, 18–20. [Google Scholar] [CrossRef]
  49. Španěl, P.; Smith, D. Selected ion flow tube: A technique for quantitative trace gas analysis of air and breath. Med. Biol. Eng. Comput. 1996, 34, 409–419. [Google Scholar] [CrossRef] [PubMed]
  50. Smith, D.; Španěl, P. Application of ion chemistry and the SIFT technique to the quantitative analysis of trace gases in air and on breath. Int. Rev. Phys. Chem. 1996, 15, 231–271. [Google Scholar] [CrossRef]
  51. Španěl, P.; Smith, D. On the features, successes and challenges of selected ion flow tube mass spectrometry. Eur. J. Mass Spectrom. 2013, 19, 225–246. [Google Scholar] [CrossRef]
  52. Smith, D.; Španěl, P. Status of selected ion flow tube MS: Accomplishments and challenges in breath analysis and other areas. Bioanalysis 2016, 8, 1183–1201. [Google Scholar] [CrossRef]
  53. Španěl, P.; Dryahina, K.; Smith, D. A quantitative study of the influence of inhaled compounds on their concentrations in exhaled breath. J. Breath Res. 2013, 7, 017106. [Google Scholar] [CrossRef] [PubMed]
  54. Storer, M.; Dummer, J.; Sturney, S.; Epton, M. Validating SIFT-MS analysis of volatiles in breath. Curr. Anal. Chem. 2013, 9, 576–583. [Google Scholar] [CrossRef]
  55. Smith, D.; Španěl, P.; Herbig, J.; Beauchamp, J. Mass spectrometry for real-time quantitative breath analysis. J. Breath Res. 2014, 8, 027101. [Google Scholar] [CrossRef]
  56. Belluomo, I.; Boshier, P.R.; Myridakis, A.; Vadhwana, B.; Markar, S.R.; Španěl, P.; Hanna, G.B. Selected ion flow tube mass spectrometry for targeted analysis of volatile organic compounds in human breath. Nat. Protoc. 2021, 16, 3419–3438. [Google Scholar] [CrossRef]
  57. Saad, S.; (University of the West of England, Bristol, UK). Personal communication, 2022.
  58. Saad, S.; Hewett, K.; Greenman, J. Effect of mouth-rinse formulations on oral malodour processes in tongue-derived perfusion biofilm model. J. Breath Res. 2012, 6, 016001. [Google Scholar] [CrossRef]
  59. Saad, S.; Fitzgerald, M.; Hewett, K.; Greenman, J.; Vandeven, M.; Trivedi, H.M.; Masters, J.G. Short- and long-term effects of a dentifrice containing dual zinc plus arginine on intra-oral halitosis: Improvements in breath quality. J. Clin. Dent. 2018, 29, A46–A54. [Google Scholar]
  60. Smith, D.; Wang, T.; Pysanenko, A.; Španěl, P. A selected ion flow tube mass spectrometry study of ammonia in mouth- and nose-exhaled breath and in the oral cavity. Rapid Commun. Mass Spectrom. 2008, 22, 783–789. [Google Scholar] [CrossRef]
  61. Pysanenko, A.; Španěl, P.; Smith, D. A study of sulfur-containing compounds in mouth- and nose-exhaled breath and in the oral cavity using selected ion flow tube mass spectrometry. J. Breath Res. 2008, 2, 046004. [Google Scholar] [CrossRef]
  62. Padayachee, D.; Langford, V.S. SIFTing through flavor—Exploring real-time, real-life processes using SIFT-MS. In Dynamic Flavor: Capturing Aroma Release using Real-Time Mass Spectrometry; Beauchamp, J.D., Ed.; American Chemical Society: Washington, DC, USA, 2021; pp. 51–65. [Google Scholar] [CrossRef]
  63. Barringer, S. From mold worms to fake honey: Using SIFT-MS to improve food quality. In Dynamic Flavor: Capturing Aroma Release using Real-Time Mass Spectrometry; Beauchamp, J.D., Ed.; American Chemical Society: Washington, DC, USA, 2021; pp. 99–105. [Google Scholar] [CrossRef]
  64. Wood, J.E.; Gill, B.D.; Longstaff, W.M.; Crawford, R.A.; Indyk, H.E.; Kissling, R.C.; Lin, Y.-H.; Bergonia, C.; Davis, L.M.; Matuszek, A. Dairy product quality using screening of aroma compounds by selected ion flow tube-mass spectrometry: A chemometric approach. Int. Dairy J. 2021, 121, 105107. [Google Scholar] [CrossRef]
  65. Langford, V.S.; Du Bruyn, C.; Padayachee, D. An evaluation of selected ion flow tube mass spectrometry for rapid instrumental determination of paper type, origin and sensory attributes. Packag. Technol. Sci. 2021, 34, 245–260. [Google Scholar] [CrossRef]
  66. Mannebeck, D. Olfactometers according to EN 13725. In Springer Handbook of Odor; Buettner, A., Ed.; Springer: Cham, Switzerland, 2017; pp. 545–551. [Google Scholar]
  67. Smith, D.; Van Doren, J.M.; Španěl, P. A selected ion flow tube study of the reactions of H3O+, NO+, and O2+ with saturated and unsaturated aldehydes and subsequent hydration of the product ions. Int. J. Mass Spectrom. 2002, 213, 163–176. [Google Scholar] [CrossRef]
  68. Michel, E.; Schoon, N.; Amelynck, C.; Guimbaud, C.; Catoire, V.; Arijs, E. A selected ion flow tube study of the reactions of H3O+, NO+ and O2+ with methyl vinyl ketone and some atmospherically important aldehydes. Int. J. Mass Spectrom. 2005, 244, 50–59. [Google Scholar] [CrossRef]
  69. Španěl, P.; Smith, D. SIFT studies of the reactions of H3O+, NO+ and O2+ with a series of volatile carboxylic acids and esters. Int. J. Mass Spectrom. Ion Process. 1998, 172, 137–147. [Google Scholar] [CrossRef]
  70. Smith, D.; Chippendale, T.W.E.; Španěl, P. Selected ion flow tube, SIFT, studies of the reactions of H3O+, NO+ and O2+ with some biologically active isobaric compounds in preparation for SIFT-MS analyses. Int. J. Mass Spectrom. 2011, 303, 81–89. [Google Scholar] [CrossRef]
  71. van Gemert, L.J. Odour Thresholds: Compilations of Odour Threshold Values in Air, Water and Other Media; Oliemans, Punter and Partners: Utrecht, The Netherlands, 2011. [Google Scholar]
