Next Article in Journal
Volatile Compounds in Monovarietal Wines of Two Amarone Della Valpolicella Terroirs: Chemical and Sensory Impact of Grape Variety and Origin, Yeast Strain and Spontaneous Fermentation
Previous Article in Journal
Bootstrap Resampling of Temporal Dominance of Sensations Curves to Compute Uncertainties
Previous Article in Special Issue
Development of Optimal Digesting Conditions for Microplastic Analysis in Dried Seaweed Gracilaria fisheri
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Comprehensive Strategy for Sample Preparation for the Analysis of Food Contaminants and Residues by GC–MS/MS: A Review of Recent Research Trends

1
State Key Laboratory of Supramolecular Structure and Materials, College of Chemistry, Jilin University, Changchun 130012, China
2
College of Food Science and Engineering, Jilin University, Changchun 130062, China
3
College of Plant Protection, Jilin Agricultural University, Changchun 130118, China
4
Jilin Institute for Food Control, Changchun 130103, China
*
Authors to whom correspondence should be addressed.
Foods 2021, 10(10), 2473; https://doi.org/10.3390/foods10102473
Submission received: 25 May 2021 / Revised: 6 October 2021 / Accepted: 8 October 2021 / Published: 15 October 2021

Abstract

:
Food safety and quality have been gaining increasing attention in recent years. Gas chromatography coupled to tandem mass spectrometry (GC–MS/MS), a highly sensitive technique, is gradually being preferred to GC–MS in food safety laboratories since it provides a greater degree of separation on contaminants. In the analysis of food contaminants, sample preparation steps are crucial. The extraction of multiple target analytes simultaneously has become a new trend. Thus, multi-residue analytical methods, such as QuEChERs and adsorption extraction, are fast, simple, cheap, effective, robust, and safe. The number of microorganic contaminants has been increasing worldwide in recent years and are considered contaminants of emerging concern. High separation in MS/MS might be, in certain cases, favored to sample preparation selectivity. The ideal sample extraction procedure and purification method should take into account the contaminants of interest. Moreover, these methods should cooperate with high-resolution MS, and other sensitive full scan MSs that can produce a more comprehensive detection of contaminants in foods. In this review, we discuss the most recent trends in preparation methods for highly effective detection and analysis of food contaminants, which can be considered tools in the control of food quality and safety.

1. Introduction

Since the 1970s, gas chromatography (GC) coupled to mass spectrometry (MS) has been applied to the detection of most contaminants and residues routinely found in foods. However, in the last ten years, the majority of food control laboratories moved from GC–MS to GC–MS/MS as the preferred analytical technique to address GC amenable compounds, mainly due to the interference in single-step GC–MS analysis caused by coeluting matrix compounds. Certain compounds cannot be separated in single-step MS because they coincide with selected ions in GC–MS. In order to analyze multiple compounds by GC–MS, it is necessary to ensure that all target compounds can be adequately separated qualitatively and quantitatively.
A sample preparatory step in food contaminants detection can be included to enable high recovery and good reproducibility, which should ideally be rapid, inexpensive, simple, easy to automatize, and environmentally friendly. High separation in MS/MS may substitute the importance of preparation selectivity; however, the choice of the method for the sample preparatory step is also determined by the type of food matrix as well as by the contaminants of interest. Thus, multi-residue preparation methods, such as QuEChERs and adsorption extraction, are a most important technology enabling the simultaneous extraction of as many targets as possible [1]. When a large number of contaminants is to be detected in samples, MS/MS must operate in multi-reaction monitoring (MRM) mode, and the use of multi-residue analysis methods in analysis providing high sensitivity and high specificity has become the new trend.
Among emerging chemical hazards in foods, contaminants that may constitute a threat to human health and the environment in the future can be included. These hazards might include not only contaminants of emerging concern, such as brominated flame retardants (BFRs), perfluorochemicals (PFOS), endocrine disrupters (ERs), but also those of biological origin. Thus, the ideal sample extraction procedure and purification method should take into account the contaminants of interest. Moreover, these methods should cooperate with high-resolution MS, and other sensitive full scan MSs that can produce a more comprehensive detection of contaminants in foods.
GC–MS/MS has become a major technique for the analysis of contaminants and residues in foods due to their high sensitivity and selectivity, being widely used for the analysis of low-polarity, volatile, and thermally stable compounds. Considering the nature of contaminants detected in foods, we summarized herein and discussed two commonly used processes: (1) analysis of volatile organic compounds (VOCs) by headspace (HS) injection with/without derivatization; (2) and analysis of semi-volatile organic compounds (SVOCs) or thermally stable compounds after extraction and clean-up. In general, foods, such as grains, vegetables, fruits, sugars, beverages, edible fungi, flavorings, medicinal plants, and foods of animal origins, are often complex matrices. Moreover, foods can be classified according to their form into solid and liquid food matrices, and effective analytical strategies must take into account the type of food matrix.
The aim of this review article was to discuss preparation methods for the analysis of contaminants and residue in foods by GC–MS/MS with an emphasis on literature published in recent years. Promising future trends and perspectives are also discussed.

2. Preparation Methods for the Analysis of VOCs

VOCs in food contaminants mainly include phthalate esters (PAEs), polycyclic aromatic hydrocarbons (PAHs), polychlorinated biphenyls (PCBs), aldehydes, and certain pesticides. In order to achieve a practical and reliable method for the determination of VOCs in food samples, several preparation methods have been developed, such as HS extraction and solid-phase microextraction (SPME).

2.1. HS Extraction

HS extraction is a sample preparation procedure that has demonstrated its usefulness for a broad range of VOCs in the headspace, which has been shown to reduce interference of the matrix for food substrates [2]. HS extraction is chiefly based on the adsorption of analytes on fiber coating. After establishing equilibrium between the HS of the sample and the fiber coating, components are desorbed from the fiber into a chromatography column. HS methods can be divided into static HS and dynamic HS (DHS) extraction. Static HS sampling is a conventional sample preparation method used for the analysis of VOCs from herbs and foods. It is a rapid and solvent-free method that requires only a small aliquot of samples [3]. DHS extraction can be performed by continuously sweeping the HS of the sample with a significant quantity of gas. Then, the extracted gas is loaded onto a selective adsorbent where analytes are trapped. Thermal desorption of trapped analytes is then required before conducting cryofocus GC–MS analysis. This approach has already been used to determine VOCs in fish and wine as well as to characterize olive oil [4,5,6].

2.2. Solid-Phase Microextraction (SPME)

SPME has been used for sample preparation of a wide range of foods due to its sensitivity and convenience of quantitation [7]. SPME can be regarded as a short GC column turned inside out. A fiber coating can be used as a filter to extract chemicals from different samples. SPME has the advantages of being simple to operate, highly efficient, solvent-free, and employed reagents can be reused. SPME can combine sampling, extraction, pre-enrichment, and injection into one step. Nowadays, commercial SPME fiber coatings, such as polydimethylsiloxane (PDMS) and PDMS/divinylbenzene (PDMS/DVB), are available, but some shortcomings, e.g., fragile needle, deciduous coating, short column lifetime, and insufficient thermal or solvent stability, limit their practical use [8,9,10]. To overcome these, many studies have focused in recent years on metals, such as gold, silver, platinum, titanium, copper, and stainless steel, as substrates for SPME fibers [11]. In addition, a growing trend in developing SPME coating deals with obtaining substrates with high corrosion resistance and high stability as well as high chemical activity and simple surface modification [12,13] (Table 1).

3. Preparation Methods for the Analysis of SVOCs or Thermally Stable Compounds

Common SVOCs or thermally stable compounds mostly include pesticides, ERs, carcinogens, such as mycotoxin, process contaminants, among others, which can lead to cancer or impair neurodevelopment in humans. In order to achieve a practical and reliable method, several preparation methods have been developed. Extraction by a solvent is the classic sample preparation technique, which includes liquid–liquid extraction (LLE), soxhlet extraction, solid–liquid extraction (SLE), microwave-assisted extraction (MAE), ultrasonic extraction (USE), accelerated solvent extraction (ASE), and supercritical fluid extraction (SFE) [14,15,16,17,18,19,20,21,22]. Improved extraction methods, such as pressurized liquid extraction (PLE), can increase the diffusion rate and solubility of interferences into the matrix [23,24]. Increasing common clean-up automation and improving instrument design has created a surge in the use of solid-phase extraction (SPE) in a variety of applications has been observed [25,26]. Based on SPE, novel miniaturized SPE methodologies, such as micro-solid phase extraction (MSPE), dispersive-MSPE (DMSPE), matrix solid-phase dispersion extraction (MSPD), and stir bar sorptive extraction (SBSE), have the advantage of requiring low amount of analytes, sorbent, and organic solvents [27,28,29]. Based on miniaturized sorbent-based extraction techniques, several methods have been developed for the analysis of contaminants in real food samples. The development of novel materials, e.g., magnetic molecularly imprinted polymers and other magnetic nanometer materials, with high selectivity to analytes that can at the same time eliminate the interference of the matrix and increase sensitivity and accuracy of the method, is still a promising research field [30,31].
As a general trend, ideal sample preparation techniques should be clean, selective, time-saving, cheap, simple, and environmentally friendly [32]. Compared to MS technology, MS/MS has been shown to accurately detect contaminants in multi-residue analysis and has become an analytical reality for food samples. Thus, in this section, we focused on multi-residue analysis suitable for GC–MS/MS developed in recent years and its matrix effect.

3.1. SPE

SPE is mainly based on solid-phase materials acting as sorbents of analytes which are further released under specific conditions. SPE employs a low consumption of organic solvent compared to conventional extraction techniques. However, the steps in the SPE procedure include activation of the SPE column, sample elution, and elution evaporation steps, which implicates a laborious procedure. Moreover, preventing high back pressure is difficult due to the tight packing of the SPE filler. Thus, SPE is combined with other extraction or clean-up procedures, such as LLE and ASE, to obtain more accurate results [33,34,35,36,37,38,39,40].
Based on SPE, the addition of a magnetic adsorbent to the sample with further dispersion with the aid of a vortex, shaker, or sonicator, upon which an external magnetic field is then applied to facilitate efficient retrieval of the magnetic adsorbent particles. Covalent organic frameworks, metal–organic frameworks (MOFs), and molecularly imprinted nanoparticles with uniform morphology offer superior selectivity, large adsorption capacity, and fast binding kinetics that can be used for selective recognition of analytes as well as enrichment and determination of many organic contaminants or pesticide residues [41,42]. Magnetic SPE coupled with GC–MS/MS enables group-selective extractions and detection with enhanced hydrophilicity, dispersibility, adsorptivity, and selectivity, which results in high recovery, precision, and sensitivity of analysis of food samples, being thus a promising alternative for reliable, efficient analysis.

3.2. MSPD

As a further development of the SPE method, MSPD is regarded as a promising technique that has been gaining extensive recognition due to the ability to reduce waste of samples and organic solvent. However, the biggest disadvantage of conventional MSPD dispersants (silica gel, C8, C18, etc.) is the lack of selectivity, which may cause interference from non-target substances with similar structures. MOFs, multi-walled carbon nanotubes (MWCNTs), and other newly developed nanomaterials have a high specific surface area and good chemical, mechanical, and thermostability properties and can serve as an adsorbent for enrichment and removal of organic contaminants. These nanomaterials can also be used in MSPD extraction for the pre-treatment of food samples, which have been shown to have short adsorption time, excellent selective recognition, simple operation, low cost, fast extraction efficiency, and low solvent consumption [43].