  72. Langford, V.S.; Billiau, C.; McEwan, M.J. Evaluation of the efficacy of SIFT-MS for speciation of wastewater treatment plant odors in parallel with human sensory analysis. Environments 2020, 7, 90. [Google Scholar] [CrossRef]
  73. Van Huffel, K.; Heynderickx, P.M.; Dewulf, J.; Van Langenhove, H. Measurement of odorants in livestock buildings: SIFT-MS and TD-GC-MS. Chem. Eng. 2012, 30, 67–72. [Google Scholar] [CrossRef]
  74. Heynderickx, P.M.; Van Huffel, K.; Dewulf, J.; Van Langenhove, H. SIFT-MS for livestock emission characterization: Application of similarity coefficients. Chem. Eng. Trans. 2012, 30, 157–162. [Google Scholar] [CrossRef]
  75. Heynderickx, P.M.; Van Huffel, K.; Dewulf, J.; Van Langenhove, H. Application of similarity coefficients to SIFT-MS data for livestock emission characterization. Biosyst. Eng. 2013, 114, 44–54. [Google Scholar] [CrossRef]
  76. Volckaert, D.; Heynderickx, D.; Fathi, E.; Van Langenhove, H. SIFT-MS a logical tool for monitoring and evaluating a biofilter performance. In Proceedings of the Biotechniques for Air Pollution Control and Bioenergy (5th Conference), Nimes, France, 10−13 September 2013. [Google Scholar]
  77. Heynderickx, P.M.; Spanel, P.; Van Langenhove, H. Quantification of octanol–water partition coefficients of several aldehydes in a bubble column using selected ion flow tube mass spectrometry. Fluid Phase Equilibria 2014, 367, 22–28. [Google Scholar] [CrossRef]
  78. Bruneel, J.; Walgraeve, C.; Van Huffel, K.; Van Langenhove, H. Determination of the gas-to-liquid partitioning coefficients using a new dynamic absorption method (DynAb Method). Chem. Eng. J. 2016, 283, 544–552. [Google Scholar] [CrossRef]
  79. Walgraeve, C.; Bruneel, J.; Van Huffel, K.; Demeestere, K.; Vincze, L.; De Meulenaer, B.; Van Langenhove, H. Sorption behaviour of targeted volatile organic compounds on airborne particulate matter using selected ion flow tube mass spectrometry. Biosyst. Eng. 2015, 131, 84–94. [Google Scholar] [CrossRef]
  80. Volckaert, D.; Ebude, D.E.L.; Van Langenhove, H. SIFT-MS analysis of the removal of dimethyl sulphide, n-hexane and toluene from waste air by a two phase partitioning bioreactor. Chem. Eng. J. 2016, 290, 346–352. [Google Scholar] [CrossRef]
  81. Volckaert, D.; Heynderickx, P.M.; Fathi, E.; Van Langenhove, H. SIFT-MS a novel tool for monitoring and evaluating a biofilter performance. Chem. Eng. J. 2016, 304, 98–105. [Google Scholar] [CrossRef]
  82. Bruneel, J.; Walgraeve, C.; Dumortier, S.; Stockman, J.; Demeyer, P.; Van Langenhove, H. Increasing mass transfer of volatile organic compounds in air scrubbers: A fundamental study for different gas–liquid systems. J. Chem. Technol. Biotechnol. 2018, 93, 1468–2180. [Google Scholar] [CrossRef]
  83. Bruneel, J.; Walgraeve, C.; Demeyer, P.; Van Langenhove, H. Increasing mass transfer of volatile organic compounds in air scrubbers: Relation between partition coefficient and mass transfer coefficient in a pilot-scale scrubber. J. Chem. Technol. Biotechnol. 2018, 93, 2170–2180. [Google Scholar] [CrossRef]
  84. Bruneel, J.; Walgraeve, C.; Mukurarinda, J.; Boon, N.; Van Langenhove, H. Biofiltration of hexane, acetone and dimethyl sulphide using wood, compost and silicone foam. J. Chem. Technol. Biotechnol. 2018, 93, 2234–2243. [Google Scholar] [CrossRef]
  85. Bruneel, J.; Follert, J.L.H.; Laforce, B.; Vincze, L.; Van Langenhove, H.; Walgraeve, C. Dynamic performance of a fungal biofilter packed with perlite for the abatement of hexane polluted gas streams using SIFT-MS and packing characterization with advanced X-ray spectroscopy. Chemosphere 2020, 253, 126684. [Google Scholar] [CrossRef]
  86. Atzeni, M.; Langford, V.; Prince, B.; Mayer, D. Rapid Continuous Chemical Analysis of Meat Chicken Shed Emissions by SIFT–MS; Publication no. 16/052; Rural Industries Research and Development Corporation (Australian Government): Canberra, Australia, 2016. Available online: https://agrifutures.com.au/wp-content/uploads/publications/16-052.pdf (accessed on 14 January 2023).
  87. Sharma, N.K.; Choct, M.; Dunlop, M.W.; Wu, S.-B.; Castada, H.Z.; Swick, R.A. Characterisation and quantification of changes in odorants from litter headspace of meat chickens fed diets varying in protein levels and additives. Poult. Sci. 2017, 96, 851–860. [Google Scholar] [CrossRef]
  88. Sharma, N.K.; Keerqin, C.; Wu, S.-B.; Choct, M.; Swick, R.A. Emissions of volatile odorous metabolites by Clostridium perfringens-in vitro study using two broth cultures. Poult. Sci. 2017, 96, 3291–3297. [Google Scholar] [CrossRef]
  89. Son, H.D.; An, J.G.; Ha, S.Y.; Kim, G.B.; Yim, U.H. Development of real-time and simultaneous quantification of volatile organic compounds in ambient with SIFT-MS (selected ion flow tube-mass spectrometry). J. Korean Soc. Atmos. Environ. 2018, 34, 393–405. [Google Scholar] [CrossRef]
  90. Langford, V.S.; McEwan, M.J.; Askey, M.; Barnes, H.A.; Olerenshaw, J.G. Comprehensive instrumental odor analysis using SIFT-MS: A case study. Environments 2018, 5, 43. [Google Scholar] [CrossRef]
  91. United States Environmental Protection Agency. Initial List of Hazardous Air Pollutants with Modifications. Available online: https://www.epa.gov/haps/initial-list-hazardous-air-pollutants-modifications (accessed on 10 January 2023).