3.3. QuEChERS

The dispersive solid-phase extraction (d-SPE) technique employs acetonitrile extraction partitioning and requires few steps, which reduces the time required to complete the extraction and clean-up procedures [44,45,46,47]. The typical QuEChERS methodology was introduced as a green, user-friendly, and cheap approach to meet the challenge of analyzing trace residue of various organic compounds in foods of plant and animal origin, which are complex matrices. In recent years, the QuEChERS methodology has become increasingly popular as a method to determine contaminants in all kinds of food matrices [48,49,50,51,52,53,54,55,56,57,58,59,60,61,62,63,64,65,66,67,68,69,70,71,72,73,74,75,76,77,78,79,80,81,82,83]. The QuEChERS multi-method is predominantly suitable for the analysis of polar analytes. Target analytes for QuEChERS also include multi-class contaminants, such as pesticide residues, N-nitrosamines, veterinary drug residues, prohibited flavor compounds.
Certain substrates present in foods, such as tea, honeybees, meat, or leek, introduce a heavy matrix interference, which could cause poor peak shape or suppress ionization of analytes or low content target analytes. In this context, further dilution of extracts or the use of a clean-up process, such as USE, multiplug filtration clean-up (m-PFC), robotic clean-up, deep-frozen, SPE methods, are recommended to effectively reduce matrix interference [84,85,86]. The m-PFC method provides a more practical and effective way to perform the clean-up process than the conventional d-SPE method, offering a compromise between the clean-up performance of SPE and the convenience of d-SPE [44]. Since the QuEChERS method is mainly based on the penetration of water and acetonitrile into the sample tissue (at room temperature), if samples have poor water solubility, e.g., beeswax, analysis by the common QuEChERS method might not be efficient [18].
New materials as adsorbents are constantly being developed, among which can be cited tributylamine-functionalized graphene oxide (tri-BuA-rGO), graphitized MWCNTs, and ZrOx cluster [87]. These materials can be used as sorbents in sample preparation since they offer relatively high sensitivity for the analysis of pesticides [67]. Thus, the QuEChERS method is currently a widely used sample preparation method that can be coupled with GC–MS/MS detection and that allows simultaneous determination of a large number of contaminants in a sample.

3.4. SBSE

SBSE is a solid-phase extraction technique first described in 1999 [88]. In the past 15–20 years, miniaturized and solventless sample preparation techniques based on sorptive extraction methods, especially SBSE, have gained increasing popularity as simple and environmentally friendly alternatives to the aforementioned conventional methods. A glass-encased magnetic stir bar coated with a polymer, typically PDMS, is employed in this extraction method. In these techniques, extraction and concentration are performed in one step, while the use of a polar PDMS sorbent (which preferentially extracts non-polar compounds) does not require a clean-up process [89]. With SBSE, the sample preparation procedure and the number of solvents used are lower compared with multiple-step extraction procedures that are typically employed for the analysis of persistent organic pollutants (POPs) in solid samples. SBSE coupled with GC–MS has been used in the extraction and analysis of mainly hydrophobic organic compounds in aqueous samples.

3.5. Single-Drop Microextraction (SDME)

SDME is an analytical technique that uses only a small amount of water-immiscible solvent for concentrating analytes in aqueous samples. Its advantages include simplicity, low cost, the potential for automation, and the high yield of analytes from the sample matrix. Complex matrices, such as tea, require a step of extraction preparation. Extraction is limited by the partition coefficient of the analytes between the PDMS coating and the sample matrix, as well as by the phase ratio between the PDMS coating and the sample volume [90].

3.6. SPME Arrow

The SPME Arrow is a new SPME-based device that combines large sorption phase volumes used in SBSE and the main advantages of the conventional SPME method. The SPME Arrow device consists of a steel rod coated with a sorbent material protected by an outer tube which, together with the arrow-shaped tip, forms the needle. Classical SPME coatings are commercially available for the SPME Arrow. After analytes extraction, the stir bar is removed from the sample matrix and is placed in the thermal desorption unit that extracts analytes from the stir bar. Subsequently, the cooled injection system cryofocuses the desorbed analytes and injects them into the GC–MS/MS system for separation and detection. The SPME Arrow enhances sensitivity since it employs sorption phases that are larger and is hence a more robust technique. Since SPME has emerged only recently, few studies have demonstrated the suitability of this technique in the analysis of different kinds of compounds and samples, which include the analysis of organic compounds and PAHs in water and biogenic VOCs in atmospheric air [91,92].

3.7. Other Methods

Directly suspended droplet microextraction was developed to condense contaminants from foods through d-SPE prior to analysis by GC−MS/MS. The extractant is intentionally dispersed into the sample solution in the form of globules through high-speed agitation. This procedure increases the contact area between the binary phases and shortens equilibrium time [90]. The principle is similar to that of the dispersive liquid–liquid microextraction (DLLME) procedure; samples are dissolved into a dispersive solvent (e.g., acetonitrile) and then dissolved in water after being centrifuged and filtered. Lastly, chloroform is used to extract analytes in the DLLME method [93,94] (Table 2).

4. Separate, Clean-up, and Derivation Steps

Separate, clean-up, and derivation steps are usually used in further sample purification prior to food contaminants detection.

4.1. Gel Permeation Chromatography (GPC)

The GPC clean-up system can be used for the detection of contaminants accumulated in fat or liquid phases of foods. After coextraction of the target analyte and sample fat/liquid, the latter is removed in GPC by the solvent. The GPC clean-up system has been applied in foods of animal origin for further purification and detection of PAHs, chlorinated PAHs, polybrominated diphenyl ethers (PBDEs), and hexabromocyclododecanes (HBCDs) [95,96,97]. GPC has also been used in the combined detection of BaP, benz[a]anthracene (BaA), benzo[b]fluoranthene (BbF), and chrysene (Chr) in grains [98]. The amount of extracted fat depends on the properties of the target analyte, which thus influences the choice of the solvent.

4.2. Freezing-Lipid Method

Fat from foods of animal origin usually affects the results of food contaminants detection. Therefore, to separate lipids efficiently, the extract is often deep-frozen, during which a fairly significant amount of non-polar substances are lost [22]. A novel material, namely bond elute enhanced matrix removal (EMR), has a stronger selective adsorption affinity for lipids and has recently been used in the analysis of pesticides in kale, pork, salmon, chicken eggs, and avocado, also used to detect antibacterial drugs in cream disinfection products, specifically adsorbing C5 and long-chain hydrocarbons from lipids [99,100,101].

4.3. Derivatization

For single residue or selective residue detection, a derivatization is a useful tool for chromatography analysis which increases extraction rate while reducing detection limit. Especially for analytes with low volatility and high polarity, such as perfluoroalkyl carboxylic acids (PFCAs), PAEs, and 4-methylimidazole, the use of direct GC–MS/MS with selective ion monitoring (SIM) mode without derivatization of samples might not prove a sensitive method for adequate quantification [102].

5. Matrix Effect

In the multi-residue analysis of agricultural products, the matrix-induced chromatographic response enhancement, also termed matrix effect (ME), is a major issue that reduces the accuracy and precision of analytical results in food contaminants detection [103]. In general, the term matrix indicates miscellaneous substances that are extracted from samples, and the ME is defined as the direct or indirect alteration or interference in response due to the presence of unintended analytes (for analysis) or other interfering substances in the sample [104]. ME is determined by comparing the slope obtained for the standard calibration curve of the procedure and that of the solvent standard calibration curve, according to the following equation [105,106]:
Matrix   Effect % = ( Slope   of   calibration   curve   in   matrix Slope   of   calibration   curve   in   solvent 1 ) × 100 %
Several methods have been proposed to compensate for or overcome ME for various types of matrices, such as matrix-matched calibration, extensive clean-up, dilution, analyte protectant, standard addition method, and stable isotope-labeled pesticides standard mixtures. Dilution is a simple and effective method for reducing ME, but it requires highly sensitive analytical equipment to detect the maximum residue limit of pesticides in foods. Matrix-matched calibration, one of the most popular methods for the multi-residue analysis of agricultural products, is based on the theoretical concept that the ratio of the analytical response to the concentration of a target pesticide must match if the nature of the matrix of both sample and standard calibration solutions is identical. Moreover, internal isotope standards are very expensive and, usually, not available for all substances. Hence, matrix-matched calibration is likely the best strategy to overcome ME and generate highly accurate results [38].
Moreover, a matrix-like effect caused by the presence of pesticides in standard calibration solutions might result in unexpected response alterations and quantification errors, especially when using matrix-free, solvent standard solutions. Many studies suggest that non-polar compounds that have relatively low molecular weights are not susceptible to the matrix-like effect. Small molecules have a shorter residence time in the inlet liner (particularly glass wool), which thereby decreases the probability of ME. Both polarity and stability of a compound are important factors that influence the degree of response alteration caused by the matrix-like effect. It has been shown that highly polar compounds, such as organophosphates, have the potential for high adsorption interaction with active sites, and a fraction of these decompose in the GC system and are susceptible to response alteration induced by the food matrix and analyte protectants. Overcoming and/or compensating for matrix-like effects by improving the GC hardware or using certain food matrices and surrogate compounds (e.g., isotope-labeled pesticide standards) might be important topics for future research [107].

6. Emerging Risks

Emerging contaminants may represent a threat to human health and the environment in the future. Such contaminants, such as POPs, BFRs, PFOS, ERs, and other contaminants of unknown biological activity, may have a wide range of physicochemical properties. For instance, POPs are highly stable organic chemicals that resist photolytic, biological, and chemical degradation. POPs might persist in the environment, bioaccumulate through the food chain, and may adversely impact human health and the environment. Among POPs, PAHs, organochlorine pesticides (OCPs), PFCAs, and other compounds in dietary supplement samples can be included [47]. GC–MS/MS has been adopted to accurately quantify these emerging contaminants in order to overcome the low concentration and the complexity of the matrix, which limit the GC–MS/MS analysis [108].

6.1. BFRs

BFRs are a group of synthetic chemicals that comprise more than 75 compounds and that have been commercially available since 2003. PBDEs were among the first group of chemicals used worldwide to prevent flammability and have been used as flame retardant additives in a wide array of products, such as plastic materials, fabrics, and furniture. PBDEs comprise over 209 possible congeners produced in three major technical mixtures characterized by different degrees of bromination (penta-BDE, octa-BDE, and deca-BDE). As a consequence, PBDEs can be abundantly found in the environment, wildlife, and human tissue. Novel brominated compounds have been increasingly developed as an alternative to legacy BFRs. OH-PBDEs have oestrogenic, anti-prostagenic, antiandrogenic, and anti-glucocortogenic activity. PBDE congeners have been detected in foods consumed by the general population. Importantly, both PBDEs and OH-PBDEs can alter calcium homeostasis and disrupt intracellular communication [109,110,111]. QuEChERS is a suitable preparation method for PBDEs detection; however, when using other extraction methods, lipids and proteins need to be removed from the sample by using KOH solution/n-hexane partition or GPC process. Lipids and pigments that remain in the extract can be removed using the Florisil cartridge [23,112].

6.2. PFOS

PFCAs are a class of PFOS in which hydrogen atoms on the carbon skeleton are completely replaced by fluorine atoms and linked to the carboxylic acid functional groups. The unique physical and chemical characteristics imparted by the fluorinated region of the molecule include water and oil repellency, heat resistance, and surfactant properties, making these compounds suitable for a wide range of industrial and commercial applications. The hardly-degradable characteristic and wide distribution of PFCAs contribute to the persistence of these compounds in the environment and to their bioaccumulation [101]. QuEChERS is the most recommended extraction and separation method for detecting PFCAs [113].

6.3. ERs

Many chemical pollutants in the food chain can be considered ERs, which may include certain POPs and their metabolites, pesticides, PAEs, and hormones [114]. PAEs are widely used as plasticizers in various plastic products and packaging and are considered to produce developmental toxicity due to their potential to interfere with hormone homeostasis. Among PAEs, bisphenol A (BPA, 2,2-bis(4-hydroxyphenyl)propane) is used worldwide. QuEChERS, SPE, coupled with solvent extraction methods, are commonly used for PAEs and BPA extraction. Due to the relatively non-volatile and polar nature of PAEs and BPA, derivatization of samples for GC–MS/MS analysis is usually required [115].
PCBs are a group of ubiquitous and organic pollutants with chlorine atoms present in different frequencies and positions on two coupled biphenyl rings. PCBs have been produced for many years and are widely employed in the industry as heat exchange fluids in electric transformers and capacitors, as well as additives in pesticides, paints, sealants, and plastics. Conventional techniques, such as d-SPE or QuEChERS, represent a good choice for the analysis of PCBs in terms of selectivity, accuracy, and precision [116,117].

6.4. Mycotoxins

Mycotoxins are secondary toxic metabolites produced by various fungal species (mainly Aspergillus, Penicillium, and Fusarium) and represent an important group of contaminants, particularly in food. Mycotoxins are generally stable, resistant to various processing methods employed in the food industry and could even be found in thermally-processed foods. Several extraction procedures have been reported, such as LLE, SPE, and QuEChERS. However, the QuEChERS method has several advantages over other methods since it is a rapid protocol that requires only small amounts of organic solvents and adequately yields recoveries for all compounds [34,118].

6.5. Process Contaminants

Process contaminants are chemicals that are generated when food constituents undergo chemical changes during processing. Prime contaminants, such as acrylamide, PAHs, or furan and furan derivatives, are well-studied. The processes that often introduce PAHs into foods are smoking and grilling [119]. Unrefined plant oils obtained from oilseeds, such as soybeans, rapeseeds, olive seeds, and sunflower seeds, are known to contain high levels of polyaromatic hydrocarbons. PAHs can migrate through the food chain due to their high lipophilicity and accumulate in specific tissues. A new class of PAHs derivatives, known as chlorinated PAHs (Cl-PAHs), have been attracting increasing interest. Emerging Cl-PAHs are also ubiquitous and hazardous pollutants akin to PAHs and other halogenated aromatic compounds, such as polychlorinated dibenzo-p-dioxins, dibenzofurans, and PCBs; the QuEChERS method has been used for sample preparation [120].
Fatty acid esters of monochloropropanediol (MCPDEs) and glycidol (GEs) are emerging process contaminants that are often found in oil-containing products. An SPE clean-up process was used to remove partial glycerides to a certain extent [121]. Ice bath-assisted sodium hydroxide purification can be used for sample extraction for the analysis of ethyl carbamate (EC) and N-nitrosoamines (NAs), which are toxic contaminants found in fermented alcoholic beverages [122].

6.6. Contaminants with Unknown Biological Activity

9,10-anthraquinone (AQ) is a new contaminant of unknown sources occurring in tea globally [123]. Moreover, AQ is ubiquitous and currently used as a raw chemical in the paper, pulp, and dye industries. AQ may contribute to the carcinogenic potential of foods. A low level of AQ found in tea plants may be one of the sources of AQ contamination in tea (Table 3).

7. Conclusions

This review presents a discussion of current sample preparation methods for the analysis of food contaminants and residues by GC–MS/MS. The main criteria to consider on preparation methods for VOCs analysis, HS extraction, and SPME methods are sensitivity and easiness of quantitation of target analytes. Moreover, GC–MS/MS is gradually being preferred to GC–MS for its higher sensitivity and better separation rather than preparation selectivity. Thus, for the analysis of more SVOCs or thermally stable compounds, it might be advisable to develop methods for the analysis of multiple contaminants of different classes considering a single sample preparation technique and preferably one chromatographic run. Multi-residue preparation methods, such as QuEChERs and adsorption extraction coupled with SBSE and SDME, are the most important technologies currently available, enabling the extraction of as many targets as possible simultaneously. Due to the complexity of certain food matrices, the use of separation, clean-up, and derivatization processes might be useful to increase the extraction rate and reduce the detection limit. In addition, we also proposed guidelines for determining whether ME might occur and interfere with the results; matrix-matched calibration is one of the most popular methods for the multi-residue analysis of agricultural products. Moreover, a discussion of emerging contaminants that may threaten human health and the environment in the future is provided. QuEChERs might be employed to detect these emerging contaminants. In conclusion, a comprehensive strategy is required for sample preparation for the analysis of food contaminants and residues by GC–MS/MS in order to effectively achieve food safety.