  92. ⟨467⟩ Residual Solvents; United States Pharmacopeia: Rockville, MD, USA, 2019.
  93. ICH Q3C (R8) International Council for Harmonisation of Technical Requirements for Pharmaceuticals for Human Use (2021). Guideline for Residual Solvents: Q3C (R8) Step 5 Version (and Annexes There to). Available online: https://www.ema.europa.eu/en/ich-q3c-r8-residual-solvents (accessed on 10 January 2023).
  94. Španěl, P.; Rolfe, P.; Rajan, B.; Smith, D. The selected ion flow tube (SIFT)—A novel technique for biological monitoring. Ann. Occup. Hyg. 1996, 40, 615–626. [Google Scholar] [CrossRef] [PubMed]
  95. Španěl, P.; Cocker, J.; Rajan, B.; Smith, D. Validation of the SIFT technique for trace gas analysis of breath using the syringe injection technique. Ann. Occup. Hyg. 1997, 41, 373–382. [Google Scholar] [CrossRef]
  96. Smith, D.; Španěl, P.; Thompson, J.M.; Rajan, B.; Cocker, J.; Rolfe, P. The selected ion flow tube method for workplace analyses of trace gases in air and breath: Its scope, validation, and applications. Appl. Occup. Environ. Hyg. 1998, 13, 817–823. [Google Scholar] [CrossRef]
  97. Wilson, P.F.; Freeman, C.G.; McEwan, M.J.; Milligan, D.B.; Allardyce, R.A.; Shaw, G.M. Alcohol in breath and blood: A selected ion flow tube mass spectrometric study. Rapid Commun. Mass Spectrom. 2001, 15, 413–417. [Google Scholar] [CrossRef]
  98. Wilson, P.F.; McEwan, M.J.; Milligan, D.B.; Allardyce, R.A.; Shaw, G.M. Real time in vivo measurement of VOCs by SIFT-MS. In Proceedings of the American Society for Mass Spectrometry (ASMS) Conference, Chicago, IL, USA, 27–31 May 2001. [Google Scholar]
  99. Wilson, P.F.; Freeman, C.G.; McEwan, M.J.; Allardyce, R.A.; Shaw, G.M. SIFT-MS measurement of VOC distribution coefficients in human blood constituents and urine. Appl. Occup. Environ. Hyg. 2003, 18, 759–763. [Google Scholar] [CrossRef]
  100. Milligan, D.B.; Francis, G.J.; Prince, B.J.; McEwan, M.J. Demonstration of selected ion flow tube MS detection in the parts per trillion range. Anal. Chem. 2007, 79, 2537–2540. [Google Scholar] [CrossRef]
  101. Mück, O.; Stock, A. Occupational safety in import containers containing fumigants, other gases and volatile substances: Practical experiences. In Proceedings of the 9th. Int. Conf. on Controlled Atmosphere and Fumigation in Stored Products, Antalya, Türkiye, 15–19 October 2012; Available online: http://ftic.co.il/2012AntalyaPDF/SESSION%2002%20PAPER%2010.pdf (accessed on 16 January 2023).
  102. Hinz, R.; ‘t Mannetje, A.; Glass, B.; McLean, D.; Pearce, N.; Douwes, J. Exposures to fumigants and residual chemicals in workers handling cargo from shipping containers and export logs in New Zealand. Ann. Work. Expo. Health 2020, 64, 826–837. [Google Scholar] [CrossRef]
  103. Hinz, R.; ‘t Mannetje, A.; Glass, B.; McLean, D.; Douwes, J. Airborne fumigants and residual chemicals in shipping containers arriving in New Zealand. Ann. Work. Expo. Health 2022, 66, 481–494. [Google Scholar] [CrossRef]
  104. Anicich, V.G. Evaluated bimolecular ion-molecule gas phase kinetics of positive ions for use in modeling planetary atmospheres, cometary comae, and interstellar clouds. J. Phys. Chem. Ref. Data 1993, 22, 1469–1569. [Google Scholar] [CrossRef]
  105. Keer, S.; Taptiklis, P.; Glass, B.; McLean, D.; McGlothlin, J.D.; Douwes, J. Determinants of airborne solvent exposure in the collision repair industry. Ann. Work. Expo. Health 2018, 62, 871–883. [Google Scholar] [CrossRef]
  106. Doepke, A.; Streicher, R.P. Source apportionment and quantification of liquid and headspace leaks from closed system drug-transfer devices via selected ion flow tube mass spectrometry (SIFT-MS). PLoS ONE 2021, 16, e0258425. [Google Scholar] [CrossRef] [PubMed]
  107. Lehnert, A.S.; Behrendt, T.; Ruecker, A.; Pohnert, G.; Trumbore, S.E. SIFT-MS optimization for atmospheric trace gas measurements at varying humidity. Atmos. Meas. Tech. 2020, 13, 3507–3520. [Google Scholar] [CrossRef]
  108. Valisure, Inc. Citizen Petition on Benzene in Dry Shampoo Products (31 October 2022). Available online: https://assets-global.website-files.com/6215052733f8bb8fea016220/6360f7f49903987d8f4f4309_Valisure%20FDA%20Citizen%20Petition%20on%20Benzene%20in%20Dry%20Shampoo%20Products_103122.pdf (accessed on 14 January 2023).
  109. Yeoman, A.M.; Shaw, M.; Carslaw, N.; Murrells, T.; Passant, N.; Lewis, A.C. Simplified speciation and atmospheric volatile organic compounds emission rates from non-aerosol personal care products. Indoor Air 2020, 30, 459–472. [Google Scholar] [CrossRef]
  110. Yeoman, A.M.; Heeley-Hill, A.C.; Shaw, M.; Andrews, S.J.; Lewis, A.C. Inhalation of VOCs from facial moisturizers and the influence of dose proximity. Indoor Air 2022, 32, e12948. [Google Scholar] [CrossRef]
  111. Harding-Smith, E.; O’Leary, C.; Shaw, M.; Carslaw, N.; Davies, H.L.; Dillon, T.J.; Mehra, A.; Phillips, G.; Ye, L.; Jones, B. Impacts of cleaning on indoor air quality: Regular versus ‘green’ cleaning products. In Proceedings of the Indoor Air 2022, Kuopio, Finland, 12–16 June 2022; 2022. [Google Scholar]
  112. Harding-Smith, E.; O’Leary, C.; Davies, H.; Shaw, D.; Shaw, M.; Dillon, T.J.; Carslaw, N. Dynamic headspace SIFT-MS coupled with indoor air modelling for bottom-up estimations of indoor air pollution from cleaning. In Proceedings of the AGU Fall Meeting, Chicago, IL, USA, 12–16 December 2022. [Google Scholar]
  113. Winkless, R.; Thompson, M.; O’Leary, C.; Kumar, A.; Davies, H.; Shaw, M.; Carslaw, N.; Dillon, T.J. The kinetics of cooking–emission rates for Harmful VOC. In Proceedings of the 26th International Symposium on Gas Kinetics and Related Phenomena–GK2022, Rennes, France, 28 August–1 September 2022. [Google Scholar]
  114. O’Leary, C.; Davies, H.L.; Mehra, A.; Harding-Smith, E.; Ye, L.; Shaw, M.; Dillon, T.J.; Phillips, G.; Jones, B.; Carslaw, N. Investigating primary emissions and secondary chemistry following cooking. In Proceedings of the Indoor Air 2022, Kuopio, Finland, 12–16 June 2022. [Google Scholar]
  115. O’Leary, C.; Davies, H.L.; Mehra, A.; Harding-Smith, E.; Winkless, R.; Ye, L.; Shaw, M.; Dillon, T.J.; Phillips, G.; Jones, B.; et al. Investigating the impacts of cooking on indoor air quality. In Proceedings of the EnvChem2022: Chemistry of the Whole Environment Research, York, UK, 14–15 July 2022. [Google Scholar]
  116. Ghislain, M.; Reyrolle, M.; Sotiropoulos, J.-M.; Pigot, T.; Plaisance, H.; Le Bechec, M. Study of the chemical ionization of organophosphate esters in air using selected ion flow tube−mass spectrometry for direct analysis. J. Am. Soc. Mass Spectrom. 2022, 33, 865–874. [Google Scholar] [CrossRef]
  117. Pasetto, L.V.; Richard, R.; Pic, J.-S.; Manero, M.-H.; Violleau, F.; Simon, V. Ozone quantification by selected ion flow tube mass spectrometry: Influence of humidity and manufacturing gas of ozone generator. Anal. Chem. 2019, 91, 15518–15524. [Google Scholar] [CrossRef] [PubMed]
  118. Zhu, J.H.; Nones, C.; Li, Y.; Milligan, D.; Prince, B.; Polster, M.; Dearth, M. Ultra-trace real time VOC measurements by SIFT-MS for VIAQ. SAE Int. J. Engines 2017, 10, 1815–1819. [Google Scholar] [CrossRef] [Green Version]
  119. U.S. Environmental Protection Agency. Compendium of Methods for the Determination of Toxic Organic Compounds in Ambient Air: Method TO-15, 2nd ed.; U.S. Environmental Protection Agency: Research Triangle Park, NC, USA, 1997; EPA 600/625/R-96-010b.
  120. Langford, V.S. Real-time monitoring of volatile organic compounds in ambient air using direct injection mass spectrometry. LCGC N. Am. 2022, 40, 174–179. Available online: https://www.chromatographyonline.com/view/real-time-monitoring-of-volatile-organic-compounds-in-ambient-air-using-direct-injection-mass-spectrometry (accessed on 14 January 2023). [CrossRef]
  121. Hwang, K.; An, J.G.; Lee, S.; Choi, W.; Yim, U.H. Study on the ozone formation potential of volatile organic compounds in Busan using SIFT-MS. J. Korean Soc. Atmos. Environ. 2020, 36, 645–668. [Google Scholar] [CrossRef]
  122. Shi, Z.; Vu, T.; Kotthaus, S.; Harrison, R.M.; Grimmond, S.; Yue, S.; Zhu, T.; Lee, J.; Han, Y.; Demuzere, M.; et al. Introduction to the special issue “In-depth study of air pollution sources and processes within Beijing and its surrounding region (APHH-Beijing)”. Atmos. Chem. Phys. 2019, 19, 7519–7546. [Google Scholar] [CrossRef]
  123. Shaw, M.; Acton, J.; Squires, F.; Carpenter, L.; Lewis, A.; Hopkins, J. Evaluation of Direct Mass spectrometry for VOC concentration and emission determination in Beijing. In Proceedings of the Atmospheric Science Conference (UK), York, UK, 3−4 July 2018. [Google Scholar]
  124. Crilley, L.R.; Kramer, L.J.; Ouyang, B.; Duan, J.; Zhang, W.; Tong, S.; Ge, M.; Tang, K.; Qin, M.; Xie, P.; et al. Intercomparison of nitrous acid (HONO) measurement techniques in a megacity (Beijing). Atmos. Meas. Tech. 2019, 12, 6449–6463. [Google Scholar] [CrossRef]
  125. Hien, T.T.; Huy, D.H.; Dominutti, P.A.; Chi, N.D.T.; Hopkins, J.R.; Shaw, M.; Forster, G.; Mills, G.; Le, H.A.; Oram, D. Comprehensive volatile organic compound measurements and their implications for ground-level ozone formation in the two main urban areas of Vietnam. Atmos. Environ. 2022, 269, 118872. [Google Scholar] [CrossRef]
  126. Dominutti, P.A.; Hopkins, J.R.; Shaw, M.; Mills, G.; Le, H.A.; Huy, D.H.; Forster, G.L.; Keita, S.; Hien, T.T.; Oram, D.E. Evaluating major anthropogenic VOC emission sources in densely populated Vietnamese cities. Environ. Pollut. 2023, 318, 120927. [Google Scholar] [CrossRef] [PubMed]
  127. Borrás, E.; Tortajada-Genaro, L.A.; Ródenas, M.; Vera, T.; Speak, T.; Seakins, P.; Shaw, M.D.; Lewis, A.C.; Muñoz, A. On-line solid phase microextraction derivatization for the sensitive determination of multi-oxygenated volatile compounds in air. Atmos. Meas. Tech. 2021, 14, 4989–4999. [Google Scholar] [CrossRef]
  128. Wagner, R.L.; Farren, N.J.; Davison, J.; Young, S.; Hopkins, J.R.; Lewis, A.C.; Carslaw, D.C.; Shaw, M.D. Application of a mobile laboratory using a selected-ion flow-tube mass spectrometer (SIFT-MS) for characterisation of volatile organic compounds and atmospheric trace gases. Atmos. Meas. Tech. 2021, 14, 6083–6100. [Google Scholar] [CrossRef]
  129. Perkins, M.J.; Langford, V.S.; McEwan, M.J. High-throughput analysis of volatile compounds in air, water and soil using SIFT-MS. Curr. Trends Mass Spectrom. 2018, 37, 24–29. Available online: https://www.chromatographyonline.com/view/high-throughput-analysis-volatile-compounds-air-water-and-soil-using-sift-ms (accessed on 14 January 2023).