Author Contributions

Conceptualization, M.-L.X.; writing—original draft preparation, M.-L.X. and Y.G.; writing—review and editing, X.X.H. and X.W.; funding acquisition, B.Z. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by the National Natural Science Foundation of the People’s Republic of China, grant number 22011540378, 21773080, 21773079, the Development Program of the Science and Technology of the Jilin Province, grant number 20190201215jc, 20200404193YY. The APC was funded by 22011540378.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Feng, C.; Xu, Q.; Qiu, X.; Jin, Y.; Ji, J.; Lin, Y.; Le, S.; Wang, G.; Lu, D. Comprehensive strategy for analysis of pesticide multi-residues in food by GC–MS/MS and UPLC–Q–Orbitrap. Food Chem. 2020, 320, 126576–126584. [Google Scholar] [CrossRef] [PubMed]
  2. Yue, Q.; Huang, Y.-Y.; Shen, X.-F.; Yang, C.; Pang, Y.-H. In situ growth of covalent organic framework on titanium fiber for headspace solid-phase microextraction of 11 phthalate esters in vegetables. Food Chem. 2020, 318, 126507–126515. [Google Scholar] [CrossRef] [PubMed]
  3. Li, J.; Zhang, Z.; Sun, M.; Zhang, B.; Fan, C. Use of a headspace solid-phase microextraction-based methodology followed by gas chromatography–tandem mass spectrometry for pesticide multiresidue determination in teas. Chromatographia 2018, 81, 809–821. [Google Scholar] [CrossRef]
  4. Guiffard, I.; Geny, T.; Veyrand, B.; Marchand, P.; Pellouin-Grouhel, A.; Bizec, B.L.; Bichon, E. Quantification of light polycyclic aromatic hydrocarbons in seafood samples using on–line dynamic headspace extraction, thermodesorption, gas chromatography tandem mass spectrometry, based on an isotope dilution approach. J. Chromatogr. A 2020, 1619, 460906–460914. [Google Scholar] [CrossRef]
  5. Moyano, L.; Serratosa, M.P.; Marquez, A.; Zea, L. Optimization and validation of a DHS–TD–GC–MS method to wineomics studies. Talanta 2019, 192, 301–307. [Google Scholar] [CrossRef] [PubMed]
  6. Sales, C.; Portoles, T.; Johnsen, L.G.; Danielsen, M.; Beltran, J. Olive oil quality classification and measurement of its organoleptic attributes by untargeted GC–MS and multivariate statistical–based approach. Food Chem. 2019, 271, 488–496. [Google Scholar] [CrossRef] [PubMed]
  7. Trabalón, L.; Nadal, M.; Borrull, F.; Pocurull, E. Determination of benzothiazoles in seafood species by subcritical water extraction followed by solid–phase microextraction–gas chromatography–tandem mass spectrometry: Estimating the dietary intake. Anal. Bioanal. Chem. 2017, 409, 5513–5522. [Google Scholar] [CrossRef]
  8. Hernandes, K.C.; Souza-Silva, É.A.; Assumpção, C.F.; Zini, C.A.; Welke, J.E. Validation of an analytical method using HS–SPME–GC/MS–SIM to assess the exposure risk to carbonyl compounds and furan derivatives through beer consumption. Food Addit. Contam. A 2019, 36, 61–68. [Google Scholar] [CrossRef]
  9. Ma, T.-T.; Shen, X.-F.; Yang, C.; Qian, H.-L.; Pang, Y.-H.; Yan, X.-P. Covalent immobilization of covalent organic framework on stainless steel wire for solid–phase microextraction GC–MS/MS determination of sixteen polycyclic aromatic hydrocarbons in grilled meat samples. Talanta 2019, 201, 413–418. [Google Scholar] [CrossRef] [PubMed]
  10. Guo, J.-X.; Qian, H.-L.; Zhao, X.; Yang, C.; Yan, X.-P. In situ room–temperature fabrication of a covalent organic framework and its bonded fiber for solid–phase microextraction of polychlorinated biphenyls in aquatic products. J. Mater. Chem. A 2019, 7, 13249–13256. [Google Scholar] [CrossRef]
  11. Wang, Z.; Jin, P.; Zhou, S.; Wang, X.; Du, X. Controlled growth of a porous hydroxyapatite nanoparticle coating on a titanium fiber for rapid and efficient solid–phase microextraction of polar chlorophenols, triclosan and bisphenol A from environmental water. Anal. Methods 2018, 10, 3237–3248. [Google Scholar] [CrossRef]
  12. Castro, Ó.; Trabalón, L.; Schilling, B.; Borrull, F.; Pocurull, E. Solid phase microextraction arrow for the determination of synthetic musk fragrances in fish samples. J. Chromatogr. A 2019, 1591, 55–61. [Google Scholar] [CrossRef]
  13. Žnideršič, L.; Mlakar, A.; Prosen, H. Development of a SPME–GC–MS/MS method for the determination of some contaminants from food contact material in beverages. Food Chem. Toxicol. 2019, 134, 110829–110840. [Google Scholar] [CrossRef] [PubMed]
  14. Wu, P.; Gu, M.; Wang, Y.; Xue, J.; Wu, X. Transfer of organochlorine pesticide residues during household and industrial processing of ginseng. J. Food Qual. 2020, 2020, 5946078–5946087. [Google Scholar] [CrossRef]
  15. García-Córcoles, M.T.; Cipa, M.; Rodríguez-Gómez, R.; Rivas, A.; Olea-Serrano, F.; Vílchez, J.L.; Zafra-Gómez, A. Determination of bisphenols with estrogenic activity in plastic packaged baby food samples using solid–liquid extraction and clean–up with dispersive sorbents followed by gas chromatography tandem mass spectrometry analysis. Talanta 2018, 178, 441–448. [Google Scholar] [CrossRef]
  16. Christodoulou, L.D.; Kourouzidou, O.; Hadjigeorgiou, M.; Hadjiloizou, P.; Constantinou, M.; Constantinou, P.; Kika, K.; Klavarioti, M. Multi–residue analysis of pesticide residues in fruits and vegetables using gas and liquid chromatography with mass spectrometric detection. Accredit. Qual. Assur. 2018, 23, 145–175. [Google Scholar] [CrossRef]
  17. Lakhmanov, D.; Varakina, Y.; Aksenov, A.; Sorokina, T.; Sobolev, N.; Kotsur, D.; Plakhina, E.; Chashchin, V.; Thomassen, Y. Persistent organic pollutants (POPs) in fish consumed by the indigenous peoples from Nenets Autonomous Okrug. Environments 2020, 7, 3. [Google Scholar] [CrossRef] [Green Version]
  18. Issa, M.M.; Taha, S.M.; El- Marsafy, A.M.; Khalil, M.M.H.; Ismail, E.H. Acetonitrile–ethyl acetate based method for the residue analysis of 373 pesticides in beeswax using LC–MS/MS and GC–MS/MS. J. Chromatogr. B 2020, 1145, 122106–122112. [Google Scholar] [CrossRef]
  19. Panseri, S.; Chiesa, L.; Ghisleni, G.; Marano, G.; Boracchi, P.; Ranghieri, V.; Malandra, R.M.; Roccabianca, P.; Tecilla, M. Persistent organic pollutants in fish: Biomonitoring and cocktail effect with implications for food safety. Food Addit. Contam. A 2019, 36, 601–611. [Google Scholar] [CrossRef]
  20. Zhao, X.; Wang, B.; Xie, K.; Liu, Y.; Zhang, Y.; Wang, Y.; Liu, C.; Guo, Y.; Bu, X.; Zhang, G.; et al. Development of an ASE–GC–MS/MS method for detecting dinitolmide and its metabolite 3–ANOT in eggs. J. Mass. Spectrom. 2018, 53, 976–985. [Google Scholar] [CrossRef] [PubMed]
  21. Chau, N.D.G.; Hop, N.V.; Long, H.T.; Duyen, N.T.M.; Raber, G. Multi–residue analytical method for trace detection of new–generation pesticides in vegetables using gas chromatography–tandem mass spectrometry. J. Environ. Sci. Heal. B 2020, 55, 417–428. [Google Scholar] [CrossRef]
  22. Sturm, J.; Wienhold, P.; Frenzel, T.; Speer, K. Ultra turrax® tube drive for the extraction of pesticides from egg and milk samples. Anal. Bioanal. Chem. 2018, 410, 5431–5438. [Google Scholar] [CrossRef] [PubMed]
  23. Zhang, M.; Guo, W.; Wei, J.; Shi, J.; Zhang, J.; Ge, H.; Tao, H.; Liu, X.; Hu, Q.; Cai, Z. Determination of newly synthesized dihydroxylated polybrominated diphenyl ethers in sea fish by gas chromatography–tandem mass spectrometry. Chemosphere 2020, 240, 124878. [Google Scholar] [CrossRef] [PubMed]
  24. Julia, M.; Luis, S.J.; Luis, M.J.; Irene, A.; Esteban, A. Determination of bisphenol A, its chlorinated derivatives and structural analogues in vegetables by focussed ultrasound solid–liquid extraction and GC–MS/MS. Environ. Chem. 2019, 17, 266–277. [Google Scholar] [CrossRef]
  25. Shin, Y.; Lee, J.; Kim, J.-H. A simultaneous multiresidue analysis for 203 pesticides in soybean using florisil solid–phase extraction and gas chromatography–tandem mass spectrometry. Appl. Biol. Chem. 2018, 61, 543–548. [Google Scholar] [CrossRef] [Green Version]
  26. Qin, X.; Luo, X.; Han, J.; Chen, Y.; Zhang, K.; Hu, D. Residual determination of pyrethrins in Lyciumbarbarum (goji) by GC–MS/MS and a dietary risk assessment of Chinese goji consumption. Food Addit. Contam. A 2020, 37, 478–487. [Google Scholar] [CrossRef]
  27. Li, W.-K.; Zhang, H.-X.; Shi, Y.-P. Simultaneous determination of bifenox, dichlobenil and diclofop methyl by hollow carbon nanospheres enhanced magnetic carboxylic multi–walled carbon nanotubes. Anal. Chim. Acta 2018, 1011, 40–49. [Google Scholar] [CrossRef] [PubMed]
  28. Li, W.; Liu, D.; Li, J.; Gao, J.; Zhang, C.; Wang, P.; Zhou, Z. Matrix solid–phase dispersion combined with GC–MS/MS for the determination of organochlorine pesticides and polychlorinated biphenyls in marketed seafood. Chromatographia 2017, 80, 813–824. [Google Scholar] [CrossRef]
  29. Rahbar, N.; Behrouz, E.; Ramezani, Z. One–step synthesis of zirconia and magnetite nanocomposite immobilized chitosan for micro–solid–phase extraction of organophosphorous pesticides from juice and water samples prior to gas chromatography/mass spectroscopy. Food Anal. Method 2017, 10, 2229–2240. [Google Scholar] [CrossRef]
  30. Xu, X.; Duhoranimana, E.; Zhang, X. Preparation and characterization of magnetic molecularly imprinted polymers for the extraction of hexamethylenetetramine in milk samples. Talanta 2017, 163, 31–38. [Google Scholar] [CrossRef]
  31. Huang, X.; Liu, Y.; Liu, H.; Liu, G.; Xu, X.; Li, L.; Lv, J.; Gao, H.; Xu, D. Magnetic solid—Phase extraction of pyrethroid insecticides from tea infusions using ionic liquid–modified magnetic zeolitic imidazolate framework—8 as an adsorbent. RSC Adv. 2019, 9, 39272–39281. [Google Scholar] [CrossRef] [Green Version]
  32. Luo, P.; Liu, X.; Kong, F.; Tang, L.; Wang, Q.; Li, W.; Xu, W.; Wen, S.; Chen, L.; Li, Y. Multi–residue determination of 325 pesticides in chicken eggs with EMR–Lipid clean–up by UHPLC–MS/MS and GC–MS/MS. Chromatographia 2020, 83, 593–599. [Google Scholar] [CrossRef]
  33. Han, L.; Sapozhnikova, Y. Semi–automated high–throughput method for residual analysis of 302 pesticides and environmental contaminants in catfish by fast low–pressure GC–MS/MS and UHPLC–MS/MS. Food Chem. 2020, 319, 126592–126600. [Google Scholar] [CrossRef]
  34. Wang, B.; Wang, Y.; Xie, X.; Diao, Z.; Xie, K.; Zhang, G.; Zhang, T.; Dai, G. Quantitative analysis of spectinomycin and lincomycin in poultry eggs by accelerated solvent extraction coupled with gas chromatography tandem mass spectrometry. Foods 2020, 9, 651. [Google Scholar] [CrossRef]