  130. Lee, S.-H.; Shin, E.-J.; Zoh, K.-D.; Kang, Y.-S.; Choi, J.-W. Direct mass spectrometry with online headspace sample pretreatment for continuous water quality monitoring. Water 2020, 12, 1843. [Google Scholar] [CrossRef]
  131. Perkins, M.J. Methanolic Extraction of Soils by Automated Selected Ion Flow Tube Mass Spectrometry (SIFT-MS). Anatune Application Note. Available online: http://bit.ly/3XeCaSI (accessed on 14 January 2023).
  132. U.S. Environmental Protection Agency. Method 5021A-Volatile Organic Compounds in Various Sample Matrices Using Equilibrium Headspace Analysis, Revision 2. 2014. Available online: https://www.epa.gov/sites/production/files/2015-12/documents/5021a.pdf (accessed on 14 January 2023).
  133. La Nasa, J.; Lomonaco, T.; Manco, E.; Ceccarini, A.; Fuoco, R.; Corti, A.; Modugno, F.; Castelvetro, V.; Degano, I. Plastic breeze: Volatile organic compounds (VOCs) emitted by degrading macro- and microplastics analyzed by selected ion flow-tube mass spectrometry. Chemosphere 2021, 270, 128612. [Google Scholar] [CrossRef]
  134. Castelvetro, V.; Corti, A.; Biale, G.; Ceccarini, A.; Degano, I.; La Nasa, J.; Lomonaco, T.; Manariti, A.; Manco, E.; Modugno, F.; et al. New methodologies for the detection, identification, and quantification of microplastics and their environmental degradation by-products. Environ. Sci. Pollut. Res. 2021, 28, 46764–46780. [Google Scholar] [CrossRef]
  135. Perkins, M.J.; Padayachee, D.; Langford, V.S.; McEwan, M.J. High throughput residual solvent and residual monomer analysis using selected ion flow tube mass spectrometry. Chromatogr Today 2017, 10, 2–5. Available online: https://www.chromatographytoday.com/article/gc-ms/46/syft-technologies-ltd/phigh-throughput-residualnbspsolvent-and-residual-monomer-analysis-using-selected-ionnbspflow-tube-mass-spectrometryp/2242/download (accessed on 3 January 2023).
  136. Langford, V.S.; Perkins, M.J. Demystifying sample preparation for headspace analysis using direct injection mass spectrometry. The Column 2022, 18, 24–28. Available online: https://www.chromatographyonline.com/view/demystifying-sample-preparation-for-headspace-analysis-using-direct-injection-mass-spectrometry (accessed on 14 January 2023).
  137. Residual Solvents—Verification of Compendial Procedures and Validation of Alternative Procedures 〈1467〉; United States Pharmacopeia: Rockville, MD, USA, 2019.
  138. European Medicines Agency. Questions and Answers for Marketing Authorisation Holders/Applicants on the CHMP Opinion for the Article 5(3) of Regulation (EC) No 726/2004 Referral on Nitrosamine Impurities in Human Medicinal Products, EMA/409815/2020. Available online: https://www.ema.europa.eu/en/documents/referral/nitrosamines-emea-h-a53-1490-questions-answers-marketing-authorisation-holders/applicants-chmp-opinion-article-53-regulation-ec-no-726/2004-referral-nitrosamine-impurities-human-medicinal-products_en.pdf (accessed on 15 September 2022).
  139. United States Food and Drug Administration. Guidance for Industry. Control of Nitrosamine Impurities in Human Drugs. 2021. Available online: https://www.fda.gov/media/141720/download (accessed on 15 September 2022).
  140. Office of New Drugs, United States Food and Drug Administration. Nitrosamines as Impurities in Drugs−Health Risk Assessment and Mitigation Public Workshop, 29–30 March 2021. Available online: https://www.fda.gov/media/150932/download (accessed on 21 March 2022).
  141. Langford, V.S.; Gray, J.D.C.; Maclagan, R.G.A.R.; Milligan, D.B.; McEwan, M.J. Real-time measurements of nitrosamines in air. Int. J. Mass Spectrom. 2015, 377, 490–495. [Google Scholar] [CrossRef]
  142. Perkins, M.J.; Langford, V.S. Simple, Rapid Analysis of N-nitrosodimethylamine (NDMA) Impurity in Ranitidine Products Using SIFT-MS. Syft Technologies Application Note. 2022. Available online: http://bit.ly/3GB8wzW (accessed on 12 January 2023).
  143. Perkins, M.J.; Langford, V.S. Simple, Rapid Analysis of NDMA in a Recalled Valsartan Product Using SIFT-MS. Syft Technologies Application Note. 2022. Available online: https://bit.ly/3QFclbR (accessed on 12 January 2023).
  144. Wikipedia. Ethylene Oxide. Available online: https://en.wikipedia.org/wiki/Ethylene_oxide (accessed on 20 September 2022).
  145. United States Pharmacopeia. Polysorbate 80. Available online: https://www.usp.org/sites/default/files/usp/document/harmonization/excipients/polysorbate_80.pdf (accessed on 20 September 2022).
  146. ⟨228⟩ Ethylene Oxide and Dioxane; United States Pharmacopeia: Rockville, MD, USA, 2023.
  147. Silva, L.; Connor, A.; Bastian, C.; Williams, C.; Morseth, Z.; Franco, E. Simple, Rapid Analysis of Ethylene Oxide in a Polysorbate 80 Excipient Using SIFT-MS. Syft Technologies Application Note. 2022. Available online: https://bit.ly/3QLWRmC (accessed on 16 January 2023).
  148. National Toxicology Program, U.S. Department of Health and Human Services. 12th Report on Carcinogens; 10 June 2011. Available online: https://ntrl.ntis.gov/NTRL/dashboard/searchResults/titleDetail/PB2011111646.xhtml (accessed on 14 January 2023).
  149. European Commission. Proposal for a Directive of the European Parliament and of the Council Amending Directive 2004/37/EC on the Protection of Workers from the Risks Related to Exposure to Carcinogens or Mutagens at Work, COM/2018/0171 final-2018/081 (COD). Available online: https://eur-lex.europa.eu/legal-content/EN/TXT/?uri=CELEX:52018PC0171 (accessed on 14 January 2023).
  150. Panigrahi, K.C.; Patra, C.N.; Jena, G.K.; Ghose, D.; Jena, J.; Panda, S.K.; Sahu, M. Gelucire: A versatile polymer for modified release drug delivery system. Future J. Pharm. Sci. 2018, 4, 102–108. [Google Scholar] [CrossRef]
  151. Langford, V.S.; Perkins, M.J.; Milligan, D.B.; Prince, B.J.; McEwan, M.J. High-speed formaldehyde analysis for the process-line and laboratory: SIFT-MS. In Proceedings of the International Society of Automation (ISA) 62nd Analysis Division Symposium, Pasadena, CA, USA, 23−27 April 2017; Available online: http://bit.ly/3INcYyh (accessed on 16 January 2023).
  152. Perkins, M.J.; Langford, V.S. Simple, Rapid Analysis of Formaldehyde Impurities in Gelucire Excipient Using SIFT-MS. Syft Technologies Application Note. 2022. Available online: http://bit.ly/3Xca3nd (accessed on 16 January 2023).
  153. Bettenhausen, C.A. Finding benzene everywhere we look. Chem. Eng. News 2022, 100, 24–26. [Google Scholar] [CrossRef]
  154. Silva, L.; Langford, V. High-Throughput, Quantitative Analysis of Benzene in Personal Care Products Using Headspace-SIFT-MS. Syft Technologies Application Note. 2022. Available online: http://bit.ly/3Xv3Cew (accessed on 16 January 2023).
  155. Perkins, M.J.; Langford, V.S. Multiple headspace extraction-selected ion flow tube mass spectrometry (MHE-SIFT-MS). Part 1: A protocol for method development and transfer to routine analysis. Rev. Sep. Sci. 2022, 4, e22001. [Google Scholar] [CrossRef]
  156. Langford, V.S.; Perkins, M.J. Untargeted SIFT-MS headspace analysis: High-throughput differentiation of virgin and recycled polyethylene pellets. Rapid Commun. Mass Spectrom. 2022, 36, e9230. [Google Scholar] [CrossRef]
  157. Bell, K.J.M.; Ma, J.; Padayachee, D.; Langford, V.S.; Perkins, M.J. Time-Resolved Thermal Extraction of Residual Formaldehyde Using Thermal Desorption-SIFT-MS. Syft Technologies Application Note. 2019. Available online: http://bit.ly/3knBTOQ (accessed on 16 January 2023).
  158. European Food Safety Authority Panel on Contaminants in the Food Chain (CONTAM). Scientific Opinion on Mineral Oil Hydrocarbons in Food. EFSA J. 2012, 10, 2704. [Google Scholar] [CrossRef]
  159. Perkins, M.J.; Langford, V.S. Simple, Rapid Analysis of Packaging for Volatile MOAH and MOSH Contaminants. Syft Technologies Application Note. 2022. Available online: http://bit.ly/3iEIGD1 (accessed on 16 January 2023).
  160. Wicks, J.; Perkins, M.J.; Langford, V.S. Validation of a High-Throughput SIFT-MS Method for Analysis of Formaldehyde Leachate. Syft Technologies Application Note. 2019. Available online: http://bit.ly/3khxZ9Z (accessed on 16 January 2023).
  161. Kharbach, M.; Kamal, R.; Mansouri, A.M.; Marmouzi, I.; Viaene, J.; Cherrah, Y.; Alaoui, K.; Vercammen, J.; Bouklouze, A.; Vander Heyden, Y. Selected ion flow tube mass spectrometry (SIFT-MS) fingerprinting versus chemical profiling for geographic traceability of Moroccan Argan oils. Food Chem. 2018, 263, 8–17. [Google Scholar] [CrossRef]
  162. Bajoub, A.; Medina-Rodriguez, S.; Ajal, E.A.; Cuadros-Rodríguez, L.; Monasterio, R.P.; Vercammen, J.; Fernández-Gutiérreza, A.; Carrasco-Pancorbo, A. A metabolic fingerprinting approach based on selected ion flow tube mass spectrometry (SIFT-MS) and chemometrics: A reliable tool for Mediterranean origin-labelled olive oils authentication. Food Res. Int. 2018, 106, 233–242. [Google Scholar] [CrossRef]
  163. Kharbach, M.; Yu, H.W.; Kamal, R.; Marmouzi, I.; Alaoui, K.; Vercammen, J.; Bouklouze, A.; Vander Heyden, Y. Authentication of extra virgin Argan oil by selected-ion flow-tube mass-spectrometry fingerprinting and chemometrics. Food Chem. 2022, 383, 132565. [Google Scholar] [CrossRef] [PubMed]
  164. Langford, V.S.; Bell, K.J.M. Rapid determination of strawberry flavour integrity using static headspace-selected ion flow tube mass spectrometry. Chromatogr. Today 2019, 12, 2–6. Available online: https://www.chromatographytoday.com/article/gc-ms/46/syft-technologies-ltd/rapid-determination-of-strawberry-flavour-integrity-using-static-headspace-selected-ion-flow-tube-mass-spectrometry/2558/download (accessed on 3 January 2023).