  35. Zhao, X.; Wang, B.; Xie, K.; Liu, Y.; Zhang, Y.; Wang, Y.; Guo, Y.; Bu, X.; Liu, C.; Zhang, G.; et al. Determination of dinitolmide and its metabolite 3–ANOT in chicken tissues via ASE–SPE–GC–MS/MS. J. Food Compos. Anal. 2018, 71, 94–103. [Google Scholar] [CrossRef]
  36. Sapozhnikova, Y. High–throughput analytical method for 265 pesticides and environmental contaminants in meats and poultry by fast low pressure gas chromatography and ultrahigh–performance liquid chromatography tandem mass spectrometry. J. Chromatogr. A 2018, 1572, 203–211. [Google Scholar] [CrossRef]
  37. Wang, X.; Wang, Z.; Di, S.; Xue, X.; Min, S. Determination of 14 lipophilic pesticide residues in raw propolis by selective sample preparation and gas chromatography–tandem mass spectrometry. Food Anal. Method 2020, 13, 1726–1735. [Google Scholar] [CrossRef]
  38. Ly, T.-K.; Ho, T.-D.; Behra, P.; Nhu-Trang, T.-T. Determination of 400 pesticide residues in green tea leaves by UPLC–MS/MS and GC–MS/MS combined with QuEChERS extraction and mixed–mode SPE clean–up method. Food Chem. 2020, 326, 126928–126936. [Google Scholar] [CrossRef]
  39. Kang, H.S.; Kim, M.K.; Kim, E.J. High–throughput simultaneous analysis of multiple pesticides in grain, fruit, and vegetables by GC–MS/MS. Food Addit. Contam. A 2020, 37, 963–972. [Google Scholar] [CrossRef]
  40. Wong, Y.-T.; Law, W.-K.; Lai, S.S.-L.; Wong, S.-P.; Lau, K.-C.; Ho, C. Ultra–trace determination of sodium fluoroacetate (1080) as monofluoroacetate in milk and milk powder by GC–MS/MS and LC–MS/MS. Anal. Methods 2018, 10, 3514–3524. [Google Scholar] [CrossRef]
  41. Pang, Y.-H.; Yue, Q.; Huang, Y.; Yang, C.; Shen, X.-F. Facile magnetization of covalent organic framework for solid–phase extraction of 15 phthalate esters in beverage samples. Talanta 2020, 206, 120194–120204. [Google Scholar] [CrossRef]
  42. Lu, C.; Tang, Z.; Gao, X.; Ma, X.; Liu, C. Computer–aided design of magnetic dummy molecularly imprinted polymers for solid–phase extraction of ten phthalates from food prior to their determination by GC–MS/MS. Microchim. Acta 2018, 185, 373–384. [Google Scholar] [CrossRef] [PubMed]
  43. Liang, T.; Wang, S.; Chen, L.; Niu, N. Metal organic framework–molecularly imprinted polymer as adsorbent in matrix solid phase dispersion for pyrethroids residue extraction from wheat. Food Anal. Method 2019, 12, 217–228. [Google Scholar] [CrossRef]
  44. Han, Y.; Song, L.; Liu, S.; Zou, N.; Li, Y.; Qin, Y.; Li, X.; Pan, C. Simultaneous determination of 124 pesticide residues in Chinese liquor and liquor–making raw materials (sorghum and rice hull) by rapid multi–plug filtration cleanup and gas chromatography–tandem mass spectrometry. Food Chem. 2018, 241, 258–267. [Google Scholar] [CrossRef]
  45. Zhu, B.; Xu, X.; Luo, J.; Jin, S.; Chen, W.; Liu, Z.; Tian, C. Simultaneous determination of 131 pesticides in tea by on–line GPC–GC–MS/MS using graphitized multi–walled carbon nanotubes as dispersive solid phase extraction sorbent. Food Chem. 2019, 276, 202–208. [Google Scholar] [CrossRef] [PubMed]
  46. Wang, F.; Li, S.; Feng, H.; Yang, Y.; Xiao, B.; Chen, D. An enhanced sensitivity and cleanup strategy for the nontargeted screening and targeted determination of pesticides in tea using modified dispersive solid–phase extraction and cold–induced acetonitrile aqueous two–phase systems coupled with liquid chromatography–high resolution mass spectrometry. Food Chem. 2019, 275, 530–538. [Google Scholar] [CrossRef]
  47. Zacs, D.; Rozentale, I.; Reinholds, I.; Bartkevics, V. Multi–walled carbon nanotubes as effective sorbents for rapid analysis of polycyclic aromatic hydrocarbons in edible oils using dispersive solid–phase extraction (d–SPE) and gas chromatography—tandem mass spectrometry (GC–MS/MS). Food Anal. Method 2018, 11, 2508–2517. [Google Scholar] [CrossRef]
  48. Naik, R.H.; Pallavi, M.S.; Bheemanna, M.; PavanKumar, K.; Reddy, V.C.S.; Nidoni, R.U.; Paramasivam, M.; Yadav, S. Simultaneous determination of 79 pesticides in pigeonpea grains using GC–MS/MS and LC–MS/MS. Food Chem. 2021, 347, 128986–128997. [Google Scholar] [CrossRef]
  49. Gao, G.; Chen, H.; Dai, J.; Jin, L.; Chai, Y.; Zhu, L.; Liu, X.; Lu, C. Determination of polychlorinated biphenyls in tea using gas chromatography–tandem mass spectrometry combined with dispersive solid phase extraction. Food Chem. 2020, 316, 126290–126298. [Google Scholar] [CrossRef]
  50. Song, N.-E.; Lee, J.Y.; Mansur, A.R.; Jang, H.W.; Lim, M.-C.; Lee, Y.; Yoo, M.; Nam, T.G. Determination of 60 pesticides in hen eggs using the QuEChERS procedure followed by LC–MS/MS and GC–MS/MS. Food Chem. 2019, 298, 125050–125060. [Google Scholar] [CrossRef]
  51. Yu, X.; Zhang, Z.; Li, W.; Zhang, R.; Jiao, H.; Zhao, J.; Sun, A.; Shi, X.; Chen, J. Development and application of the dispersive solid–phase extraction method based on molecular imprinted polymers for removal of matrix components of bivalve shellfish extracts in the GC–MS/MS analysis of amide/dinitroaniline/substituted urea herbicides. Chromatographia 2019, 82, 961–970. [Google Scholar] [CrossRef]
  52. Gaweł, M.; Kiljanek, T.; Niewiadowska, A.; Semeniuk, S.; Goliszek, M.; Burek, O.; Posyniak, A. Determination of neonicotinoids and 199 other pesticide residues in honey by liquid and gas chromatography coupled with tandem mass spectrometry. Food Chem. 2019, 282, 36–47. [Google Scholar] [CrossRef]
  53. Chiesa, L.M.; Panseri, S.; Nobile, M.; Ceriani, F. ArioliDistribution of POPs, pesticides and antibiotic residues in organic honeys from different production areas. Food Addit. Contam. A 2018, 35, 1340–1355. [Google Scholar] [CrossRef]
  54. Jadhav, M.R.; Pudale, A.; Raut, P.; Utture, S.; Shabeer, T.P.A.; Banerjee, K. A unified approach for high–throughput quantitative analysis of the residues of multi–class veterinary drugs and pesticides in bovine milk using LC–MS/MS and GC–MS/MS. Food Chem. 2019, 272, 292–305. [Google Scholar] [CrossRef] [PubMed]
  55. Li, P.; Duan, Y.; Ge, H.; Zhang, Y.; Wu, X. Multiresidue Analysis of 113 pesticides in different maturity levels of mangoes using an optimized QuEChERS method with GC–MS/MS and UHPLC–MS/MS. Food Anal. Methods 2018, 11, 2742–2757. [Google Scholar] [CrossRef]
  56. Lachter, D.R.; Nudi, A.H.; Porto, G.F.; Ribeiro, R.A.L.; Wagener, A.L.R.; Masone, C.G. Multiresidue method for triazines and pyrethroids determination by solid–phase extraction and gas chromatography–tandem mass spectrometry. J. Environ. Sci. Heal. B 2020, 55, 865–875. [Google Scholar] [CrossRef] [PubMed]
  57. Kumari, D.; John, S. Health risk assessment of pesticide residues in fruits and vegetables from farms and markets of Western Indian Himalayan region. Chemosphere 2019, 224, 162–167. [Google Scholar] [CrossRef] [PubMed]
  58. Hakme, E.; Lozano, A.; Uclés, S.; Fernández-Alba, A.R. Further improvements in pesticide residue analysis in food by applying gas chromatography triple quadrupole mass spectrometry (GC–QqQ–MS/MS) technologies. Anal. Bioanal. Chem. 2018, 410, 5491–5506. [Google Scholar] [CrossRef] [PubMed]
  59. Lin, X.-Y.; Mou, R.-X.; Cao, Z.-Y.; Cao, Z.-Z.; Chen, M.-X. Analysis of pyrethroid pesticides in Chinese vegetables and fruits by GC–MS/MS. Chem. Pap. 2018, 72, 1953–1962. [Google Scholar] [CrossRef]
  60. Wei, Q.; Wu, M.; Xiao, F.; Wang, D. Development of a fast method for the determination of pesticide multiresidues in tomatoes using QuEChERS and GC–MS/MS. Eur. Food Res. Technol. 2020, 246, 1563–1572. [Google Scholar] [CrossRef]
  61. Saito-Shida, S.; Kashiwabara, N.; Shiono, K.; Nemoto, S.; Akiyama, H. Development of an analytical method for determination of total ethofumesate residues in foods by gas chromatography–tandem mass spectrometry. Food Chem. 2020, 313, 126132–126140. [Google Scholar] [CrossRef]
  62. Murcia-Morales, M.; Cutillas, V.; Fernández-Alba, A.R. Supercritical fluid chromatography and gas chromatography coupled to tandem mass spectrometry for the analysis of pyrethroids in vegetable matrices: A comparative study. J. Agr. Food Chem. 2019, 67, 12626–12632. [Google Scholar] [CrossRef]
  63. Reichert, B.; Nunes, M.G.P.; Pizzutti, I.R.; Costabeber, I.H.; Fontana, M.Z.; Jänich, B.D.; Panciera, M.P.; Arbusti, D.; Cardoso, C.D.; Chim, J.F. Pesticide residues determination in common bean using an optimized QuEChERS approach followed by solvent exchange and GC–MS/MS analysis. J. Sci. Food Agric. 2020, 100, 2425–2434. [Google Scholar] [CrossRef]
  64. Ramadan, M.F.A.; Abdel-Hamid, M.M.A.; Altorgoman, M.M.F.; AlGaramah, H.A.; Alawi, M.A.; Shati, A.A.; Shweeta, H.A.; Awwad, N.S. Evaluation of pesticide residues in vegetables from the Asir region, Saudi Arabia. Molecules 2020, 25, 205. [Google Scholar] [CrossRef] [Green Version]
  65. Song, L.; Zhong, Z.; Han, Y.; Zheng, Q.; Qin, Y.; Wu, Q.; He, X.; Pan, C. Dissipation of sixteen pesticide residues from various applications of commercial formulations on strawberry and their risk assessment under greenhouse conditions. Ecotox. Environ. Safe. 2020, 188, 109842–109852. [Google Scholar] [CrossRef]
  66. Varela-Martínez, D.A.; González-Curbelo, M.Á.; González-Sálamo, J.; Hernández-Borges, J. Determination of pesticides in dried minor tropical fruits from Colombia using the Quick, Easy, Cheap, Effective, Rugged, and Safe method–gas chromatography tandem mass spectrometry. J. Sep. Sci 2020, 43, 929–935. [Google Scholar] [CrossRef]
  67. Mao, X.; Yan, A.; Wan, Y.; Luo, D.; Yang, H. Dispersive solid–phase extraction using microporous sorbent UiO–66 coupled to gas chromatography–tandem mass spectrometry: A QuEChERS–type method for the determination of organophosphorus pesticide residues in edible vegetable oils without matrix interference. J. Agric. Food Chem. 2019, 67, 1760–1770. [Google Scholar] [CrossRef]
  68. Biswas, S.; Mondal, R.; Mukherjee, A.; Sarkar, M.; Kole, R.K. Simultaneous determination and risk assessment of fipronil and its metabolites in sugarcane, using GC–ECD and confirmation by GC–MS/MS. Food Chem. 2019, 272, 559–567. [Google Scholar] [CrossRef] [PubMed]
  69. Geng, Y.; Jiang, L.; Jiang, H.; Wang, L.; Peng, Y.; Wang, C.; Shi, X.; Gu, J.; Wang, Y.; Zhu, J.; et al. Assessment of heavy metals, fungicide quintozene and its hazardous impurity residues in medical Panax notoginseng (Burk) F.H.Chen root. Biomed. Chromatogr. 2019, 33, e4378–e4386. [Google Scholar] [CrossRef] [PubMed]
  70. Li, S.; Yu, P.; Zhou, C.; Tong, L.; Li, D.; Yu, Z.; Zhao, Y. Analysis of pesticide residues in commercially available chenpi using a modified QuEChERS method and GC–MS/MS determination. J. Pharm. Anal. 2020, 10, 60–69. [Google Scholar] [CrossRef] [PubMed]
  71. Duan, X.; Tong, L.; Li, D.; Yu, Z.; Zhao, Y. A Multiresidue method for simultaneous determination of 116 pesticides in notoginseng radix et rhizoma using modified QuEChERS coupled with gas chromatography tandem mass spectrometry and census 180 batches of sample from Yunnan Province. Chromatographia 2018, 81, 545–556. [Google Scholar] [CrossRef]
  72. Rutkowska, E.; Łozowicka, B.; Kaczyński, P. Modification of multiresidue QuEChERS protocol to minimize matrix effect and improve recoveries for determination of pesticide residues in dried herbs followed by GC–MS/MS. Food Anal. Methods 2018, 11, 709–724. [Google Scholar] [CrossRef]
  73. Fu, Y.; Dou, X.; Zhang, L.; Qin, J.; Yang, M.; Luo, J. A comprehensive analysis of 201 pesticides for different herbal species–ready application using gas chromatography–tandem mass spectrometry coupled with QuEChERs. J. Chromatogr. B 2019, 1125, 121730–121742. [Google Scholar] [CrossRef]
  74. Shabeer, T.P.A.; Girame, R.; Utture, S.; Oulkar, D.; Banerjee, K.; Ajay, D.; Arimboor, R.; Menon, K.R.K. Optimization of multi–residue method for targeted screening and quantitation of 243 pesticide residues in cardamom (Elettaria cardamomum) by gas chromatography tandem mass spectrometry (GC–MS/MS) analysis. Chemosphere 2018, 193, 447–453. [Google Scholar] [CrossRef] [PubMed]
  75. Abdalla, A.A.; Afify, A.S.; Hasaan, I.E.; Amr, M. Studying the effect of household–type treatment and processing on the residues of ethion and profenofos pesticides and on the contents of capsaicinoids in green chili pepper using GC–MS/MS and HPLC. Food Anal. Methods 2018, 11, 382–393. [Google Scholar] [CrossRef]
  76. Ahlawat, S.; Gulia, S.; Malik, K.; Rani, S.; Chauhan, R. Persistence and decontamination studies of chlorantraniliprole in Capsicum annum using GC–MS/MS. J. Food Sci. Tech. 2019, 56, 2925–2931. [Google Scholar] [CrossRef] [PubMed]
  77. Song, L.; Han, Y.; Yang, J.; Qin, Y.; Zeng, W.; Xu, S.; Pan, C. Rapid single–step cleanup method for analyzing 47 pesticide residues in pepper, chili peppers and its sauce product by high performance liquid and gas chromatography–tandem mass spectrometry. Food Chem. 2019, 279, 237–245. [Google Scholar] [CrossRef]
  78. Tian, F.; Qiao, C.; Luo, J.; Xie, H. Method development and validation of ten pyrethroid insecticides in edible mushrooms by modified QuEChERS and gas chromatography–tandem mass spectrometry. Sci. Rep. 2020, 10, 7042–7052. [Google Scholar] [CrossRef]
  79. Bella, G.D.; Potortì, A.G.; Tekaya, A.B.; Beltifa, A.; Mansour, H.B.; Sajia, E.; Bartolomeo, G.; Naccari, C.; Dugo, G.; Turco, V.L. Organic contamination of Italian and Tunisian culinary herbs and spices. J. Environ. Sci. Heal. B 2019, 54, 345–356. [Google Scholar] [CrossRef]
  80. Beneta, A.; Pavlović, D.M.; Periša, I.; Petrović, M. Multiresidue GC–MS/MS pesticide analysis for evaluation of tea and herbal infusion safety. Int. J. Environ. An. Chem. 2018, 98, 987–1004. [Google Scholar] [CrossRef]
  81. Rutkowska, E.; Łozowicka, B.; Kaczyński, P. Three approaches to minimize matrix effects in residue analysis of multiclass pesticides in dried complex matrices using gas chromatography tandem mass spectrometry. Food Chem. 2019, 279, 20–29. [Google Scholar] [CrossRef] [PubMed]
  82. Li, J.; Liu, H.; Wang, C.; Yang, J.; Han, G. Stable isotope labeling–assisted GC/MS/MS method for determination of methyleugenol in food samples. J. Sci. Food Agric. 2018, 98, 3485–3491. [Google Scholar] [CrossRef]
  83. Chen, H.; Ling, Y.; Zhang, F.; Liu, T.; Wang, J.-F.; Wu, H.-Q.; Hong, Y.-H.; Cheng, Y. Simultaneous detection of eight prohibited flavor compounds in foodstuffs using gas chromatography–tandem mass spectrometry. J. Food Prot. 2019, 82, 331–338. [Google Scholar] [CrossRef] [PubMed]
  84. Garcia, M.D.G.; Galera, M.M.; Uclés, S.; Lozano, A.; Fernández-Alba, A.R. Ultrasound–assisted extraction based on QuEChERS of pesticide residues in honeybees and determination by LC–MS/MS and GC–MS/MS. Anal. Bioanal. Chem. 2018, 410, 5195–5210. [Google Scholar] [CrossRef] [PubMed]
  85. Song, L.; Pan, C.; Yang, J.; Zeng, S.; Han, Y. Dual–layer column filtration cleanup and gas chromatography–tandem mass spectrometry detection for the analysis of 39 pesticide residues in porcine meat. J. Sep. Sci. 2020, 43, 1306–1315. [Google Scholar] [CrossRef]
  86. Han, L.; Sapozhnikova, Y.; Nuñez, A. Analysis and occurrence of organophosphate esters in meats and fish consumed in the United States. J. Agric. Food Chem. 2019, 67, 12652–12662. [Google Scholar] [CrossRef] [PubMed]
  87. Ma, G.; Zhang, M.; Zhu, L.; Chen, H.; Liu, X.; Lu, C. Facile synthesis of amine–functional reduced graphene oxides as modified quick, easy, cheap, effective, rugged and safe adsorbent for multi–pesticide residues analysis of tea. J. Chromatogr. A 2018, 1531, 22–31. [Google Scholar] [CrossRef] [PubMed]
  88. Baltussen, E.; Sandra, P.; David, F.; Cramers, C. Stir bar sorptive extraction (SBSE), a novel extraction technique for aqueous samples: Theory and principles. J. Microcolumn Sep. 1999, 11, 737–747. [Google Scholar] [CrossRef]
  89. Hao, W.; ‘Skip’ Kingston, H.M.; Dillard, A.; Stuff, J.; Pamuku, M. Quantification of persistent organic pollutants in dietary supplements using stir bar sorptive extraction coupled with GC–MS/MS and isotope dilution mass spectrometry. Food Addit. Contam. A 2020, 37, 1202–1215. [Google Scholar] [CrossRef] [PubMed]
  90. Li, J.; Shan, J.; Kong, Z.; Fan, C.; Zhang, Z.; Fan, B. Determining multi–pesticide residues in teas by dispersive solid–phase extraction combined with speed–regulated directly suspended droplet microextraction followed by gas chromatography–tandem mass spectrometry. J. Sep. Sci. 2020, 43, 486–495. [Google Scholar] [CrossRef] [PubMed]
  91. Barreira, L.M.F.; Duporté, G.; Rönkkö, T.; Parshintsev, J.; Hartonen, K.; Hyrsky, L.; Heikkinen, E.; Jussila, M.; Kulmala, M.; Riekkola, M.-L. Field measurements of biogenic volatile organic compounds in the atmosphere using solid–phase microextraction Arrow. Atmos. Meas Tech. 2018, 11, 881–893. [Google Scholar] [CrossRef] [Green Version]
  92. Khademi, S.M.S.; Salemi, A.; Jochmann, M.; Joksimoski, S.; Telgheder, U. Development and comparison of direct immersion solid phase micro extraction Arrow–GC–MS for the determination of selected pesticides in water. Microchem. J. 2021, 164, 106006. [Google Scholar] [CrossRef]
  93. Abdallah, O.I. Simultaneous determination of nine dinitroaniline herbicides in environmental samples using a validated vortex–assisted dispersive liquid–liquid microextraction procedure coupled with GC–MS/MS. Chem. Pap. 2020, 74, 2311–2326. [Google Scholar] [CrossRef]
  94. Meng, L.; Chen, S.; Zhu, B.; Zhang, J.; Mei, Y.; Cao, J.; Zheng, K. Application of dispersive liquid–liquid microextraction and GC–MS/MS for the determination of GHB in beverages and hair. J. Chromatogr. B 2020, 1144, 122058–122065. [Google Scholar] [CrossRef]
  95. Tan, J.; Lu, X.; Fu, L.; Yang, G.; Chen, J. Quantification of Cl–PAHs and their parent compounds in fish by improved ASE method and stable isotope dilution GC–MS. Ecotox. Environ. Safe. 2019, 186, 109775–109784. [Google Scholar] [CrossRef] [PubMed]
  96. Rozentale, I.; Zacs, D.; Bartkiene, E.; Bartkevics, V. Polycyclic aromatic hydrocarbons in traditionally smoked meat products from the Baltic states. Food Addit. Contam. A 2018, 11, 138–145. [Google Scholar] [CrossRef] [PubMed]
  97. Tavoloni, T.; Stramenga, A.; Stecconi, T.; Siracusa, M.; Bacchiocchi, S.; Piersanti, A. Single sample preparation for brominated flame retardants in fish and shellfish with dual detection: GC–MS/MS (PBDEs) and LC–MS/MS (HBCDs). Anal. Bioanal. Chem. 2020, 412, 397–411. [Google Scholar] [CrossRef] [PubMed]
  98. Rozentale, I.; Zacs, D.; Perkons, I.; Bartkevics, V. A comparison of gas chromatography coupled to tandem quadrupole mass spectrometry and high–resolution sector mass spectrometry for sensitive determination of polycyclic aromatic hydrocarbons (PAHs) in cereal products. Food Chem. 2017, 221, 1291–1297. [Google Scholar] [CrossRef]
  99. Zhao, L.; Szakas, T.; Churley, M.; Lucas, D. Multi–class multi–residue analysis of pesticides in edible oils by gas chromatography–tandem mass spectrometry using liquid–liquid extraction and enhanced matrix removal lipid cartridge cleanup. J. Chromatogr. A 2019, 1584, 1–12. [Google Scholar] [CrossRef] [PubMed]
  100. Zhu, F.; Wu, X.S.; Li, F.; Wang, W.; Ji, W.L.; Huo, Z.L.; Xu, Y. Simultaneous determination of 12 antibacterial drugs in cream disinfection products with EMR–Lipid cleanup using ultra–high–performance liquid chromatography tandem mass spectrometry. Anal. Methods 2019, 11, 4084–4092. [Google Scholar] [CrossRef]
  101. Ji, Y.; Cui, Z.; Li, X.; Wang, Z.; Zhang, J.; Li, A. Simultaneous determination of nine perfluoroalkyl carboxylic acids by a series of amide acetals derivatization and gas chromatography tandem mass spectrometry. J. Chromatogr. A 2020, 1622, 461132–461140. [Google Scholar] [CrossRef]
  102. Choi, S.J.; Jung, M.Y. Simple and fast sample preparation followed by gas chromatography–tandem mass spectrometry (GC–MS/MS) for the analysis of 2– and 4–methylimidazole in cola and dark beer. J. Food Sci. 2017, 82, 1044–1052. [Google Scholar] [CrossRef] [PubMed]
  103. Rahman, M.M.; Abd El-Aty, A.M.; Shim, J.-H. Matrix enhancement effect: A blessing or a curse for gas chromatography? A review. Anal. Chim. Acta 2013, 801, 14–21. [Google Scholar] [CrossRef] [PubMed]
  104. Shah, V.P.; Midha, K.K.; Findlay, J.W.; Hill, H.M.; Hulse, J.D.; McGilveray, I.J.; McKay, G.; Miller, K.J.; Patnaik, R.N.; Powell, M.L.; et al. Bioanalytical method validation—A revisit with a decade of progress. Pharm. Res. 2000, 17, 1551–1557. [Google Scholar] [CrossRef] [PubMed]
  105. Shendy, A.H.; Eltanany, B.M.; Al-Ghobashy, M.A.; Gadalla, S.A.; Mamdouh, W.; Lotfy, H.M. Coupling of GC–MS/MS to principal component analysis for assessment of matrix effect: Efficient determination of ultra–low levels of pesticide residues in some functional foods. Food Anal. Methods 2019, 12, 2870–2885. [Google Scholar] [CrossRef]
  106. Kittlaus, S.; Schimanke, J.; Kempe, G.; Speer, K. Assessment of sample cleanup and matrix effects in the pesticide residue analysis of foods using postcolumn infusion in liquid chromatography–tandem mass spectrometry. J. Chromatogr. A 2011, 1218, 8399–8410. [Google Scholar] [CrossRef]
  107. Akutsu, K.; Yoshimitsu, M.; Kitagawa, Y.; Takatori, S.; Fukui, N.; Osakada, M.; Yamaguchi, S.; Kajimura, K.; Obana, H.; Watanabe, T. Evaluation of the matrix–like effect in multiresidue pesticide analysis by gas chromatography with tandem mass spectrometry. J. Sep. Sci. 2017, 40, 1293–1300. [Google Scholar] [CrossRef] [PubMed]
  108. Albero, B.; Sánchez-Brunete, C.; Miguel, E.; Tadeo, J.L. Application of matrix solid–phase dispersion followed by GC–MS/MS to the analysis of emerging contaminants in vegetables. Food Chem. 2017, 217, 660–667. [Google Scholar] [CrossRef]
  109. Chiesa, L.M.; Nobile, M.; Malandra, R.; Pessina, D.; Panseri, S.; Labella, G.F.; Arioli, F. Food safety traits of mussels and clams: Distribution of PCBs, PBDEs, OCPs, PAHs and PFASs in sample from different areas using HRMS–Orbitrap® and modified QuEChERS extraction followed by GC–MS/MS. Food Addit. Contam. A 2018, 35, 959–971. [Google Scholar] [CrossRef]
  110. Burket, S.R.; Sapozhnikova, Y.; Zheng, J.S.; Chung, S.S.; Brooks, B.W. At the intersection of urbanization, water, and food security: Determination of select contaminants of emerging concern in mussels and oysters from Hong Kong. J. Agric. Food Chem. 2018, 66, 5009–5017. [Google Scholar] [CrossRef] [PubMed]
  111. Fernandes, V.C.; Luts, W.; Delerue-Matos, C.; Domingues, V.F. Improved QuEChERS for analysis of polybrominated diphenyl ethers and novel brominated flame retardants in capsicum cultivars using gas chromatography. J. Agric. Food Chem. 2020, 68, 3260–3266. [Google Scholar] [CrossRef]
  112. Solaesa, A.G.; Fernandes, J.O.; Sanz, M.T.; Benito-Román, Ó.; Cunha, S.C. Green determination of brominated flame retardants and organochloride pollutants in fish oils by vortex assisted liquid–liquid microextraction and gas chromatography–tandem mass spectrometry. Talanta 2019, 195, 251–257. [Google Scholar] [CrossRef] [PubMed]
  113. Chiesa, L.M.; Lin, S.-K.; Ceriani, F.; Panseri, S.; Arioli, F. Levels and distribution of PBDEs and PFASs in pork from different European countries. Food Addit. Contam. A 2018, 35, 2414–2423. [Google Scholar] [CrossRef] [PubMed]
  114. Ye, X.; Shao, H.; Zhou, T.; Xu, J.; Cao, X.; Mo, W. Analysis of organochlorine pesticides in tomatoes using a modified QuEChERS method based on n–doped graphitized carbon coupled with GC–MS/MS. Food Anal. Method 2020, 13, 823–832. [Google Scholar] [CrossRef]
  115. Fan, J.; Jin, Q.; He, H.; Ren, R.; Wang, S. Detection of 20 phthalate esters in different kinds of food packaging materials by GC–MS/MS with five internal standards. J. AOAC Int. 2019, 102, 255–261. [Google Scholar] [CrossRef] [PubMed]
  116. Cui, Y.; Wang, Z.; Cong, J.; Wang, L.P.; Liu, Y.; Wang, X.C.; Xie, J. Determination of polychlorinated biphenyls in fish tissues from shanghai seafood markets using a modified QuEChERS method. Anal. Sci. 2017, 33, 973–977. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  117. Li, W.; Zhang, Z.; Zhang, R.; Jiao, H.; Sun, A.; Shi, X.; Chen, J. Effective removal matrix interferences by a modified QuEChERS based on the molecularly imprinted polymers for determination of 84 polychlorinated biphenyls and organochlorine pesticides in shellfish samples. J. Hazard. Mater. 2020, 384, 121241–121251. [Google Scholar] [CrossRef] [PubMed]
  118. Mahmoud, A.F.; Escrivá, L.; Rodríguez-Carrasco, Y.; Moltó, J.C.; Berrada, H. Determination of trichothecenes in chicken liver using gas chromatography coupled with triple–quadrupole mass spectrometry. LWT 2018, 93, 237–242. [Google Scholar] [CrossRef]
  119. Szternfeld, P.; Marchi, J.; Malysheva, S.V.; Joly, L. Modular Method for the Determination of polycyclic aromatic hydrocarbons in spices and dried herbs by gas chromatography–tandem mass spectrometry. Food Anal. Method 2019, 12, 2383–2391. [Google Scholar] [CrossRef]
  120. Silva, M.C.D.; Oliveira, M.L.G.; Augusti, R.; Faria, A.F. Simultaneous extraction of pesticides and polycyclic aromatic hydrocarbons in brazilian cachaça using a modified QuEChERS method followed by gas chromatography coupled to tandem mass spectrometry quantification. J. Agric. Food Chem. 2019, 67, 399–405. [Google Scholar] [CrossRef] [PubMed]
  121. Nguyen, K.H.; Fromberg, A. Monochloropropanediol and glycidyl esters in infant formula and baby food products on the Danish market: Occurrence and preliminary risk assessment. Food Control. 2020, 110, 106980–106988. [Google Scholar] [CrossRef]
  122. Xian, Y.; Wu, Y.; Dong, H.; Liang, M.; Wang, B.; Wang, L.; Bai, W.; Zeng, X.; Qian, M.; Zhao, X. Ice–bath assisted sodium hydroxide purification coupled with GC–MS/MS analysis for simultaneous quantification of ethyl carbamate and 12 N–nitrosoamines in yellow rice wine and beer. Food Chem. 2019, 300, 125200–125209. [Google Scholar] [CrossRef]
  123. Wang, X.; Zhou, L.; Luo, F.; Zhang, X.; Sun, H.; Yang, M.; Lou, Z.; Chen, Z. 9,10–Anthraquinone deposit in tea plantation might be one of the reasons for contamination in tea. Food Chem. 2018, 244, 254–259. [Google Scholar] [CrossRef]
Table 1. Application of preparation methods for volatile organic compounds (VOCs) analysis in food by GC–MS/MS during the past three years.
Table 1. Application of preparation methods for volatile organic compounds (VOCs) analysis in food by GC–MS/MS during the past three years.
Food GroupsFood MatricesAnalytePreparation MethodLimit of DetectionLimit of QuantitationRecoveriesRSDRef.
Animal origin foodBeerAcetaldehyde, acrolein, ethyl carbamate, formaldehydeHeadspace (HS)-Solid-phase microextraction (SPME)0.03−0.5 μg/L1.0−2.5 μg/L90−105%0.9−12.0%[8]
Fish products6 polycyclic aromatic hydrocarbons (PAHs)Dynamic HS (DHS) extraction0.01–0.60 ng/g/dw13−62%[4]
Grilled meat samples16 PAHsSPME0.02−1.66 ng/L0.07–5.52 ng/L85.1–102.8%2.6–8.5% (intra-day), 4.5–9.4% (inter-day)[9]
Aquatic productsPolychlorinated biphenyls (PCBs)SPME0.07–0.35 ng/L87.1–99.7%3.8–9.7%[10]
Fish samplesSynthetic musk fragrancesSPME Arrow0.5–2.5 ng/g2.5–5 ng/g<23%[12]
Seafood speciesBenzothiazolesSubcritical water extractionSPME1 and 10 ng/g (dw) for hake, 0.5 and 10 ng/g (dw) for salmon5–50 ng/g (dw)2–20%<21%[7]
VegetablesTomatoes, cucumbers and lettuce11 Phthalate estersSPME0.001–0.430 μg/L>95.2%<10.8%[2]
BeveragesTea128 Pesticide multi-residueHS-SPME1–5 μg/kg70–120%<20%[3]
Liquor, beer, wine, vinegar, tinctureParabens, phenolic antioxidants, sulfonamide plasticizer, and flame retardantSPME0.005–0.2 μg/L0.01–0.5 μg/L98–109%0.8–5.4%[13]
Table 2. Application of preparation methods for semi-volatile organic compounds (SVOCs) analysis in food by GC–MS/MS during the last three years.
Table 2. Application of preparation methods for semi-volatile organic compounds (SVOCs) analysis in food by GC–MS/MS during the last three years.
Food groupsFood MatricesAnalytePreparation MethodLimit of DetectionLimit of QuantitationRecoveriesRSDRef.
ExtractionClean-up
Animal origin foodCatfish219 Pesticides and metabolites (178 pesticides and 41 environmental contaminants)QuEChERsSPE<50 ng/g for 90% analytes1–20 ng/g70−120% for 80% analytes<20% for 80% analytes[33]
Fish, shrimp and shellfishOrganochlorine pesticides (OCPs) and polychlorinated
biphenyls (PCBs)
Matrix solid-phase dispersion (MSPD)0.011–0.046 ng/g0.037–0.153 ng/g70–120%<20%[28]
Bivalve shellfish samplesAmide/Dinitroaniline/Substituted Urea HerbicidesQuEChERS0.3–8.88 μg/kg81–109%<8%[51]
Egg (egg white, egg yolk, and whole egg)Dinitolmide residue and its metaboliteASE0.8–2.8 μg/kg3.0–10.0 μg/kg>80%2.96–5.21% (intra-day)
3.94–6.34% (inter-day)
[20]
Chicken eggs80 PesticidesAcetonitrile with 5% formic acidBond elute enhanced matrix removal-lipid0.02−9.725 µg/kg0.066–30.261 µg/kg65.3–124%4.3−24%[32]
Poultry egg (whole egg, albumen and yolk)Spectinomycin and lincomycinASESPE2.3−4.3 μg/kg6.0−9.5 μg/kg80.0−95.7%1.0−3.4%[34]
Meat (chicken, pork, and beef) and fish (catfish and salmon)Organophosphate esters (OPEs)QuEChERSAutomated robotic clean-up0.5−1 ng/g70−120%≤20%[86]
Chicken tissuesDinitolmide and its metaboliteASESPE0.8–2.5 μg/kg2.7–8.0 μg/kg81.96–94.31%1.72–5.37%[35]
Meats and poultry200 pesticides and 65 environmental contaminantsQuEChERSSPE<5 ng/g (for 90% analytes)70–120%≤20%[36]
Raw propolis14 Lipophilic pesticidesn-hexaneSPE0.002–0.020 μg/g61.0–106.8%≤ 16.9%[37]
10 Beeswax samples160 PesticidesAcetonitrile–ethyl acetate (1:3, v/v)<20 μg/kg80−110% for most analytes<8% for most analytes[18]
HoneybeesPesticide residuesUSEQuEChERS5 μg/kg70–120%<20%[84]
122 Honey samples53 Pesticide residuesQuEChERS 0.001–0.01 mg/kg70.0–120.0%≤20%[52]
Organic honeysPOPs, pesticides and antibiotic residuesQuEChERS7.15–9.80 ng/g82–120%<20%[53]
Bovine milk78 Drugs and 238 pesticidesQuEChERS0.1–10 ng/g70–120%<20%[54]
Hen eggs60 PesticidesQuEChERS0.001–0.004 mg/kg<10 µg/kg for 83% analytes70–120%<20%[50]
Porcine meat39 Pesticide residuesQuEChERSRapid multiplug filtration clean-up (m–PFC)0.01 mg/kg except pyrimethanil74–118% except
pyrimethanil
1−16%[85]
Grain, Vegetables and fruits207 Vegetable samples10 New–generation pesticidesMAE1.4–3.6 ng/g80–111%<11%[21]
249 Grain, beans, fruit and vegetables samples365 Pesticide residuesacetonitrileSPE0.0001–0.0414 mg/kg0.0002–0.1367 mg/kg70–120% for 95% analytes<20%[39]
Mangoes113 PesticidesQuEChERS<4 μg/kg<10.0 μg/kg70–120%<20%[55]
Pigeonpea grains79 PesticidesQuEChERS0.53–3.97 µg/kg1.60–10.05 µg/kg70–120%<15%[48]
Apples; mangos; strawberries; cucumbers and tomatoes41 Triazines and pyrethroids residuesQuEChERS0.03 − 10.22  μg/kg[56]
Fruits (apple and grapes) and vegetables (apple, grapes, cauli–flower, cabbage, peas, potato)Cypermethrin, chlorpyrifos, methyl parathion, ethion, captan, malathion, and triazophosQuEChERS0.0011–0.012 μg/kg0.0012–0.035 mg/kg94–99%3.3–8.1%[57]
69 Fruits and vegetables samples203 PesticidesQuEChERS2 μg/kg70–120% in tomato, apple, and orange for 97% compounds. low recoveries for orange<20% except for biphenyl, butylate,
chlozolinate, and pyrifenox; <20% (inter-day) for 97% analytes
[58]
Chinese vegetables and fruitsPyrethroid pesticidesQuEChERS0.3–4.9 μg/kg~10 μg/kg78.8–118.6%<14.8%[59]
Tomato9 Dinitroaniline herbicidesVortex-assisted dispersive liquid–liquid microextraction (VA-DLLME)0.3–3.3 µg/L2–10 µg/kg64.1–87.9%≤15.1% (inter-day)[90]
Tomatoes20 PesticidesQuEChERS2.6–31.3 μg/kg6.9–93.8 μg/kg72.5–119.7%1.17–14.62%[60]
Garlic, onion, and sugar beetTotal ethofumesate residuesSPE0.0005 mg/kg0.01 mg/kg94−113%1.8−5.7%[61]
Vegetables14 PyrethroidsQuEChERS2–10 μg/kg in tea, 2 μg/kg In tomato, pear, and zucchini[62]
16 Common bean samples142 Pesticide residuesQuEChERs20–100 µg/kg70−120% for 61.4% analytes<20% for 61.4% analytes[63]
211 Vegetable samples12 Pesticide residuesQuEChERS0.0005–0.0023 mg/kg0.0009–0.0047 mg/kg74–120% at 0.01 mg/kg, 75–123% mg/kg2–9% at 0.01 mg/kg, 0.5–16 mg/kg[64]
Greenhouse strawberries16 Pesticide residuesQuEChERS0.1–0.8 µg/kg0.3–2.8 µg/kg80.7–117.2 µg/kg0.6–14.6%[65]
Dried fruits38 Multi-class pesticidesQuEChERS0.02–5 µg/kg70–120%<20% for 92% samples[66]
BeveragesTeaPyrethroid insecticidesmagnetic SPE0.0065–0.1017 µg/L<9.7% (intra-day), <11.95% (inter-day)[31]
Chinese liquor and liquor–making raw materials (sorghum and rice hull)124 Pesticide residuesd–SPE0.00003–0.015 mg/kg0.0001–0.05 mg/kg71–121%<16.8% except cyprodinil, di-flufenican and prothioconazole[44]
Tea131 Pesticidesd-SPE0.5–5.0 µg/kg1.5–16.7 µg/kg78.2–113.9%<15.8%[45]
Tea11 Pesticidesd-SPE0.10–2.10 µg/kg0.29–6.20 µg/kg73.4–106.4%1.9–6.6% (within-run precision), 12.1% (between-run precision)[46]
38 Tea samples45 Pesticide residuesd-SPESpeed-regulated directly suspended droplet microextraction (SR-DSDME)0.1−47 µg/kg70–120%<20%[90]
Green tea203 Pesticide residuesQuEChERSSPE0.33–16.67 μg/kg1–50 μg/kg70−120% for most analytes<20% for most analytes[38]
OilSoybean203 PesticidesSPE<0.01 mg/kg70–120%<20%[25]
Edible oilsOrganophosphorus pesticide residues (OPPs)QuEChERS0.16−1.56 ng/g0.61−5.00 ng/g81.1−113.5%<8.2 (intra–day) <13.9% (inter–day)[67]
Edible oilsPesticidesLLEEnhanced matrix removal (EMR)-lipid cartridge1 ng/g70–120%<20%[22]
SugarSugarcaneFipronil and its metabolitesQuEChERS0.0015–0.002 µg/g0.005 µg/g80.7–98.5%1.80–12.81% (intra–day), 1.2–16.5% (inter–day)[68]
Medicinal plantsLycium barbarum (goji)6 Active ingredients of pyrethrinsSPE0.24–2.1 µg/kg0.8–7 µg/kg88.3–111.5%0.4–8.3%[26]
Panax notoginseng (Burk) F.H.Chen rootPesticide residuesQuEChERS0.0015 mg/kg0.005 mg/kg94–125% for quintozene, 84–119% for hexachlorobenzene (HCB)6.2–16.1%[69]
Chenpi133 Pesticide residuesQuEChERS0.005–0.01 mg/kg70–112.2%0.2–14.4%[70]
Notoginseng Radix et Rhizome116 Pesticide residuesQuEChERS0.01–0.05 mg/kg64.3–119.4%<18.3%[71]
Dried Herbs235 PesticidesQuEChERs0.0003–0.0007 mg/kg0.001–0.002 mg/kg70–120%<20%[72]
Herbal species–ready application201 PesticidesQuEChERS≤10 ng/mL70.0–120.0%≤20%[73]
Cardamom243 Pesticide residuesQuEChERS10 mg/kg70.0–120%<20%[74]
CondimentCapsicum annumChlorantraniliproleQuEChERS0.005 mg/kg0.01 mg/kg85–91%<2% (intra-day and inter-day)[76]
Pepper, chili peppers and its sauce product47 Pesticide residuesQuEChERS0.01 mg/kg70–120% (except for pyrimethanil)<17%[77]
Edible fungiEdible mushrooms10 Pyrethroid
insecticides
QuEChERs0.015−1.67 μg/kg0.051−5.57 μg/kg72.8–103.6%<13%[78]
Condiment and medicinal plantsSpices and herbs140 Organic contaminationQuEChERS0.04–5.20 ng/g0.08–17.19 ng/g80–137%<20%[79]
Muti-matricesTea and herbal infusion300 PesticidesQuEChERS0.018–40 µg/kg0.06–135 µg/kg70–120%<20%[80]
Dried herbs and dried fruit236 PesticidesQuEChERS0.001 mg/kg0.005 mg/kg62–125%1–19%[81]
Beverages, ‘pesto’ sauces, meat preparationMethyleugenolQuEChERS0.4 μg/kg1 μg/kg94.29–100.27%<9%[82]
Beef jerky, cod liver oil, candy8 Prohibited flavor compoundsQuEChERS0.005−0.2 μg/kg0.03−0.8 μg/kg80.2–110.6% (beef jerky), 82.3–94.1% (cod liver oil), 83.6–104.1% (candy)[83]
Table 3. Application of preparation methods for emerging risks analysis in food by GC–MS/MS during the past three years.
Table 3. Application of preparation methods for emerging risks analysis in food by GC–MS/MS during the past three years.
Food TypeSampleAnalytePreparation MethodLimit of DetectionLimit of QuantitationRecoveriesRSDRef.
ExtractionClean-up
Animal origin foodFishPersistent organic pollutants (POPs)Accelerated solvent extraction (ASE)0.01–4.44 ng/g70–120%<20%[19]
FishPOPsHexane–acetoneFlorisil and silica gel0.001–0.040 ng/g0.004–0.12 ng/g60–127%≤20%[17]
Sea fishDihydroxylated PBDEsPressurized liquid
extraction (PLE)
Florisil cartridge3.98–38.74 pg/g11.95–116.22 pg/g19–101% in 10 ng, 28–88% in 20 ng, 42–90% in 40 ng[23]
77 Smoked meat productsPAHsExtracted by dichlormethane/hexaneGel permeation chromatography (GPC)0.02–0.03 µg/kg0.06–0.09 µg/kg97–115%3–9% (intra-day), 6–9% (inter-day)[95]
Fish shellfish and muscle of terrestrial animalsPBDEs and hexabromocyclododecanes (HBCDs)QuEChERSGPC10 pg/g, for BDE-206; 100 pg/g for BDE-20972–97%9–22%[96]
5 Kinds of marine products9 Pefluoroalkyl carboxylic acids (PCAs)Alkaline digestionSPE0.04–0.10 ng/g54.72−107.29%1.53−11.89%[100]
233 Fish and aquatic invertebrate samples6 Polychlorinated biphenyls (PCBs)QuEChERs3–13 ng/g9–40 ng/g75–113%2–12%[116]
Mussels and clamsPCBs, Polybrominated diphenyl ethers (PBDEs), organochlorine pesticides (OCPs), PAHs, and perfluoroalkyl substances (PFASs)QuEChERS0.5–5 ng/g70–120%<20%[108]
Marine bivalves211 Analytes, including pesticides, PCBs, PAHs, PBDEs, and other flame retardantsQuEChERS0.2–10 μg/kg80–120%[109]
Shellfish samples84 PCBs and OCPsQuEChERS0.004–2.705 μg/kg0.01–9.02 μg/kg70–120%<10%[117]
PorkPBDEs and PFASsQuEChERS5–50 pg/g15–150 pg/g80–119%6–19% (intra-day), 9–20% (inter-day)[112]
Plastic packaged baby food samplesBisphenols (BPs)Liquid extractionDispersive sorbents0.1–1 ng/g0.5–4 ng/g91–110%<13%[114]
Chicken meat and edible offal8 TrichothecenesQuEChERS0.05–0.15 μg/g0.25–0.75 μg/kg85.1–108.4%<8%[118]
60 Infant formula and baby foodproductsMonochloropropanediol (MCPDEs) and glycidol (GEs)SPE0.1−0.6 µg/kg1, 2, and 1.2 μg/kg in baby food; 1.2, 1, and 0.5 μg/kg in infant formula91−106% for baby food; 94−99% for infant formula1.2−7.8%[121]
Milk and milk powderSodium fluoroacetate (1080)SPE0.0013–0.0025 µg/kg0.0042–0.0085 µg/kg90–105%<6%[40]
MilkHexamethylenetetramine (HMT)Magnetic molecularly imprinted polymers0.3 μg/kg1.0 μg/kg88.7–111.4%2.6–5.2% (intra-day), 3.6–11.5 (inter-day)[30]
Grains,
vegetables and fruits
Wheat flour samplesBifenox, dichlobenil and diclofop methylMSPE0.39 ng/g (DCB), 0.24 ng/g (BFO), 0.68 ng/g (DCM)1.33 ng/g (DCB), 0.76 ng/g (BFO), 2.18 ng/g (DCM)88.8–96.6%<3.5%[27]
Cereal productsSum of BaP, benz[a]anthracene (BaA), benzo[b]fluoranthene (BbF), and chrysene (Chr)Extracted dichlormethane/hexane (1:1, v/v)GPC0.002–0.006 μg/kg0.07–0.75 μg/kg92–103%4–19%[97]
12 Commercially available plant extract-based dietary supplement samples21 POPsStir-bar sorptive extraction (SBSE)0.00899−0.0931 ng/g4.48−12.9%[89]
Vegetables17 Emerging contaminantsUltrasound-assisted matrix solid-phase dispersion (UAE-MSPD)0.1–0.4 ng/g0.1–0.8 ng/g55–138%≤13% (intra-day), ≤16% (inter-day)[47]
Tomatoes20 Organochlorine pesticidesQuEChERS0.001–0.1 μg/kg0.01–0.33 µg/kg71.2–95.3%<20%[113]
Carrots, turnips and potatoesBisphenol A, its chlorinated derivatives and structural analoguesFocused
ultrasound solid–liquid extraction
Dispersive solid-phase extraction (d-SPE)0.02–0.33 ng/g/dw0.05–1 ng/g/dw74–105 %<12%[115]
CondimentCapsicum cultivars12 Brominated flame retardants (BFRs)QuEChERS1.4–9.3 μg/kg4.6–30.9 µg/kg66–104%<20%[110]
OilsEdible OilsPAHsd-SPE0.06–0.21 μg/kg0.19–0.71 μg kg−198–108%2–5% (intra-day), 4–6% (inter-day)[49]
Edible OilsPAHsd-SPE0.06–0.21 μg/kg0.19–0.71 μg/kg98–108%2–5% (intra-day), 4–6% (inter-day)[107]
Fish oilsBFRs and organochloride pollutantsVortex assisted liquid–liquid microextraction (VALLME) technique0.2–0.7 ng/g76–90%<20%[111]
Beverages30 Tea samples38 PCBsd-SPE0.1–2.9 μg/kg2.0–10 μg/kg73−113%5−20%[49]
54 BeveragesGamma-hydroxybutyrate (GHB)Dispersive liquid–liquid microextraction (DLLME)0.5 ng/mL78.2−84.7%4.9−5.7%[93]
Beverage samples15 PAEsSPE0.005–2.748 µg/L0.018–9.151 µg/L79.3–121.8%<8.8% (intra–day),<9.9% (inter-day)[15]
Brazilian Cachaça93 Pesticides and 6 PAHsQuEChERS2.5 µg/L10.0 µg/L86.7–118.2%≤20%[120]
Yellow rice wineEthyl carbamate (EC) and N-nitrosoamines (NAs)Ice bath-assisted sodium hydroxide purification0.1–0.5 μg/kg0.5–1.5 μg/kg81.5–121%2.2–9.4% (intra-day), 1.6–7.9% (inter-day)[122]
Tea9,10-Anthraquinone (AQ)Solvent extraction10 μg/kg (tea shoots, tea), 0.4 μg/L (tea brew)0.01 mg/kg (tea shoots, tea), 0.4 mg/kg (tea brew)87.0–110.8%2.3–14.6%[123]
Medicinal plantsGinseng5 Organochlorine pesticideLLE0.02–0.12 µg/L in liquid samples, 0.001–0.004 μg/kg in solid samples70.3–85.6% in liquid samples, 83.4–106.9% in solid samples[14]
Muti-matricesSpices and dried herbsPAHsSPE0.25 μg/kg0.5 μg/kgclose to
100%
<22%[119]
Cow milk, plastic bottled beverage, and edible oilPAEsSPE0.15–1.64 ng/g73.7–98.1%1.7–10.2%.[41]
Food Packaging MaterialsFood Packaging Materials20 PAEsSolvent extraction1.7–62.5 µg/kg5.5–208.3 µg/kg82.1–110.8%0.3–9.7%.[42]
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Xu, M.-L.; Gao, Y.; Wang, X.; Han, X.X.; Zhao, B. Comprehensive Strategy for Sample Preparation for the Analysis of Food Contaminants and Residues by GC–MS/MS: A Review of Recent Research Trends. Foods 2021, 10, 2473. https://doi.org/10.3390/foods10102473

AMA Style

Xu M-L, Gao Y, Wang X, Han XX, Zhao B. Comprehensive Strategy for Sample Preparation for the Analysis of Food Contaminants and Residues by GC–MS/MS: A Review of Recent Research Trends. Foods. 2021; 10(10):2473. https://doi.org/10.3390/foods10102473

Chicago/Turabian Style

Xu, Meng-Lei, Yu Gao, Xiao Wang, Xiao Xia Han, and Bing Zhao. 2021. "Comprehensive Strategy for Sample Preparation for the Analysis of Food Contaminants and Residues by GC–MS/MS: A Review of Recent Research Trends" Foods 10, no. 10: 2473. https://doi.org/10.3390/foods10102473

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