  165. Baerenzung dit Baron, T.; Yobrégat, O.; Jacques, A.; Simon, V.; Geffroy, O. A novel approach to discriminate the volatilome of Vitis vinifera berries by [SIFT-MS] analysis and chemometrics. Food Res. Int. 2022, 157, 111434. [Google Scholar] [CrossRef] [PubMed]
  166. Geffroy, O.; Baerenzung dit Baron, T.; Yobrégat, O.; Denat, M.; Simon, V.; Jacques, A. The SIFT-MS fingerprint of Vitis vinifera L. cv. Syrah berries is stable over the second part of maturation under warm conditions of climate. OENO One 2022, 56, 139–146. [Google Scholar] [CrossRef]
  167. Stefanuto, P.H.; Zanella, D.; Vercammen, J.; Henket, M.; Schleich, F.; Louis, R.; Focant, J.F. Multimodal combination of GC x GC-HRTOFMS and SIFT-MS for asthma phenotyping using exhaled breath. Sci. Rep. 2020, 10, 16159. [Google Scholar] [CrossRef]
  168. Segers, K.; Slosse, A.; Viaene, J.; Bannier, M.A.G.E.; Van de Kant, K.D.G.; Dompeling, E.; Van Eeckhaut, A.; Vercammen, J.; Heyden, Y.V. Feasibility study on exhaled-breath analysis by untargeted selected-ion flow-tube mass spectrometry in children with cystic fibrosis, asthma, and healthy controls: Comparison of data pretreatment and classification techniques. Talanta 2021, 225, 122080. [Google Scholar] [CrossRef]
Figure 1. A schematic illustration of the SIFT-MS technique, showing the three zones in this continuous-analysis approach. Used with permission of Syft Technologies.
Figure 1. A schematic illustration of the SIFT-MS technique, showing the three zones in this continuous-analysis approach. Used with permission of Syft Technologies.
Chemosensors 11 00111 g001
Figure 2. Specificity in SIFT-MS analysis is achieved through a unique combination of attributes.
Figure 2. Specificity in SIFT-MS analysis is achieved through a unique combination of attributes.
Chemosensors 11 00111 g002
Figure 3. The relationship between the dominant sulfur odorants (hydrogen sulfide and methyl mercaptan) observed at a WWTP using SIFT-MS for the primary (green circles), secondary water (red crosses), and secondary solids (blue diamonds) odor sources. Note the use of odor activity values (OAVs) and logarithmic axes. From Ref. [72]; used with permission of the authors.
Figure 3. The relationship between the dominant sulfur odorants (hydrogen sulfide and methyl mercaptan) observed at a WWTP using SIFT-MS for the primary (green circles), secondary water (red crosses), and secondary solids (blue diamonds) odor sources. Note the use of odor activity values (OAVs) and logarithmic axes. From Ref. [72]; used with permission of the authors.
Chemosensors 11 00111 g003
Figure 4. Illustration of the diversity of gaseous fumigants and several TICs based on their polarity (as indicated by the dipole moment) and ionization energy (in electron volts, eV). The ionization energies of NO and O2, corresponding to the NO+ and O2+• reagent ions of SIFT-MS, are shown for reference.
Figure 4. Illustration of the diversity of gaseous fumigants and several TICs based on their polarity (as indicated by the dipole moment) and ionization energy (in electron volts, eV). The ionization energies of NO and O2, corresponding to the NO+ and O2+• reagent ions of SIFT-MS, are shown for reference.
Chemosensors 11 00111 g004
Figure 5. Mixing ratios (concentrations) for ambient air in Beijing during winter 2016, as measured using DC-GC-FID and SIFT-MS [123]. Legend text color matches relevant data series. Unpublished conference poster illustration reproduced with permission of the authors.
Figure 5. Mixing ratios (concentrations) for ambient air in Beijing during winter 2016, as measured using DC-GC-FID and SIFT-MS [123]. Legend text color matches relevant data series. Unpublished conference poster illustration reproduced with permission of the authors.
Chemosensors 11 00111 g005
Figure 6. Headspace SIFT-MS enables analysis of ethylene oxide in Polysorbate 80 samples to be conducted at significantly higher throughput, while eliminating very slow sample preparation, due to enhanced sensitivity. Used with permission of Syft Technologies.
Figure 6. Headspace SIFT-MS enables analysis of ethylene oxide in Polysorbate 80 samples to be conducted at significantly higher throughput, while eliminating very slow sample preparation, due to enhanced sensitivity. Used with permission of Syft Technologies.
Chemosensors 11 00111 g006
Figure 7. Benzene headspace concentrations (in ppbV) determined using automated SIFT-MS for various packaging materials. See Ref. [160] for full details. Used with permission of Syft Technologies.
Figure 7. Benzene headspace concentrations (in ppbV) determined using automated SIFT-MS for various packaging materials. See Ref. [160] for full details. Used with permission of Syft Technologies.
Chemosensors 11 00111 g007
Figure 8. A simplified schematic representation of the parallels between chemosensors (enoses) and untargeted SIFT-MS sample classification approaches. SIFT-MS has an advantage in that compounds cause sample discrimination can often be identified.
Figure 8. A simplified schematic representation of the parallels between chemosensors (enoses) and untargeted SIFT-MS sample classification approaches. SIFT-MS has an advantage in that compounds cause sample discrimination can often be identified.
Chemosensors 11 00111 g008
Table 1. The most common ion–molecule reaction mechanisms of the SIFT-MS reagent ions and the resulting product ion m/z, either relative to the molecular weight or a standard product ion.
Table 1. The most common ion–molecule reaction mechanisms of the SIFT-MS reagent ions and the resulting product ion m/z, either relative to the molecular weight or a standard product ion.
Mechanism NameReagent Ion(s)General EquationShift Relative to Parent Ion
Proton transfer (PT)H3O+AH+ + B → BH+ + A+1 (simple PT)
−17 (loss of OH moiety)
Electron transfer (ET)NO+, O2+•, O2−•A+/− + B → B+/− + A0
Dissociative ETNO+, O2+•A+ + B → C+ + ProductsVaries
Hydride abstractionNO+, O2+•A+ + B-H → B+ + HA−1
AssociationNO+, O−•, O2−•, OHA+/− + B + M → {B.A}+/− + M+Reagent m/z
(e.g., +30 for NO+)
Proton abstractionO−•, O2−•, OH, NO2, NO3A + BH → B + AH−1
Mechanism NameReagent Ion(s)General EquationStandard Product * m/z
Hydrogen atom transferO−•A + BH → AH + BReagent ion m/z − 1 **
DisplacementO−•, OHA + RB → RA + B
(R = alkyl)
−35 and −37 for B = Cl
−79 and −81 for B = Br
EliminationO−•, OHA + RB → R’ + R”A + B
(R’ = alkene; R” = alkyl)
−35 and −37 for B = Cl
−79 and −81 for B = Br
Associative detachmentO−•, O2−•, OHA + B → AB + e
(e = electron)
No product detected (e)
* The reactions in the lower section of the table yield a standard product ion. For displacement and elimination mechanisms, Cl and Br product ions have been most commonly encountered to date, so they are used for illustrative purposes. ** The reagent ion m/z is defined as a negative value here. Hence for O−• (m/z = −16) the product ion following hydrogen atom transfer is OH (m/z = −17).
Table 2. Examples of compounds and the SIFT-MS reagent ions by which they are practically analyzed (i.e., sufficient sensitivity and selectivity).
Table 2. Examples of compounds and the SIFT-MS reagent ions by which they are practically analyzed (i.e., sufficient sensitivity and selectivity).
CompoundReference(s)H3O+NO+O2+•O−•O2−•OHNO2NO3
Ammonia[11,12]
Benzene[13]
Formaldehyde[14]
Hydrogen sulfide[15,16,17]
Ethanol[18]
Pyridine[11,19]
Methyl bromide[20,21]
Carbon dioxide[22,23,24,25]
Hydrogen chloride[25]
Sulfur dioxide[23,25,26]
Table 3. Primary product ions formed from reaction of acetone and propanal with H3O+, NO+, and O2+• [14]. The NO+ reagent ion is most frequently utilized to achieve selective analysis.
Table 3. Primary product ions formed from reaction of acetone and propanal with H3O+, NO+, and O2+• [14]. The NO+ reagent ion is most frequently utilized to achieve selective analysis.
Reagent IonAcetone Propanal
Formulam/zbr (%)Formulam/zbr (%)
H3O+(CH3)2CO.H+59100CH3CH2CHO.H+59100
NO+(CH3)2CO.NO+88100CH3CH2CO+57100
O2+•(CH3)2CO+•
CH3CO+•
58
43
60
40
CH3CH2CHO+•
CH3CH2CO+
58
57
50
50
Table 4. The complementary nature of SIFT-MS and chemosensor detection of odorous compounds in the application areas discussed in Section 3.
Table 4. The complementary nature of SIFT-MS and chemosensor detection of odorous compounds in the application areas discussed in Section 3.
ApplicationSIFT-MSChemosensorsComment(s)
Breath analysisBreath research
Validation of sensor technologies
Medical laboratory testing (centralized)
Routine clinic use (distributed)Breath research is usually conducted in parallel with GC/MS (biomarker discovery)
Chemosensor instrumentation is likely more affordable for clinics
Food, flavor, and fragranceHigh-throughput sensory screening
Product R&D/new product formulation
High-throughput sensory screeningBoth technologies use pattern recognition for sensory analysis; SIFT-MS often adds ability to identify odorants
Materials and packagingHigh-throughput sensory screeningHigh-throughput sensory screeningSee comment above
Environment (outdoor)Odor emissions and mitigation research
Speciated odorant monitoring
Sensory analysis
Mobile laboratory
Distributed monitoring systems
Sensory analysis
Highly portable odor assessment
Table 5. The complementary nature of SIFT-MS and chemosensor detection of hazardous volatiles in the application areas discussed in Section 4.
Table 5. The complementary nature of SIFT-MS and chemosensor detection of hazardous volatiles in the application areas discussed in Section 4.
ApplicationSIFT-MSChemosensorsComment(s)
Occupational safety and healthOn-site (fixed or mobile) or remote laboratory analysis of diverse toxic compounds
Research
Devices for personal protection (limited capability per device)Complex matrices, such as shipping container air, may reduce the usefulness of sensors due to false alarms
Indoor air qualityProtection of high-value, highly susceptible products (semiconductors)
Research
Distributed monitoring systems
Highly portable analysis
High sensitivity and selectivity are essential for semiconductor manufacturing, but this is a very specialized use-case
Environment (outdoor)Speciated 24/7 continuous monitoring of volatiles
Mobile monitoring (including while driving)
Validation of sensor technologies
Distributed monitoring systems
Sensory analysis
Highly portable odor assessment
Specificity of SIFT-MS provides validation of chemosensor performance
PharmaceuticalsHigh-throughput quantitative product safety analysis (e.g., nitrosamines, ethylene oxide, residual solvents)High-throughput product quality screening based on pattern recognitionDue to lower cost, chemosensors with appropriate performance may be a better fit for process monitoring
Personal care productsHigh-throughput quantitative product safety analysis (e.g., benzene, ethylene oxide)
Product R&D/new product formulation
High-throughput product quality screening based on pattern recognitionSee comment for
Pharmaceuticals
Materials and packagingHigh-throughput quantitative product safety analysis (e.g., benzene, styrene, formaldehyde)High-throughput product quality screening based on pattern recognitionSee comment for
Pharmaceuticals
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Langford, V.S. SIFT-MS: Quantifying the Volatiles You Smell…and the Toxics You Don’t. Chemosensors 2023, 11, 111. https://doi.org/10.3390/chemosensors11020111

AMA Style

Langford VS. SIFT-MS: Quantifying the Volatiles You Smell…and the Toxics You Don’t. Chemosensors. 2023; 11(2):111. https://doi.org/10.3390/chemosensors11020111

Chicago/Turabian Style

Langford, Vaughan S. 2023. "SIFT-MS: Quantifying the Volatiles You Smell…and the Toxics You Don’t" Chemosensors 11, no. 2: 111. https://doi.org/10.3390/chemosensors11020111

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop