Next Article in Journal
Untargeted Metabolomic Approach to Determine the Regulatory Pathways on Salicylic Acid-Mediated Stress Response in Aphanamixis polystachya Seedlings
Next Article in Special Issue
A Simple and Reliable Dispersive Liquid-Liquid Microextraction with Smartphone-Based Digital Images for Determination of Carbaryl Residues in Andrographis paniculata Herbal Medicines Using Simple Peroxidase Extract from Senna siamea Lam. Bark
Previous Article in Journal
Direct Quantitation of Phytocannabinoids by One-Dimensional 1H qNMR and Two-Dimensional 1H-1H COSY qNMR in Complex Natural Mixtures
Previous Article in Special Issue
Measuring Bismuth Oxide Particle Size and Morphology in Film-Coated Tablets
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Green Extraction Techniques as Advanced Sample Preparation Approaches in Biological, Food, and Environmental Matrices: A Review

by
José S. Câmara
1,2,*,
Rosa Perestrelo
1,
Cristina V. Berenguer
1,
Carolina F. P. Andrade
1,
Telma M. Gomes
1,
Basit Olayanju
3,
Abuzar Kabir
3,4,
Cristina M. R. Rocha
5,6,
José António Teixeira
5,6 and
Jorge A. M. Pereira
1,*
1
CQM—Centro de Química da Madeira, Natural Products Research Group, Universidade da Madeira, Campus Universitário da Penteada, 9020-105 Funchal, Portugal
2
Departamento de Química, Faculdade de Ciências Exatas e Engenharia, Universidade da Madeira, Campus da Penteada, 9020-105 Funchal, Portugal
3
Department of Chemistry and Biochemistry, Florida International University, Miami, FL 33199, USA
4
Department of Pharmacy, Faculty of Allied Health Science, Daffodil International University, Dhaka 1207, Bangladesh
5
CEB—Centre of Biological Engineering, Universidade do Minho, Campus de Gualtar, 4710-057 Braga, Portugal
6
LABBELS–Associate Laboratory, Universidade do Minho, Campus de Gualtar, 4710-057 Braga, Portugal
*
Authors to whom correspondence should be addressed.
Molecules 2022, 27(9), 2953; https://doi.org/10.3390/molecules27092953
Submission received: 14 April 2022 / Revised: 28 April 2022 / Accepted: 2 May 2022 / Published: 6 May 2022

Abstract

:
Green extraction techniques (GreETs) emerged in the last decade as greener and sustainable alternatives to classical sample preparation procedures aiming to improve the selectivity and sensitivity of analytical methods, simultaneously reducing the deleterious side effects of classical extraction techniques (CETs) for both the operator and the environment. The implementation of improved processes that overcome the main constraints of classical methods in terms of efficiency and ability to minimize or eliminate the use and generation of harmful substances will promote more efficient use of energy and resources in close association with the principles supporting the concept of green chemistry. The current review aims to update the state of the art of some cutting-edge GreETs developed and implemented in recent years focusing on the improvement of the main analytical features, practical aspects, and relevant applications in the biological, food, and environmental fields. Approaches to improve and accelerate the extraction efficiency and to lower solvent consumption, including sorbent-based techniques, such as solid-phase microextraction (SPME) and fabric-phase sorbent extraction (FPSE), and solvent-based techniques (μQuEChERS; micro quick, easy, cheap, effective, rugged, and safe), ultrasound-assisted extraction (UAE), and microwave-assisted extraction (MAE), in addition to supercritical fluid extraction (SFE) and pressurized solvent extraction (PSE), are highlighted.

1. Introduction

Over the last decades of the last century, technological improvements in chromatographic instruments boosted a remarkable evolution in the analytical chemistry field. Sophisticated configurations hyphenating fast and efficient chromatographic separations with powerful detection systems able to deliver unprecedented time of analysis and analytical performance become the forefront of this revolution where sample preparation was forgotten. For another decade, conventional processes, often involving large volumes of sample and organic solvents and laborious and many times cumbersome protocols prone to originate many experimental errors, continued to be used as standard procedures. Meanwhile, growing concerns with the environmental footprint and planet sustainability are promoting a green agenda affecting the most diverse human activities. The application of the green chemistry principles to analytical chemistry has been elegantly defined under the SIGNIFICANCE acronym [1]. Accordingly, the green analytical chemistry (GAC) envisages the simplest experimental layout involving minimal or no sample preparation and maximum integration of the analytical instruments used, preferentially in an automated way to limit operator intervention, energy consumption, and waste production. In this context, miniaturization of the sample extraction procedure, therefore decreasing sample and solvent requirements, as well as, wastes produced, was an obvious consequence of the GAC principles. This trend has fostered the development of a myriad of microextraction approaches, hereby considered green extraction (GreETs) approaches. These GreETs span almost all, if not all, fields of application, covering the microextraction of selected analytes from biological samples to food matrices or environmental matrices.
On this basis, this review will provide an updated overview of the most important and used green extraction approaches reported in the literature since 2016, their principles, advantages, limitations, and examples of application. Sorbent-based techniques, such as solid-phase microextraction (SPME), stir-bar sorptive extraction (SBSE), fabric-phase sorbent extraction (FPSE), and solvent-based techniques, including µQuEChERS (micro quick, easy, cheap, effective, rugged, and safe), single-drop microextraction (SDME), hollow-fiber liquid-phase microextraction (HF-LPME), and dispersive liquid–liquid microextraction (DLLME), have been considered. Additionally, the use of emerging green solvents such as ionic liquids (ILs) and deep eutectic solvents (DES) as an alternative to conventional solvents will be discussed. Finally, a brief overview of other promising green and sustainable approaches, such as pulsed electric-field-assisted extraction (PEFAE), supercritical fluid extraction (SFE), and subcritical water extraction (SWE), will also be provided.

2. Sample Preparation: A Key Step to Getting the Correct Data

There has been an unprecedented growth in measurement techniques over the last few decades. Instrumentation, such as spectroscopy, microscopy, and chromatography, as well as microdevices and sensors, have undergone phenomenal developments. In contrast, the importance of sample treatment in the analytical layout seems to have been neglected. However, in the last decade, especially driven by the need of pharmaceutical and environmental industries, an exponential growth, and a rapid evolution in this industry, was observed. Some important steps in analytical chemistry to allow accurate, efficient, and fast determinations are commonly used in the sample preparation process, including, for example, extraction (recovering analytes from samples), clean-up (removal of compounds that can interfere with analysis), and solvent evaporation/concentration (concentration of analytes using an N2 stream), are shown in Figure 1. The procedures depend on the sample, the matrix, and the concentration level at which the analysis needs to be carried out.
Sample preparation is the source of about 30% of the experimental errors and of about 60% of the time spent on tasks in the analytical lab. For these reasons, independently of the high performance of the analytical instrument, the sampling procedure and the sample handling and pretreatment methodologies, following a carefully outlined process, are of utmost importance to acquire high-quality analytical results with high selectivity and low sensitivity limits and to ensure high accuracy and reproducibility. In addition, the selective isolation of the analytes of interest and the removal of interfering sample components are vital for eliminating the interferences and matrix effect and protecting the instrumental equipment from possible damages. However, as referred, these procedures were not always seen as key steps in the analytical process, and for that fact, the methodology followed in sample preparation did not receive the same attention as the analytical instrumentation, considered, until the last years, being the bottleneck of the whole analytical procedure. Indeed, the most widely used and commonly accepted classical extraction techniques (CETs) were liquid–liquid extraction (LLE), Soxhlet extraction, and solid-phase extraction (SPE). CETs, however, tend to be slow and labor intensive and use high amounts of hazardous organic solvents causing serious environmental concerns and present low extraction efficiency. Despite this reality, sample preparation techniques did not receive much attention until quite recently. In the last decades, to overcome the drawbacks of CETs, several novel microextraction techniques (Figure 1), which offer faster, cheaper, and “greener” pretreatment of complex samples; utilization of hazardous reagents; and less solvents with generation of less waste, maximizing the safety for operators and the environment, have been reported as alternatives to CETs.
These techniques, hereby designated as green extraction techniques (GreETs), exhibit attractive characteristics, such as simplicity, versatility, high extraction efficiency, and environmentally friendly profile, and have experienced increased development and implementation and stimulated significant progress in laboratory sample treatment. Some of them, due to their importance and growing application in the biological, food, and environmental fields, are highlighted below.

3. Green Extraction (GreETs) Techniques

The most important and used GreET meeting all green analytical chemistry (GAC) requirements, based on miniaturized SPE techniques, such as microextraction in packed syringe (MEPS), solid-phase microextraction (SPME) in direct (DI) and headspace modes (HS), stir-bar sorptive extraction (SBSE), and matrix solid-phase dispersion, in addition to liquid-phase extraction techniques, including single-drop microextraction (SDME), hollow-fiber liquid-phase microextraction (HF-LPME), dispersive liquid–liquid microextraction (DLLME), QuEChERS, solidification of floating organic drop microextraction (SFOME), and ultrasound-assisted back extraction (UABE), will be given more emphasis.

3.1. Miniaturized Sorbent-Based Techniques

SPE is one of the most used conventional extraction and preconcentration methods for the analysis of food, biological, and environmental samples [2,3]. However, this technique requires relatively large amounts of organic solvents and additional clean-up steps, which limits the automation, decreases sample throughput, and potentiates the contamination of the extracts [3]. Moreover, SPE uses large sample amounts, involving long extraction times [2,4]. Recently, new extraction methods were developed using modern techniques with less or no organic solvents to minimize environmental pollution and overcome the limitations of the conventional methods (Figure 2).

3.1.1. Fabric-Phase Sorbent Extraction

Introduced in 2014 by Kabir and Furton [5], the solid-phase dynamic extraction (SPDE) format fabric-phase sorptive extraction (FPSE) is a fast, efficient, and versatile sample preparation approach by implementing a natural or synthetic permeable and flexible fabric (e.g., polyester, fiberglass, or cellulose) substrate to host a chemically coated sol–gel organic–inorganic hybrid sorbent in the form of an ultrathin coating. FPSE allows direct extraction of analytes without sample modification, thus minimizing/eliminating the sample pretreatment steps, which are considered the primary source of major analyte loss [6]. A strong covalent interaction between the fabric substrate and sol–gel contributes to improving the extraction efficiency medium, helping expose the FPSE to extreme chemical conditions without compromising the chemical/structural integrity of the microextraction device. The main disadvantages of FPSE are low sample capacity and extensive longer sample preparation time [7].

3.1.2. Solid-Phase Extraction-Based Approaches

Solid-Phase Microextraction (SPME)

A key milestone in the development of microextraction techniques was first achieved by the seminal invention of solvent-free solid-phase microextraction, popularly known as SPME by Arthur and Pawliszyn in the early 1990s [8]. SPME is an equilibrium-based microextraction technique that involves the partitioning of the analytes from the sample solution into the sorbent coating of the SPME fiber owing to the intermolecular interaction or affinity for the sorbent material. Several configurations of SPME integrally optimize the volume of the extraction phase to improve the high surface-area-to-volume ratio of the extraction-phase coating. There are several geometries for SPME, such as planar, spherical, rod, and in-tube or cylinder [9]. The selection of the SPME geometry depends on the target analyte and matrix that will be analyzed. Usually, in the reduced diameter or length of the extraction phase, its higher surface-area-to-volume ratio can result in a smaller extraction period and higher recoveries [9].
SPME is a simple, fast, universal, sensitive, solventless, and economical technique for the preconcentration and sampling of analytes derived from various types of samples [3,10]. This technique combines extraction, enrichment, and sample injection into a single step. Other advantages of SPME are due to the reliability, sensitivity, and selectivity of this technique [4]. SPME allows the detection of semivolatile and nonvolatile compounds [10] and benefits from the constant development of new sorption coatings [3]. This procedure can be performed in different modes: (i) headspace SPME (HS-SPME) mode in which the analytes are adsorbed/absorbed from the gas phase in equilibrium with the samples (as the temperature is a parameter with a significant effect on the kinetics of the process, this is the most adequate for volatiles extraction); (ii) direct mode (DI) in which the SPME fiber is immersed directly into the bulk sample. In this case, the agitation is an important experimental parameter to facilitate the transport of the analytes from the solution to the fiber. (iii) In the third mode, membrane extraction, the extraction of less volatile compounds is facilitated by the use of a protected membrane.

Microextraction in Packed Sorbent (MEPS)

MEPS emerged as a greener alternative to the conventional SPE, consisting of a sample pretreatment technique based on the miniaturization of SPE. This technique uses the same sorbents as SPE but is considered more advantageous since sorbent integration into a liquid-handling syringe results in low void volumes, making sample manipulations easy [4]. MEPS can be applied to smaller samples and requires shorter sample preparation times and lower solvent volumes [2]. Moreover, MEPS can be performed online in a fully automated way using the same syringe for sample extraction and extract injection into the analytical instrument [4]. A typical MEPS application comprises sorbent conditioning, sample loading, washing, and analyte elution. Contrary to SPE, the two-direction flow potential in MEPS provides the duplication of each step and satisfactory sorbent conditioning, enhanced sample–sorbent interaction, sample loading, and improved analyte elution. The elution and washing steps can be performed with 20–50 µL of organic solvent, and 1–4 mg of reused sorbent material is sufficient to extract a target analyte with high efficiency [11]. More recently, µSPEed has been proposed. It represents an advance with respect to MEPS because it has a unidirectional valve that corresponds to a flow in one direction, in addition to the high pressure conferred by the small diameter of the sorbent particles. In this case, the analytes retained in the adsorbent are not altered by the aspiration of solvents, as is the case with MEPS, which results in more efficient extractions. The most remarkable improvement over MEPS procedure is the direct flow through the sorbent bed; therefore, the analytes retained in the sorbent bed are not disturbed by the solvent aspiration as it occurs with MEPS. Moreover, the high pressure and the single direction contribute to obtain more efficient extractions of the target analytes. Figure 3 represents a schematic overview and the most important aspects of µSPEed.

Solid-Phase Dispersion Extraction (SPDE)

In SPDE, the microparticles are dispersed in the solution (liquid sample) until the equilibrium between the two phases is reached. The most popular formats are the matrix solid-phase dispersion extraction (MSPD), magnetic solid-phase extraction (MSPE), and micro-solid-phase extraction (µSPE).
MSPD is an efficient and generic technique for the isolation of a wide range of drugs, pesticides, naturally occurring constituents, and other compounds from a wide variety of complex plant and animal samples. According to Barker [12], the sample is dispersed over the surface of the bonded-phase support material, producing, through hydrophobic and hydrophilic interactions of the various components, a unique mixed-character phase for conducting target analyte isolation [12].
MSPE is based on the use of sorbent materials, such as magnetic nanoparticles, carbon hemimicelles, and molecular imprinted polymers [13]. C18 functionalized magnetic nanoparticles (MNP) are used for preconcentration or cleanup of moderate and nonpolar polar pesticides due to the absence of internal diffusion resistance, the excellent absorption capacity of the target analytes, and the high surface-to-volume ratio [14]. The main advantage of MSPE is that the sorbent is composed of MNPs, often NPs of the most diverse chemistries and geometries, that can be easily recovered from a solution by a simple spin or centrifugation process.
The pipette-tip SPE is the simplest format of µSPE in which the sorbent is placed in a tip and extraction is handled by using a pipette, widely used in preclinical and clinical development programs in addition to the study of metabolomics, genomics, and proteomics. SPE tips, such as the MonoTip®, NuTip®, and ZipTip®, can be used for the purification of peptides or proteins that, using affinity and metal chelation, can be successfully selectively isolated [15].

3.1.3. Stir-Bar Sorbent Extraction (SBSE)

Stir-bar sorbent extraction (SBSE) was introduced by Baltussen, Sandra, David, and Cramers in 1999 as an alternative to SPME and became one of the most powerful microextraction and preconcentration techniques for the enrichment of volatile analytes from aqueous samples due to its simplicity, robustness, cost-effectiveness, and environmental friendliness. After that, its applications have been extended to the analysis of nonvolatile analytes and solid and liquid samples.
This extraction procedure is based on a magnetic stir (0.5–1 mm thickness) coated with polydimethylsiloxane, a nonpolymeric phase used as an extraction phase of target analytes through hydrophilic interaction. SBSE consists of two steps: extraction and desorption. Related to extraction, the coated stir bar can act in immersion mode (immersed in the sample solution) or in headspace mode (stir bar is exposed in the gas phase above the liquid or solid sample). After extraction, the target analytes adsorbed in the stir bar are desorbed by thermal desorption, followed by analysis in a chromatographic system (e.g., GC, HPLC, and CE) [11,16].
Despite that the SBSE principle is similar to SPME, SBSE exhibits higher sensitivity, recovery, and extraction efficiency. This is due to the larger amount of coated phase in SBSE, which is 50–250 times higher than the SPME fiber, making SBSE more suitable to analyze trace levels in complex matrices. On the other hand, a special interface is required for thermal desorption in gas chromatography (GC), and lower recoveries are obtained for target analytes with a logarithm of octanol–water partitioning coefficient (log Ko/w) lower than 3 [11,16].

3.1.4. Multisphere Adsorptive Microextraction (MSAμE)

A new adsorptive microextraction (AμE) technique was proposed by Nogueira et al. [17,18], which represents a great alternative for the enrichment of a wide range of polar analytes at trace levels in aqueous media, selecting appropriate sorbent phases. The new AμE approach can be used through different analytical devices presenting suitable geometry, where specific sorbents are simply sustained through sticking-based technologies. Usually, the sorbent is physically embedded on the substrate and put in the aqueous media. The solution is stirred using a stir bar or vortexed. Since most of the polar targets are nonvolatile and some of them have thermolabile properties, liquid desorption (LD), followed by HPLC is certainly the following combination of choice for analytical purposes. AμE can appear in two geometrical configurations, namely, bar adsorptive microextraction (BAμE) and multisphere adsorptive microextraction (MSAμE). Nevertheless, previous experimental data [17,18] showed that MSAµE devices present much better stability compared with BAμE, especially when they are exposed to an aggressive sample matrix because, in this case, thermal supporting promotes much higher robustness from the fixation point of view. The MSAμE showed several advantages, namely, high recovery for polar analytes, easiness to prepare, economicalness, and selectiveness, as sorbent can be selected based on the target analyte of interest. Nevertheless, the main drawback is the device’s stability as should be evaluated on a case-by-case basis.

3.2. Miniaturized LPE-Based Techniques

In addition to sorbent- or solid-based GreETs, several microextraction approaches involving a sorbent phase in a liquid state have been developed in the last decades. Similar to the SPE-based techniques, the major shortcomings of the conventional liquid–liquid extraction technique (LLE), such as emulsion formation, long preparation time, noncompliance with GAC due to the usage of a high volume of toxic organic solvents, and inevitability of solvent evaporation and sample reconstitution, have triggered research into the miniaturized and greener version of LLE. In contrast to SPME, miniaturized LPE techniques include solvent-based extraction techniques that use microliters of organic solvent to accomplish the selective isolation, preconcentration of the analytes, and clean-up of the sample (Figure 4).

3.2.1. Single-Drop Microextraction (SDME)

The first to be invented for the series of solvent-based microextraction techniques was SDME. SDME is a nearly solvent-free, quick, inexpensive, and easy-to-operate extraction technique. It can be used to highly enrich analytes in a relatively short time and uses simple laboratory equipment, which considerably lowers the cost of analysis [4,10]. This approach implies that a single drop of an extraction solvent is employed for the isolation of the analyte. It was introduced in the mid-1990s by Liu and Dasgupta [19], which used a drop of water to extract ammonia and sulfur dioxide before the spectrophotometric analysis. SDME is based on the principle of the partitioning of the analytes from the sample solution to the extraction solvent with or without mechanical aid. As presented in Figure 5a, SDME can be operated in different modes: direct immersion is employed mostly for nonvolatile analytes, being the extraction solvent immersed in the liquid sample from which the analytes are transferred, subsequently followed by the withdrawal of the drop before instrumental injection; the headspace mode (HS-SDME) is tailored for the isolation of volatile compounds [20]; the bubble-in-drop (BID-SDME) introduced by Williams et al. [21] was designed to enlarge the droplet surface area; continuous-flow microextraction (CFME) proposed by Liu and Lee [22] was designed to increase the contact area between the analyte and the extraction solvent; or the drop-to-drop liquid–liquid microextraction was developed by Wijethunga et al. [23] in which the sample volume required for analysis is reduced. Automation has also been established with SDME coupled with electrothermal atomic absorption spectrometry for the quantitation of Cr (VI) in natural water samples [24].

3.2.2. Hollow-Fiber Liquid-Phase Microextraction (HF-LPME)

Pedersen-Bjergaard and Rasmussen [25] first developed the hollow fiber LPME coupled with capillary electrophoresis for methamphetamine in biological samples, such as urine and plasma. It is another mode of solvent-based microextraction technique that is premised on the transfer of the target compounds from the sample (donor) solution via a supported liquid membrane to the acceptor phase. Since its introduction, it has gained wide popularity for the analysis of a wide range of analytes in environmental samples [26], biological samples [27,28], and food samples [29,30]. HF-LPME could be operated in two different modes, a two-phase system and a three-phase system (Figure 5b). Although the two modes share a similarity in principle in that they involve the partitioning of the analytes from the sample (donor phase) solution to other phases (acceptor phase), few lines of demarcation can be observed. In two-phase systems, the analytes are transferred from the aqueous phase to the organic acceptor phase based on their affinity for them. In turn, the three-phase mode involves partitioning from the aqueous donor across the organic solid support liquid membranes (SLMs) into the aqueous acceptor phase in the lumen of the hollow fiber [26]. This results in several advantages that accompany automation, including a lower number of operators to recruit, reduced chemical use, and accelerated analysis time, just to mention a few. Automated HF-LPME has been applied for the analysis of pharmaceutical drugs [31].

3.2.3. Dispersive Liquid–Liquid Microextraction (DLLME)

DLLME has also gained wide popularity over the years. DLLME offers several advantages, including small sample volume, high extraction efficiency, low consumption of solvents, high enrichment factor, good repeatability, and high recovery. Furthermore, this technique is simple and uses small amounts of extraction solvents, and the equilibration between the aqueous phase and extracting solvent is fast [32]. Ultrasounds can be applied to disperse the extraction solvent in the sample, avoiding the reduction of the analyte’s partition coefficient between water and the extracting solvent [32]. As a result of efforts to minimize errors incurred in analysis due to intermittent human intervention and to increase the efficiency of the overall process, studies detailing the automation of the LPE techniques have also been reported, for instance, the online sequential injection (SI) DLLME for the isolation and preconcentration of copper and lead using a series of reagents including methanol as disperser solvent mixed with 2.0% (v/v) xylene as the extraction solvent and 0.3% (m/v) ammonium diethyldithiophosphate as the complexing agent. The solvent mixture was merged with the aqueous sample, and 300 μL isobutyl methyl ketone was used to elute the complex of the analytes before the injection into the nebulizer of the flame atomic absorption spectrometry (FAAS) [33]. A similar study involving a modification was reported for the quantitative analysis of cadmium and lead in natural water samples [34].

3.2.4. QuEChERS

At the beginning of this century, Anastassiades, et al. [35] proposed an innovative sample preparation approach with attractive characteristics, a quick, easy, cheap, effective, rugged, and safe (QuEChERS) method for the quantitative measurement of pesticide residues in vegetables and fruits. It is a two-stage process of solid–liquid partitioning with a salting-out effect and a dispersive solid-phase extraction (dSPE). The extraction of the target analytes occurs in the first stage, when a mixture of salts is dispersed in the matrix and mixes thoroughly with an organic phase, often acetonitrile, till an equilibrium is reached. This is followed by a clean-up step (dSPE) using a different combination of porous sorbents and salts according to the matrix interferences that should be removed. Since its invention, QuEChERS has been applied for the analysis of a wide spectrum of analytes in different sample matrices, such as fluoxetine and carbamazepine in benthic invertebrates (Potamopyrgus antipodarum and Valvata piscinalis) [36], pharmaceuticals and personal care products in sewage and surface waters [37], sulfonamide residues in milk samples [38], or BPA in human urine [39]. More recently, several improvements to the original procedure were reported, notably, its miniaturization applied in different fields of research [40] (additional reports available in Tables S1–S3).

3.2.5. Solidification of Floating Organic Drop Microextraction (SFOME)

The SFOME method was introduced by Khalili Zanjani et al. [41] using polycyclic aromatic hydrocarbons as model compounds. In this technique, the collection of the analytes in a microdrop of an organic extraction solvent under agitation is achieved by the solidification of the suspended microdrop organic layer in ice. The solidified microdrop is allowed to melt before it is injected into the instrument for quantitative assessment. The notable characteristic of the extraction solvent, peculiar for this procedure, is its low melting point usually in the range of 10–30 °C. The use of a little amount of organic solvent indicates the compliance of this simple method with the green analytical chemistry requirements (GAC), and it has been popularly employed either individually or in conjunction with other extraction methods for the analysis of contaminants in environmental [42], food [43], and biological [44] samples. The use of SFOME is not limited to the extraction of organic compounds as it has been deployed for the isolation of inorganic metallic ions, such as lead [45].

3.2.6. UABE

Ultrasound-assisted back extraction (UABE) is another LPE-based extraction method that has been used in tandem with other sample preparation strategies, such as cloud point extraction, for the analysis of brominated flame retardants in water samples (BFRs) [46], heterocyclic aromatic amines in natural water samples [47], and DLLME for the isolation of suvorexant in urine samples [48]. The use of UABE was reported to be a necessity when the extracts contain much extraction solvent that is not compatible with the analytical instrument. Zhou, Gao, Zhang, Li, and Li [46] developed a method for the quantitative determination of BFRs in water samples using a cloud point extraction coupled with UABE before injection into the inlet of HPLC–MS/MS. An amount of 400 μL of an aqueous solution of Triton X-114 and 0.5 M ammonium acetate was added to 40 mL of water sample, which was thoroughly mixed and centrifuged at 5000 rpm for 3 min. After the aqueous phase was decanted, 200 μL of acetonitrile and 2 mL of isooctane were added to the surfactant-rich component and sonicated for 5 min. The isooctane layer was allowed to dry with the aid of N2 flow in a new centrifuge tube, while the residue was reconstituted in methanol (50 μL) after which it was injected into the HPLC system. This method gave a limit of detection of 0.3 to 3.0 ng L−1 and a recovery of 8.7%–54.7%.

3.3. Emergent Green Solvents: Ionic Liquids (ILs), DES, and NADES

ILs have evolved as potential replacements for conventional solvents over the years. They are low-melting organic salts with a combination of an organic cation and an organic or inorganic anion, occurring in the liquid state at a temperature below 100 °C. The striking features of ILs, such as negligible vapor pressure, enhanced synthesis route, fewer by-products, thermal stability, and high hydrophobicity, are factors establishing them as green solvents as they are more environmentally friendly than conventional solvents [49]. Added to the advantages of ILs is the ability to modify functionalities, which enhances the selectivity and specificity of the target molecules [50]. ILs have been shown to improve selectivity and extraction efficiency when used in tandem with a metal organic framework (MOF). For instance, [51] employed imidazole-based ILs as a guest material with Zr-MOFs for the preconcentration of sulfonamides in a water sample in a dSPE-HPLC-DAD method. A limit of detection below 0.03 μg L−1 and enrichment factors greater than 270 were obtained. Despite the promising usefulness of ILs in the field of separation science, their notable drawback is that not all ILs are nonvolatile, nonflammable, and stable in air and water as originally considered. In fact, many ILs are volatile, flammable, unstable, and even toxic, particularly to aquatic beings [52]. For this reason, deep eutectic solvents (DES) have emerged as safer alternatives, exhibiting higher stability and lower costs and toxicity [52]. DES are formed by combining different hydrogen bond acceptors and donors, and their classification as ILs is not consensual, mainly because they share more differences than similarities (reviewed in [52]). Among those differences, it is important to highlight in the context of this review that DES are less hazardous and more stable and biodegradable than IL [52]. A specific subclass of DES, composed of components of natural origin, NADES, will be the ultimate green solvent that can be used. For their greener profile and possibilities to fine-tune extraction properties by combining different ILs and, more recently, DES and NADES, the use of these innovative extraction solvents in the most diverse extraction formats is growing exponentially and constitutes one of the forefronts in sample extraction. An exhaustive list of applications using ILs, DES, and NADES is available in Tables S1–S3.

3.4. Other Advanced Extraction Techniques

Other advanced extraction techniques may act at different levels (alone or in combination), including, but not limited to, breaking the overall matrix structure or cell wall, allowing easier penetration of selective solvents with an affinity for the target analytes, fastening mass transfer and extraction kinetics; using more selective and cost-effective extraction solvents, increasing analyte solubility, increasing safety, and decreasing environmental impact (e.g., by changing the type of solvent, decreasing the chemicals needed, reducing energy consumption, or reducing wastes generation). Pulsed electric-field-assisted extraction (PEFAE) is a nonthermal technology that has been primarily applied to disintegrate cells and cellular tissues in food processing and extraction processes. It makes use of very intense electrical pulses with a very short duration. The main objective is to disrupt cell walls and tissues and increase cell permeability without heating the target samples (and analytes), thus increasing extraction efficiency. As the duration of the pulses is designed to avoid thermal effects, it can be used with heat-sensitive compounds. Other electric-field-based approaches are also available. For instance, using moderate electric fields for longer times, combining thermal and nonthermal electric effects, may also be an effective way of extracting the target analytes. This process is also known as ohmic heating. Heat is generated inside the material through the Joule effect, and the matrix is heated almost instantaneously and evenly. Simultaneously, a limited electroporation effect is also expected. It is applicable when heat is needed to achieve an efficient extraction of the target analytes, increasing extraction efficiencies and reducing thermal degradation. Furthermore, energy efficiency in ohmic-heating-based processes is significantly higher than in traditional heating processes [53]. Electric technologies may be used complementary to other more traditional approaches to increase their efficiency and selectivity and decrease the time of sample preparation [54]. Applications of electric fields can be found both in analytes’ extraction and in samples’ concentration/purification. One example, described by Xu and coworkers [55], is the application of electric fields to enhance SPE extraction of contaminating compounds (tricyclic antidepressants) in environmental waters before their identification by GC–MS [55]. Electro-enhanced solid-phase microextraction can be found applied to several other matrices and analytes, including phthalate esters and bisphenol A from blood and seawater [56] or fluoroquinolones in eggs [57]. Electroextraction of analytes across aqueous–organic phase boundaries has also been described, and membrane-based processes coupled with electric fields are also quite common in the literature [58]. Using greener non-petroleum-based solvents and tuning their properties to increase their efficiency and selectivity by using subcritical or supercritical temperatures and pressures is also an interesting approach to reduce or remove the use of toxic and/or environmental impacting solvents in the extraction step. For instance, when considering analytes with moderate polarity and low thermal sensitivity, subcritical water extraction (SWE) may be considered a viable option. Pressurized solvent extraction (PSE) consists of a liquid–liquid extraction technique where the solvent is used at temperatures higher than its “normal” boiling point (at atmospheric pressure) in pressurized systems. The pressure is kept at values above the boiling point at the selected temperature, but under the critical point, allowing to keep the fluid in the liquid state. Higher temperatures allow increasing the solubility of the analytes and the transfer rate, thus improving extraction efficiencies. Further, viscosity decreases, and the high pressures involved may facilitate the solvent penetration into the matrix from which the analytes are being extracted. SWE is a particular case of pressurized liquid extraction, using water as solvent. Water is a solvent with unique properties in these conditions, not present in other solvents. Besides the above-mentioned advantages, the dielectric constant of water decreases when temperature increases, thus increasing its affinity for less polar analytes. Though this decrease is limited, it is possible to use subcritical water as a greener replacer for “intermediate” polarity organic solvents, such as methanol or ethanol. Further, it is possible to tune water properties to meet the desired affinity for the target analyte by selecting the most appropriate temperature. On top of that, the ionization constant of water also increases, thus liberating more ions, H+ and OH, that may work as a catalyst to break down the matrix, thus improving the solvent accessibility to the analyte [59]. On the other side, for thermolabile nonpolar analytes, supercritical carbon dioxide extraction (SCE) is a relevant option, avoiding the use of solvents, such as hexane. Supercritical fluids have mixed properties between liquids and gases, facilitating extraction processes: diffusion, viscosity, and surface tension similar to gases and density and solvation power as liquids. In the particular case of CO2, the critical temperature is close to room temperature (31 ℃), and working under supercritical conditions is possible at relatively low temperatures. The solvent’s low polarity makes it ideal for nonpolar compounds. However, SCE of more polar compounds is possible using a chemical modifier or a cosolvent (such as ethanol), though decreasing the process greenness. Further evaporation or concentration steps are not needed: resuming the atmospheric conditions turns the solvent back into gas that can either return to the environment or be pressurized again to be reused while purifying/concentrating the analyte’s sample.
Greener processes tend to use greener solvents with lower environmental impact. Bio-based solvents using renewable sources or water are considered solvents with a lower environmental footprint. However, the utmost target should be using direct analytical methods not requiring reagents or solvents [60]. When the sample’s pretreatment is unavoidable, alternatives to reduce or use no solvent at all in the preparation step should be considered [61]. In this context, sample treatments such as simple pressing or extrusion, instant controlled pressure drop, PEFs, or microwaves applied directly in the matrix being analyzed, without adding extra solvents, may cause membrane or cellular structure ruptures enough to free intracellular or structural fluids containing the compounds to be analyzed.
In Table 1 are described the advantages and disadvantages of the most common GreETs used in the analysis of biological, food, and environmental samples.

4. High-Resolution Analytical Techniques

Liquid chromatography (LC) has a great benefit on the efficiency of separating complex matrices, but it is not appropriate to achieve structural information of the target analytes. In this sense, LC combined with mass spectrometry (MS) or MS/MS is certainly the most common analytical approach in the analysis of a diversity of target analytes in environmental, clinical, and food matrices, since it provides higher selectivity, mainly when isomeric mixtures were analyzed. HPLC combined with a traditional detector such as ultraviolet (UV) [62], photodiode array detector (PDA) [63], and fluorescence detector (FLD) [64] have been applied in the determination of several target analytes in environmental, clinical, and food matrices. The benefits of these traditional systems are economical, more accessible in common laboratories, efficient, faster, and easy to use. Despite the lower sensitivity attained using these traditional detectors, excellent results related to validation parameters were achieved, namely, low limit of detection (few µg/L), good accuracy (recoveries higher than 70%), and intra- and interday precisions with relative standard deviation (RSD) lower than 20%. Currently, UV, PDA (DAD), and FLD detectors have been substituted by MS and/or MS/MS detectors since they provide high selectivity, sensitivity (low LODs), and ability to provide information related to molecular mass and structural proprieties. Liquid chromatography–tandem mass spectrometry (LC–MS/MS) [65] is becoming a promising analytical approach to analyzing complex matrices due to its high separation resolution, high sensitivity (low LODs), and capacity to identify compounds and does not require any derivatization step before the analysis. Nevertheless, compared with the traditional systems, LC–MS/MS showed several drawbacks, such as the complexity of the operation, expensiveness, and strong matrix effects that promote in many cases signal suppression or enhancement.
The atmospheric pressure ionization of MS has electrospray ionization (ESI) and atmospheric pressure chemical ionization (APCI). The introduction of ESI and APCI overcame the limitation of previous interfaces by evaporating the mobile phase during the ionization process. This, combined with the orthogonal spray interface, provided a means to stop possibly interfering nonvolatile compounds, such as salts, buffers, and detergents, from entering the MS. For many target analytes, ESI provides high sensitivity, being more used than APCI [65]. APCI is more appropriate for the analysis of nonpolar analytes and volatile organic compounds (VOCs). In addition, the major benefits of the application of ESI for quantitative LC–MS are the production of protonated or deprotonated molecules with slight fragmentation, optimal selection of precursor ions, and maximizing sensitivity, the matrix effect being its main drawback.
Single quadrupole, triple quadrupole (QQQ-MS), ion trap (IT), time of flight (TOF), and quadrupole–time of flight (Q–TOF) are the most common MSs used in tandem mass spectrometry (MS/MS). The IT, TOF, and Q–TOF mass spectrometers usually are used for structural elucidation and the identification of unknown compounds. Nevertheless, achieving structural information of unknown compounds requires higher purity of matrices. More sophisticated analytical approaches, such as ultra-performance liquid chromatography–tandem mass spectrometry (UHPLC–MS/MS) [66], have been recently used in the analysis of compounds in environmental, clinical, and food matrices. UHPLC–MS/MS compared with HPLC–MS/MS provides high pressures, narrow peaks, high chromatographic separation, and lower analysis time and solvent volumes. Moreover, LC coupled to high-resolution mass spectrometry (HRMS) was used for direct determination of glyphosate and its metabolite aminomethylphosphonic acid (AMPA) in human urine by combining cold-induced phase separation (CIPS) with hydrophilic pipette tip solid-phase extraction (PT-SPE) [67]. LC–HRMS compared with LC–MS offers screening for targeted, suspect, and nontargeted analysis in a single run, producing high-resolution accurate masses, their isotopic patterns, and MS2 spectra included in online databases.
Gas chromatography (GC) coupled with a flame ionization detector (FID) [68], MS [69], or MS/MS [70] has also been used in minor extension, when compared with LC–MS and LC–MS/MS, in the analysis of environmental, clinical, and food matrices. This fact could be explained by the derivatization process required for the analysis of some target analytes in GC analysis to promote the volatility and decrease the polarity of the analytes, as well as the time of analysis.
Other analytical approaches have been used in the analysis of environmental, clinical, and food matrices, such as flame atomic absorption spectrometry (FAAS) [71], inductively coupled plasma (ICP) combined with mass spectrometry (MS) [72], or optical emission spectrometry (OES) [73]. Additional details about these and other examples are available in Tables S1–S3, covering the clinical, food, and environmental fields, respectively.

5. Applications of Green Extraction Techniques to Different Fields

As discussed in the previous sections, the use of GreETs spans a wide range of applications, covering the most diverse type of samples, from biofluids to environmental matrices and all type of foods (an exhaustive list of applications reported in the literature since 2016 is available in the Supplementary Material). In modern analytical layouts and to fulfill GAC requirements, the analysis that follows the sample preparation using GreETs should employ fast and efficient analytical instruments able to acquire huge amounts of data. As a consequence, powerful data processing and statistical analysis procedures will be required to produce consistent results (Figure 6).

5.1. Biological Samples

The application of GreETs to the clinical field has increased consistently since the beginning of the century [74]. This mostly includes body fluid samples containing lower-molecular-mass organic molecules, less than 500 g/mol, comprising drug analytes, metabolites, environmental exposure contaminants, poisons, tissues, and endogenous substances [74]. These biological samples present great complexity and moderate-to-high levels of protein, thus requiring robust sample preparation approaches able to simplify and isolate the target analytes from the matrix [75]. As discussed in more detail in the previous sections, traditional sample preparation methods are not particularly tailored for clinical applications because they are time-consuming and require various steps and extensive clean-up before analysis. In contrast, most GreETs require low sample amounts, very low or no solvent at all, and simple, fast, and user-friendly systems that can be easily automated [75]. These advantages made SPME, µSPE, MEPS, MSPE, just to name a few GreETs, particularly suitable to process biological samples. Moreover, they also allow spanning a wide range of analytes with different properties, such as drugs for clinical and forensic toxicology assays, pharmacokinetic studies, biochemical analysis, pharmaceuticals, in vivo applications, and metabolomics [75]. SPME and its different formats are particularly efficient in this field of application because they often require minimum sample pretreatment and can be easily coupled to analytical instruments (e.g., CG and LC), providing an enhanced extraction capacity and simultaneous quantification of different compounds with overall sensitivity. This includes the simultaneous identification of drugs of abuse (e.g., amphetamines, barbiturates, methadone), psychoactive substances, pharmaceuticals (e.g., antidepressants, antiepileptic agents, steroids, anorectic agents, corticosteroids, anaesthetics), substances that affect the adrenergic system, nonsteroidal anti-inflammatory substances, and so forth [76]. Examples of such applications are available as Supplementary Material (Table S1). Among the different biological matrices, microextraction of urine samples has the advantage of minimum processing, often not requiring any centrifugation or filtration before extraction. This minimizes sample handling and improves method precision. Additionally, it is suitable for a wide range of sample volumes, including volumes as small as 50 μL, and even for sampling when the volume is not accurately known. Diverse types of GreETs using urine are available in the literature, SPME, µSPE, and MEPS being the most often reported [75] (Table S1). The use of GreETs with blood sampling is also advantageous, particularly when this allows the elimination of blood-withdrawal steps from the analytical workflow, as with SPME. GreET usage also reduces the risk of analyte degradation and matrix changes due to enzymatic conversion, as well as fast sample collection and clean-up. Different examples of applications involving blood sampling using GreETs can be found in the literature, such as VOCs (SPME [77]), polycyclic aromatic hydrocarbons (PAHs, pipette-tip SPE [78]), Ni and Pb (µSPE [79]), opiates (MEPS [70]), and antidepressants (FPSE [80]). SPME has also been reported in in vivo assays with biological matrices like tissues. This can be performed with a removed tissue portion (ex vivo), direct in vivo measurement, exposing the BioSPME needle to the tissue or even inserting the probes directly into the tissue. Regarding this, Musteata [81] observed that microdialysis and SPME were not only appropriate for tissue sampling but also complementary to each other for in vivo sampling and ex situ analysis. By using this approach, the probe extracts only a slight fraction of the free analyte, minimizing disturbances of chemical equilibrium and allowing multiple measurements of analyte concentrations under physiological conditions. Moreover, the accurate determination of analyte concentration is unaffected by the sample volume. Finally, the technique is open to miniaturization, allowing its application within small living systems, sample storage and transportation, and easy coupling to portable instrumentation [75]. An example of such an approach was reported by Cudjoe, et al. [82], which used SPME to monitor neurotransmitter changes in the striatum of a rat brain after dosing antidepressants, variations in serotonin concentrations due to deep-brain stimulations, and distribution of pharmaceuticals in the striatal region and cortex. This elegant experiment shows that SPME can also be very useful in metabolomics assays, particularly at the initial stage of biomarker discovery in medical diagnosis. It is also very relevant to the quantification of different compounds simultaneously, which enables the simultaneous monitoring of drugs in complex treatments. This is possible because GreETs coupling with chromatographic methods, as shown in Section 4, can be easily achieved, allowing the analysis of a whole pharmacopoeia of drugs, such as anticancer, antibiotic, antidepressant, analgesic, anti-inflammatory, steroid, and neurotransmitter drugs. This can help to provide earlier detection of the disease, which is imperative for a successful clinical treatment, especially in some oncologic diseases, where an early diagnosis is crucial for the survival of the patient without suffering severe impacts on health and life quality. FPSE is a very promising GreET having a key advantage regarding other microextraction approaches, allowing a direct analyte extraction with no sample modification [6]. Since its introduction in 2014, many examples of applications involving biological samples have been reported in the literature, such as the cow and human breast milk sample clean-up for screening bisphenol A and residual dental restorative material [83]; the simultaneous monitoring of inflammatory bowel disease treatment drugs [84] and anticancer drugs [85] in whole blood, plasma, and urine; or the assessment of radiation exposure [86] (Table S1). The use of magnetic nanoparticles as microextraction sorbents in MSPE also results in a very simple and efficient extraction procedure because the sorbent can be tailored to extract specific analytes, and the sorbent-retained analyte complex can be easily recovered from the solution using a magnetic field or magnet [87]. MSPE has been used to extract different drugs from urine, such as nonsteroidal anti-inflammatory drugs (NSAIDs) [88], methadone [89], pseudoephedrine [90], fluoxetine [91], and statins [92], as well as antiepileptic drugs [93] or ibuprofen [94] from plasma (Table S1). GreETs involving liquid-phase sorbents, such as DLLME, are also often reported in the literature. This format, mostly assisted by ultrasounds (UA-DLLME), allows the usage of a myriad of extraction solvents, and consequently, the repertoire of applications is very broad. Mabrouk et al. [95], for instance, used UA-DLLME to extract three gliflozins (antidiabetic drugs) from plasma.

5.2. Food Samples

Food analysis is of great importance since ingestion of a growing number of compounds intentionally or not added to food can represent a risk to our health. However, beyond food safety, consumers are also more aware of the nutritional value of food and are also interested in its composition, particularly regarding the presence of bioactive compounds. For these reasons, efficient methodologies for the identification and quantification of all these analytes are required. Accordingly, GreETs have been used in the sample preparation procedures of different food matrices to extract and preconcentrate target analytes to a sufficient level to allow their analysis [96]. The µSPE technique, for instance, has been used in the determination of aflatoxins [96], pesticides [62], trace metals [73], and pollutants, such as bisphenol A [97] and PAHs [68], in a variety of food products. Additionally, it aided in the identification and quantification of rosmarinic acid in medicinal plants [98] and vitamin D3 in bovine milk [99]. MEPS is another GreET that has been employed in the analysis of foodstuffs, including the identification of herbicides in rice [100], insecticides in drinking water [101], pesticides in apple juice and coffee [102], antibiotics [103] and steroids [104] in milk, parabens in vegetable oil [105], PAHs in apple [106], caffeine in drinks [107], and polyphenols in baby food [108]. SPME has been widely used to study the volatile composition of several foods, including walnut oils [109], hongeo [110], melon [111], and dairy products [112]. Moreover, this technique has also been used to determine the composition of specific analytes, such as the x-ray induced markers 2-dodecylcyclobutanone and 2-tetradecylcyclobutanone in irradiated dairy products [113], the contaminants 1,4-dioxane and 1,2,3-trichloropropane [114], acrylamide [115], organophosphorus pesticides [116], phthalates [117], synthetic phenolic antioxidants [118], and xanthines [119]. MSPD has been reported in the literature for the extraction of flavonoids [120], polyphenols [121], mangiferin, and hyperoside in mango-processing waste [122], ergosterol in edible fungi [123], and pharmacologically active substances in microalgae [124]. This methodology has also been applied for pesticide [125] and sulfonylurea herbicide [126] extraction in several food matrices. MSPE allowed the extraction of trace metals in food products [127] (additional examples available in Table S2). Moreover, studies have shown that this technique can be used for the determination of acrylamide [78,79], bisphenols [80], PAHs [128], plant growth regulators [129], and caffeine [130]. FPSE is another GreET that has been shown to be very useful for the determination of several classes of pesticides in foods [96]. Other analytes studied using this technique include bisphenol A [131], oligomers [132], PAHs [64], steroid hormone residues [133], and tetracycline residues [134]. DLLME has been vastly applied for the determination of trace metals [71] (additional examples available in Table S2), pesticides [96], chloramphenicol [135], and nonsteroidal anti-inflammatory drugs [136] in different foods. μQuEChERS was employed in the extraction of several analytes from foods, ranging from pesticide residues in wine [137] and PAHs in coffee and tea [138] to polyphenols in baby food [139] and pyrrolizidine alkaloids in oregano [66]. The application of SDME was proved to allow the determination of unfavorable compounds and elements in foods, such as drug metabolites [140], acrylamide [141], ammonia [142], ethyl carbamate [143], formaldehyde [144], tartrazine [145], and Cu(II) [146]. Similarly to SDME, SFOME can be used for the detection of trace metals [147], as well as of β-lactam antibiotic residues [148] and organochlorine pesticides [149]. PEAE has been applied for the extraction of different bioactive compounds [63], including phenolic compounds [43], carotenoids [150], procyanidins [151], and sulforaphane [152]. The use of SFE has been used for the extraction of several antioxidant and antibacterial compounds from feijoa leaf [153], fatty acids and oils from Indian almonds [154], oleoresins from industrial food waste [155], and polar lipid fraction from blackberry and passion fruits [156]. Additionally, SFE was employed for the extraction of phytochemicals from Terminalia chebula pulp [157]. Finally, SWE is a technique largely applied to the extraction of several classes of bioactive compounds, including anthocyanins [158], fatty acids [159], hesperidin and narirutin [160], phenolic compounds [161], and scopoletin, alizarin, and rutin [162]. The extraction of antioxidant protein hydrolysates from shellfish waste [163] and pectic polysaccharides from apple pomace has been also previously accomplished by SWE [164].

5.3. Environmental Samples

Most environmental samples have complex matrix compositions and involve the determination of trace and ultra-trace analytes [3]. For instance, the determination of PAHs in water samples or pesticide analysis is challenging due to their very low concentrations [2,3]. This requires efficient clean-up and enrichment procedures before the analytes’ analysis [2]. MEPS seems to be tailored for these requirements and has been applied in the analysis of benzene, phenol and their derivates [165], diazinon [166], La3+ and Tb3+ [167], organophosphorus pesticides [168] in water samples, fipronil and fluazuron residues in wastewater [169], and PAHs in the most diverse samples (see Table S3), including Antarctic snow [170], and in the detection of phthalates in tap and river water [171]. SPME is eventually one of the most used sample extraction procedures and has been applied for the detection of different pesticides in water [172] (additional examples in Table S3), microplastic in coral reef invertebrates [173], PAHs in rainwater [174], and volatile organic compounds (VOCs) in wastewater [69]. Molecularly imprinted polymers (MIPs) have also been employed in the extraction of polychlorinated aromatic compounds from environmental samples. Some applications include the use of MIPs in the analysis of 2-chlorophenol [175], 2,4-dichlorophenoxyacetic acid [176], and endosulfans [177] in water samples and in the determination of organochlorine pesticides in environmental samples [178]. This methodology has also been reported in the preparation of soil samples to increase the extraction efficiency of triazine herbicides [179]. Multisphere adsorptive microextraction (MSAμ) has been applied in the extraction of caffeine, acetaminophen [18], pharmaceuticals, sexual steroid hormones, and antibiotics [17] in water samples. QuEChERS is known as the Swiss knife of extraction. Its µQuEChERS version is even more greener and includes applications such as the detection of insecticides in guttation fluids [65], pesticides in arthropods and gastropods [180], and VOCs in zebrafish [181].
LPME techniques, such as SDME and SLLME, use small volumes of organic solvents to extract the analytes [4]. SDME has gained a lot of interest in the last few years and is mostly used for the determination of trace analysis in environmental matrices, including Cu(II) in tap and seawater [146], PAHs in tap water [182], ranitidine in wastewater [183], and V(V) in water samples [184]. DLLME is another efficient microextraction procedure, and its ultrasound-assisted (UA) DLLME variation has been adopted in several environmental matrices for the analysis of aromatic amines [185], Cd [186], Cr [187], dyes [188], herbicides [189], polybrominated biphenyls [190], pyrethroid insecticides [191], and tetracycline [192] in water samples. SFE was applied to environmental matrices for the analysis of Ag in electronic waste [193], petroleum biomarkers in tar balls and crude oils [194], petroleum hydrocarbons in soil [195], and solanesol in tobacco residues [196]. In turn, SWE has been successfully used for the extraction of Co, Li, and Mn in spent lithium-ion batteries [197], crude oil in soil [198], oil shale in mines [199], and VOCs in sewage sludge [200]. An exhaustive list of GreETs involving environmental samples is available in Table S3.

6. Final Remarks

The sample preparation, including the extraction process, is one of the most important steps for any analysis. It is the step that determines the quality of the measurements of the target analytes, and so it should be critically considered. In recent years, a rapid transition from solvent-based extractions and solid sorbent-based extractions to miniaturized formats promoted a substantial reduction of the solvent consumption. Being simpler, faster, more economical, and user-friendly than CETs, SPE techniques have become more popular. In turn, this has boosted a continuous improvement of innovative techniques respecting the green analytical chemistry principles, such as, FPSE, SPME, SPDE, MEPS, SBSE, and MSAμE. However, there are always opportunities for improvement, and future research should be directed to, for instance, innovative sorbents/nanosorbents able to further improve retention efficiency, loading capacity, and selectivity. Also noticeable will be the use of artificial intelligence, including microfluidics and smartphones, to boost the automation of extraction procedures and the use of alternative detectors. Another challenging aspect in this field in the near future is the integration of the whole analytical procedure in injection loops able to protect operators from harmful solvents and significantly reduce their use.

Supplementary Materials

The following supporting information can be downloaded at https://www.mdpi.com/article/10.3390/molecules27092953/s1: Table S1: Representative applications of GreETs for the analysis of biological samples; Table S2: Representative applications of GreETs for the analysis of food samples; Table S3: Representative applications of GreETs for the analysis of environmental samples.

Author Contributions

J.S.C. and J.A.M.P. planned and managed the manuscript; J.A.M.P. was responsible for Section 1; J.S.C. was responsible for Section 2 and Section 6; J.S.C., R.P. and C.V.B. collaborated on Section 3.1; B.O., A.K. and J.A.M.P. collaborated on Section 3.2; C.M.R.R., J.A.T. and R.P. collaborated on Section 3.3; R.P. was responsible for Section 4; T.M.G., C.F.P.A., C.V.B., R.P. and J.A.M.P. collaborated on Section 5; J.S.C. and J.A.M.P. designed all the figures and edited the manuscript; J.S.C. made its submission. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by FCT (Fundação para a Ciência e a Tecnologia) through the CQM Base Fund, UIDB/00674/2020, Programmatic Fund, UIDP/00674/2020, and CEB—Centre of Biological Engineering, and by ARDITI (Agência Regional para o Desenvolvimento da Investigação Tecnologia e Inovação) through the project M1420-01-0145-FEDER-000005, Centro de Química da Madeira (CQM+; Madeira 14-20 Program). The authors also acknowledge FCT and the Madeira 14–20 Program to the Portuguese Mass Spectrometry Network (RNEM) through the PROEQUIPRAM program, M14-20 M1420-01-0145-FEDER-000008). ARDITI is also acknowledged for the postdoctoral fellowship granted to J.A.M.P. (Project M1420-09-5369-FSE-000001).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Not applicable.

Acknowledgments

The authors acknowledge Victoria F. Samanidou from Aristotle University of Thessaloniki for the kind invitation to participate in this Special Issue.

Conflicts of Interest

The authors declare no conflict of interest.

Abbreviations

AMPA: aminomethylphosphonic acid; APCI: atmospheric pressure chemical ionization; BIN: barrel insert needle; BID-SDME: bubble-in-drop; BFR: brominated flame retardants; CET: classical extraction techniques; CFME: continuous-flow microextraction; CIPS: cold-induced phase separation; CNFs: carbon nanofibers; CNTs: carbon nanotubes; DAD: diode-array detection; DCM: dichloromethane; DES: deep eutectic solvents; DLLME: dispersive liquid–liquid microextraction; dSPE: dispersive solid-phase extraction; DVB: divinylbenzene; ESI: electrospray ionization; ECD: electrochemical detection; EPT-SPME: effervescent pipette-tip solid-phase microextraction; EtAc: ethyl acetate; EtOH: ethanol; FA: formic acid; FAAS: flame atomic absorption spectrometry; FID: flame ionization detector; FGO-TD-PTFE: functional graphene oxide thermal desorption poly(tetrafluoroethylene); FLD: fluorescence detector; FPSE: fabric-phase sorbent extraction; GA: gallic acid; GAC: green analytical chemistry; GC–MS: gas chromatography–mass spectrometry; GreETs: green extraction techniques; HF-LPME: hollow fiber liquid-phase microextraction; HPLC: high performance liquid chromatography; HS: headspace; HRMS: high-resolution mass spectrometry; ICP: inductively coupled plasma; ILs: ionic liquids; IT: ion trap; LC–MS: liquid chromatography–mass spectrometry; LDH: layered double hydroxide; LLE: liquid–liquid extraction; LLME: liquid–liquid microextraction; LOD: limit of detection; LPE: liquid-phase extraction; M-ILs-SDME: magnetic ionic liquid single-drop microextraction; MAE: microwave-assisted extraction; MIP: molecular imprinted polymer; MIOMS-ir: molecularly imprinted ordered mesoporous silica imprint-removed silica; MISM: molecularly imprinted silica monolithic; MeOH: methanol; MEPS: microextraction by packed sorbent; MMF: multiple monolithic fiber; MNPs: magnetic nanoparticles; MOF: metal organic framework; MS: mass spectrometry; MSAμ: multisphere adsorptive microextraction; MSPD: matrix solid-phase dispersion; MSPE: magnetic solid-phase extraction; MS/MS: tandem mass spectrometry; MWCNTs: multiwalled CNTs; NADES: natural deep eutectic solvents; NPs: nanoparticles; NTD: needle trap device; NTME: needle trap microextraction; NSAIDs: nonsteroidal anti-inflammatory drugs; OES: optical emission spectrometry; PAHs: polycyclic aromatic hydrocarbons; PDA: photodiode array; PDMS: polydimethylsiloxane; PEFAE: pulsed electric-field-assisted extraction; PSE: pressurized solvent extraction; PT: pipette tip; PS/DVB-RP: reverse-phase polystyrene–divinylbenzene sorbent; QQQ: single quadrupole triple quadrupole; Q–TOF: quadrupole–time of flight; SBSE: stir-bar sorptive extraction; SCE: supercritical carbon dioxide extraction; SDME: single-drop microextraction; SFE: supercritical fluid extraction; SFOME: solidification of floating organic drop microextraction; SPE: solid-phase extraction; SPME: solid-phase microextraction; SWE: subcritical water extraction; TOF: quadrupole–time of flight; UA: ultrasound-assisted; UHPLC: ultrahigh performance liquid chromatography; UABE: ultrasound-assisted back extraction; UV: ultraviolet analysis; VOCs: volatile organic compounds; µQuEChERS: micro-QuEChERS; µSPE: micro-solid-phase extraction.

References

  1. Gałuszka, A.; Migaszewski, Z.; Namieśnik, J. The 12 principles of green analytical chemistry and the SIGNIFICANCE mnemonic of green analytical practices. TrAC Trends Anal. Chem. 2013, 50, 78–84. [Google Scholar] [CrossRef]
  2. Fu, S.; Fan, J.; Hashi, Y.; Chen, Z. Determination of polycyclic aromatic hydrocarbons in water samples using online microextraction by packed sorbent coupled with gas chromatography-mass spectrometry. Talanta 2012, 94, 152–157. [Google Scholar] [CrossRef] [PubMed]
  3. Delińska, K.; Yavir, K.; Kloskowski, A. Ionic liquids in extraction techniques: Determination of pesticides in food and environmental samples. TrAC Trends Anal. Chem. 2021, 143, 116396. [Google Scholar] [CrossRef]
  4. Casado, N.; Ganan, J.; Morante-Zarcero, S.; Sierra, I. New Advanced Materials and Sorbent-Based Microextraction Techniques as Strategies in Sample Preparation to Improve the Determination of Natural Toxins in Food Samples. Molecules 2020, 25, 702. [Google Scholar] [CrossRef] [Green Version]
  5. Kabir, A.; Furton, K.G. Fabric Phase Sorptive Extractors. U.S. Patent 20140274660A1, 18 September 2014. [Google Scholar]
  6. Kabir, A.; Samanidou, V. Fabric Phase Sorptive Extraction: A Paradigm Shift Approach in Analytical and Bioanalytical Sample Preparation. Molecules 2021, 26, 856. [Google Scholar] [CrossRef]
  7. Olia, A.E.A.; Mohadesi, A.; Feizy, J. Ochratoxin Determination in Food Samples by Fabric Phase Sorptive Extraction Coupled with High-Performance Liquid Chromatography Technique. 2022. Available online: https://www.researchsquare.com/article/rs-1478781/v1 (accessed on 1 May 2022).
  8. Arthur, C.L.; Pawliszyn, J. Solid phase microextraction with thermal desorption using fused silica optical fibers. Anal. Chem. 1990, 62, 2145–2148. [Google Scholar] [CrossRef]
  9. Billiard, K.M.; Dershem, A.R.; Gionfriddo, E. Implementing Green Analytical Methodologies Using Solid-Phase Microextraction: A Review. Molecules 2020, 25, 5297. [Google Scholar]
  10. Ribeiro, C.; Ribeiro, A.R.; Maia, A.S.; Goncalves, V.M.; Tiritan, M.E. New trends in sample preparation techniques for environmental analysis. Crit. Rev. Anal. Chem. 2014, 44, 142–185. [Google Scholar] [CrossRef]
  11. Filippou, O.; Bitas, D.; Samanidou, V. Green approaches in sample preparation of bioanalytical samples prior to chromatographic analysis. J. Chromatogr. B Anal. Technol. Biomed. Life Sci. 2017, 1043, 44–62. [Google Scholar] [CrossRef]
  12. Barker, S.A. Matrix solid phase dispersion (MSPD). J. Biochem. Biophys. Methods 2007, 70, 151–162. [Google Scholar] [CrossRef]
  13. Chang, Y.C.; Chen, D.H. Adsorption kinetics and thermodynamics of acid dyes on a carboxymethylated chitosan-conjugated magnetic nano-adsorbent. Macromol. Biosci. 2005, 5, 254–261. [Google Scholar] [CrossRef] [PubMed]
  14. Bruzzoniti, M.C.; Sarzanini, C.; Costantino, G.; Fungi, M. Determination of herbicides by solid phase extraction gas chromatography-mass spectrometry in drinking waters. Anal. Chim. Acta 2006, 578, 241–249. [Google Scholar] [CrossRef] [PubMed]
  15. Hsieh, H.C.; Sheu, C.; Shi, F.K.; Li, D.T. Development of a titanium dioxide nanoparticle pipette-tip for the selective enrichment of phosphorylated peptides. J. Chromatogr. A 2007, 1165, 128–135. [Google Scholar] [CrossRef] [PubMed]
  16. Berton, P.; Lana, N.B.; Ríos, J.M.; García-Reyes, J.F.; Altamirano, J.C. State of the art of environmentally friendly sample preparation approaches for determination of PBDEs and metabolites in environmental and biological samples: A critical review. Anal. Chim. Acta 2016, 905, 24–41. [Google Scholar] [CrossRef] [PubMed]
  17. Neng, N.R.; Silva, A.R.M.; Nogueira, J.M.F. Adsorptive micro-extraction techniques—Novel analytical tools for trace levels of polar solutes in aqueous media. J. Chromatogr. A 2010, 1217, 7303–7310. [Google Scholar] [CrossRef] [PubMed]
  18. Silva, A.R.M.; Neng, N.R.; Nogueira, J.M.F. Multi-Spheres Adsorptive Microextraction (MSAμE)—Application of a Novel Analytical Approach for Monitoring Chemical Anthropogenic Markers in Environmental Water Matrices. Molecules 2019, 24, 931. [Google Scholar] [CrossRef] [Green Version]
  19. Liu, S.; Dasgupta, P.K. Liquid Droplet. A Renewable Gas Sampling Interface. Anal. Chem. 1995, 67, 2042–2049. [Google Scholar] [CrossRef]
  20. Jeannot, M.; Przyjazny, A.; Kokosa, J. Single drop microextraction—Development, applications and future trends. J. Chromatogr. A 2009, 1217, 2326–2336. [Google Scholar] [CrossRef]
  21. Williams, D.B.G.; George, M.J.; Meyer, R.; Marjanovic, L. Bubbles in Solvent Microextraction: The Influence of Intentionally Introduced Bubbles on Extraction Efficiency. Anal. Chem. 2011, 83, 6713–6716. [Google Scholar] [CrossRef]
  22. Liu, W.; Lee, H.K. Continuous-Flow Microextraction Exceeding1000-Fold Concentration of Dilute Analytes. Anal. Chem. 2000, 72, 4462–4467. [Google Scholar] [CrossRef]
  23. Wijethunga, P.A.L.; Nanayakkara, Y.S.; Kunchala, P.; Armstrong, D.W.; Moon, H. On-Chip Drop-to-Drop Liquid Microextraction Coupled with Real-Time Concentration Monitoring Technique. Anal. Chem. 2011, 83, 1658–1664. [Google Scholar] [CrossRef] [PubMed]
  24. Pena, F.; Lavilla, I.; Bendicho, C. Immersed single-drop microextraction interfaced with sequential injection analysis for determination of Cr(VI) in natural waters by electrothermal-atomic absorption spectrometry. Spectrochim. Acta Part B At. Spectrosc. 2008, 63, 498–503. [Google Scholar] [CrossRef]
  25. Pedersen-Bjergaard, S.; Rasmussen, K.E. Liquid− liquid− liquid microextraction for sample preparation of biological fluids prior to capillary electrophoresis. Anal. Chem. 1999, 71, 2650–2656. [Google Scholar] [CrossRef] [PubMed]
  26. Madikizela, L.; Pakade, V.; Ncube, S.; Tutu, H.; Chimuka, L. Application of Hollow Fibre-Liquid Phase Microextraction Technique for Isolation and Pre-Concentration of Pharmaceuticals in Water. Membranes 2020, 10, 311. [Google Scholar] [CrossRef] [PubMed]
  27. Hrdlička, V.; Navrátil, T.; Barek, J. Application of hollow fibre based microextraction for voltammetric determination of vanillylmandelic acid in human urine. J. Electroanal. Chem. 2019, 835, 130–136. [Google Scholar] [CrossRef]
  28. Kazakova, J.; Villar-Navarro, M.; Pérez-Bernal, J.L.; Ramos-Payán, M.; Bello-López, M.Á.; Fernández-Torres, R. Urine and saliva biomonitoring by HF-LPME-LC/MS to assess dinitrophenols exposure. Microchem. J. 2021, 166, 106193. [Google Scholar] [CrossRef]
  29. Dominguez-Tello, A.; Dominguez-Alfaro, A.; Gómez-Ariza, J.L.; Arias-Borrego, A.; García-Barrera, T. Effervescence-assisted spiral hollow-fibre liquid-phase microextraction of trihalomethanes, halonitromethanes, haloacetonitriles, and haloketones in drinking water. J. Hazard. Mater. 2020, 397, 122790. [Google Scholar] [CrossRef]
  30. Piao, H.; Jiang, Y.; Li, X.; Ma, P.; Wang, X.; Song, D.; Sun, Y. Matrix solid-phase dispersion coupled with hollow fiber liquid phase microextraction for determination of triazine herbicides in peanuts. J. Sep. Sci. 2019, 42, 2123–2130. [Google Scholar] [CrossRef]
  31. Tajik, M.; Yamini, Y.; Esrafili, A.; Ebrahimpour, B. Automated hollow fiber microextraction based on two immiscible organic solvents for the extraction of two hormonal drugs. J. Pharm. Biomed. Anal. 2015, 107, 24–31. [Google Scholar] [CrossRef]
  32. Rezaee, M.; Yamini, Y.; Faraji, M. Evolution of dispersive liquid-liquid microextraction method. J. Chromatogr. A 2010, 1217, 2342–2357. [Google Scholar] [CrossRef]
  33. Anthemidis, A.N.; Ioannou, K.-I.G. On-line sequential injection dispersive liquid–liquid microextraction system for flame atomic absorption spectrometric determination of copper and lead in water samples. Talanta 2009, 79, 86–91. [Google Scholar] [CrossRef] [PubMed]
  34. Anthemidis, A.N.; Ioannou, K.-I.G. Development of a sequential injection dispersive liquid–liquid microextraction system for electrothermal atomic absorption spectrometry by using a hydrophobic sorbent material: Determination of lead and cadmium in natural waters. Anal. Chim. Acta 2010, 668, 35–40. [Google Scholar] [CrossRef] [PubMed]
  35. Anastassiades, M.; Lehotay, S.J.; Stajnbaher, D.; Schenck, F.J. Fast and easy multiresidue method employing acetonitrile extraction/partitioning and “dispersive solid-phase extraction” for the determination of pesticide residues in produce. J. AOAC Int. 2003, 86, 412–431. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  36. Berlioz-Barbier, A.; Baudot, R.; Wiest, L.; Gust, M.; Garric, J.; Cren-Olivé, C.; Buleté, A. MicroQuEChERS–nanoliquid chromatography–nanospray–tandem mass spectrometry for the detection and quantification of trace pharmaceuticals in benthic invertebrates. Talanta 2015, 132, 796–802. [Google Scholar] [CrossRef] [PubMed]
  37. Kachhawaha, A.S.; Nagarnaik, P.M.; Jadhav, M.; Pudale, A.; Labhasetwar, P.K.; Banerjee, K. Optimization of a Modified QuEChERS Method for Multiresidue Analysis of Pharmaceuticals and Personal Care Products in Sewage and Surface Water by LC-MS/MS. J. AOAC Int. 2017, 100, 592–597. [Google Scholar] [CrossRef] [PubMed]
  38. Fernandes, F.C.B.; Silva, A.S.; Rufino, J.L.; Pezza, H.R.; Pezza, L. Screening and determination of sulphonamide residues in bovine milk samples using a flow injection system. Food Chem. 2015, 166, 309–315. [Google Scholar] [CrossRef]
  39. Correia-Sa, L.; Norberto, S.; Delerue-Matos, C.; Calhau, C.; Domingues, V.F. Micro-QuEChERS extraction coupled to GC-MS for a fast determination of Bisphenol A in human urine. J. Chromatogr. B Anal. Technol. Biomed. Life Sci. 2018, 1072, 9–16. [Google Scholar] [CrossRef] [Green Version]
  40. Rial-Berriel, C.; Acosta-Dacal, A.; Zumbado, M.; Luzardo, O.P. Micro QuEChERS-based method for the simultaneous biomonitoring in whole blood of 360 toxicologically relevant pollutants for wildlife. Sci. Total Environ. 2020, 736, 139444. [Google Scholar] [CrossRef]
  41. Khalili Zanjani, M.R.; Yamini, Y.; Shariati, S.; Jönsson, J.A. A new liquid-phase microextraction method based on solidification of floating organic drop. Anal. Chim. Acta 2007, 585, 286–293. [Google Scholar] [CrossRef]
  42. Silva, L.K.; Rangel, J.H.G.; Brito, N.M.; Sousa, E.R.; Sousa, É.M.; Lima, D.L.D.; Esteves, V.I.; Freitas, A.S.; Silva, G.S. Solidified floating organic drop microextraction (SFODME) for the simultaneous analysis of three non-steroidal anti-inflammatory drugs in aqueous samples by HPLC. Anal. Bioanal. Chem. 2021, 413, 1851–1859. [Google Scholar] [CrossRef]
  43. Moghaddam, T.N.; Elhamirad, A.H.; Asl, M.R.S.; Noghabi, M.S. Pulsed electric field-assisted extraction of phenolic antioxidants from tropical almond red leaves. Chem. Pap. 2020, 74, 3957–3961. [Google Scholar] [CrossRef]
  44. Golpayegani, M.R.; Akramipour, R.; Fattahi, N. Sensitive determination of deferasirox in blood of patients with thalassemia using dispersive liquid-liquid microextraction based on solidification of floating organic drop followed by HPLC–UV. J. Pharm. Biomed. Anal. 2021, 193, 113735. [Google Scholar] [CrossRef] [PubMed]
  45. Dadfarnia, S.; Salmanzadeh, A.M.; Shabani, A.M.H. A novel separation/preconcentration system based on solidification of floating organic drop microextraction for determination of lead by graphite furnace atomic absorption spectrometry. Anal. Chim. Acta 2008, 623, 163–167. [Google Scholar] [CrossRef] [PubMed]
  46. Zhou, X.; Gao, Y.; Zhang, Q.; Li, X.; Li, H. Cloud point extraction coupled with ultrasound-assisted back-extraction for determination of trace legacy and emerging brominated flame retardants in water using isotopic dilution high-performance liquid chromatography-atmospheric pressure chemical ionization-tandem mass spectrometry. Talanta 2021, 224, 121713. [Google Scholar] [CrossRef]
  47. Canales, R.; Guiñez, M.; Bazán, C.; Reta, M.; Cerutti, S. Determining heterocyclic aromatic amines in aqueous samples: A novel dispersive liquid-liquid micro-extraction method based on solidification of floating organic drop and ultrasound assisted back extraction followed by UPLC-MS/MS. Talanta 2017, 174, 548–555. [Google Scholar] [CrossRef] [PubMed]
  48. Iqbal, M.; Ezzeldin, E.; Khalil, N.Y.; Alam, P.; Al-Rashood, K.A. UPLC-MS/MS determination of suvorexant in urine by a simplified dispersive liquid-liquid micro-extraction followed by ultrasound assisted back extraction from solidified floating organic droplets. J. Pharm. Biomed. Anal. 2019, 164, 1–8. [Google Scholar] [CrossRef] [PubMed]
  49. Trujillo-Rodriguez, M.J.; Nan, H.; Varona, M.; Emaus, M.N.; Souza, I.D.; Anderson, J.L. Advances of Ionic Liquids in Analytical Chemistry. Anal. Chem. 2019, 91, 505–531. [Google Scholar] [CrossRef]
  50. Soares da Silva Burato, J.; Vargas Medina, D.A.; de Toffoli, A.L.; Vasconcelos Soares Maciel, E.; Mauro Lancas, F. Recent advances and trends in miniaturized sample preparation techniques. J. Sep. Sci. 2020, 43, 202–225. [Google Scholar] [CrossRef]
  51. Lu, D.; Liu, C.; Qin, M.; Deng, J.; Shi, G.; Zhou, T. Functionalized ionic liquids-supported metal organic frameworks for dispersive solid phase extraction of sulfonamide antibiotics in water samples. Anal. Chim. Acta 2020, 1133, 88–98. [Google Scholar] [CrossRef]
  52. Płotka-Wasylka, J.; De la Guardia, M.; Andruch, V.; Vilková, M. Deep eutectic solvents vs ionic liquids: Similarities and differences. Microchem. J. 2020, 159, 105539. [Google Scholar] [CrossRef]
  53. Rocha, C.M.R.R.; Genisheva, Z.; Ferreira-Santos, P.; Rodrigues, R.; Vicente, A.A.; Teixeira, J.A.; Pereira, R.N. Electric field-based technologies for valorization of bioresources. Bioresour. Technol. 2018, 254, 325–339. [Google Scholar] [CrossRef] [Green Version]
  54. Wuethrich, A.; Haddad, P.R.; Quirino, J.P. The electric field—An emerging driver in sample preparation. TrAC Trends Anal. Chem. 2016, 80, 604–611. [Google Scholar] [CrossRef]
  55. Xu, R.; Lee, H.K. Application of electro-enhanced solid phase microextraction combined with gas chromatography–mass spectrometry for the determination of tricyclic antidepressants in environmental water samples. J. Chromatogr. A 2014, 1350, 15–22. [Google Scholar] [CrossRef] [PubMed]
  56. Mousa, A.; Basheer, C.; Al-Arfaj, A.R. Application of electro-enhanced solid-phase microextraction for determination of phthalate esters and bisphenol A in blood and seawater samples. Talanta 2013, 115, 308–313. [Google Scholar] [CrossRef] [PubMed]
  57. Ribeiro, C.C.; Orlando, R.M.; Rohwedder, J.J.; Reyes, F.G.; Rath, S. Electric field-assisted solid phase extraction and cleanup of ionic compounds in complex food matrices: Fluoroquinolones in eggs. Talanta 2016, 152, 498–503. [Google Scholar] [CrossRef]
  58. Huang, C.; Seip, K.F.; Gjelstad, A.; Pedersen-Bjergaard, S. Electromembrane extraction for pharmaceutical and biomedical analysis-Quo vadis. J. Pharm. Biomed. Anal. 2015, 113, 97–107. [Google Scholar] [CrossRef]
  59. Gomes-Dias, J.S.; Romaní, A.; Teixeira, J.A.; Rocha, C.M.R.R. Valorization of Seaweed Carbohydrates: Autohydrolysis as a Selective and Sustainable Pretreatment. ACS Sustain. Chem. Eng. 2020, 8, 17143–17153. [Google Scholar] [CrossRef]
  60. Nowak, P.M.; Wietecha-Posłuszny, R.; Pawliszyn, J. White Analytical Chemistry: An approach to reconcile the principles of Green Analytical Chemistry and functionality. TrAC Trends Anal. Chem. 2021, 138, 116223. [Google Scholar] [CrossRef]
  61. Chemat, F.; Fabiano-Tixier, A.S.; Vian, M.A.; Allaf, T.; Vorobiev, E. Solvent-free extraction of food and natural products. TrAC, Trends Anal. Chem. 2015, 71, 157–168. [Google Scholar] [CrossRef] [Green Version]
  62. Khiltash, S.; Heydari, R.; Ramezani, M. Graphene oxide/polydopamine-polyacrylamide nanocomposite as a sorbent for dispersive micro-solid phase extraction of diazinon from environmental and food samples and its determination by HPLC-UV detection. Int. J. Environ. Anal. Chem. 2021. [Google Scholar] [CrossRef]
  63. Barbosa-Pereira, L.; Guglielmetti, A.; Zeppa, G. Pulsed Electric Field Assisted Extraction of Bioactive Compounds from Cocoa Bean Shell and Coffee Silverskin. Food Bioprocess Technol. 2018, 11, 818–835. [Google Scholar] [CrossRef]
  64. Gazioglu, I.; Zengin, O.S.; Tartaglia, A.; Locatelli, M.; Furton, K.G.; Kabir, A. Determination of polycyclic aromatic hydrocarbons in nutritional supplements by fabric phase sorptive extraction (FPSE) with high-performance liquid chromatography (HPLC) with fluorescence detection. Anal. Lett. 2021, 54, 1683–1696. [Google Scholar] [CrossRef]
  65. Hrynko, I.; Łozowicka, B.; Kaczyński, P. Development of precise micro analytical tool to identify potential insecticide hazards to bees in guttation fluid using LC-ESI-MS/MS. Chemosphere 2021, 263, 128143. [Google Scholar] [CrossRef] [PubMed]
  66. Izcara, S.; Casado, N.; Morante-Zarcero, S.; Sierra, I. A miniaturized QuEChERS method combined with ultrahigh liquid chromatography coupled to tandem mass spectrometry for the analysis of pyrrolizidine alkaloids in oregano samples. Foods 2020, 9, 1319. [Google Scholar] [CrossRef]
  67. Chen, D.; Miao, H.; Zhao, Y.; Wu, Y. A simple liquid chromatography-high resolution mass spectrometry method for the determination of glyphosate and aminomethylphosphonic acid in human urine using cold-induced phase separation and hydrophilic pipette tip solid-phase extraction. J. Chromatogr. A 2019, 1587, 73–78. [Google Scholar] [CrossRef]
  68. Atirah Mohd Nazir, N.; Raoov, M.; Mohamad, S. Spent tea leaves as an adsorbent for micro-solid-phase extraction of polycyclic aromatic hydrocarbons (PAHs) from water and food samples prior to GC-FID analysis. Microchem. J. 2020, 159, 105581. [Google Scholar] [CrossRef]
  69. Moufid, M.; Hofmann, M.; El Bari, N.; Tiebe, C.; Bartholmai, M.; Bouchikhi, B. Wastewater monitoring by means of e-nose, VE-tongue, TD-GC-MS, and SPME-GC-MS. Talanta 2021, 221, 121450. [Google Scholar] [CrossRef]
  70. Prata, M.; Ribeiro, A.; Figueirinha, D.; Rosado, T.; Oppolzer, D.; Restolho, J.; Araújo, A.R.T.S.; Costa, S.; Barroso, M.; Gallardo, E. Determination of opiates in whole blood using microextraction by packed sorbent and gas chromatography-tandem mass spectrometry. J. Chromatogr. A 2019, 1602, 1–10. [Google Scholar] [CrossRef]
  71. Altunay, N.; Elik, A.; Gürkan, R. Monitoring of some trace metals in honeys by flame atomic absorption spectrometry after ultrasound assisted-dispersive liquid liquid microextraction using natural deep eutectic solvent. Microchem. J. 2019, 147, 49–59. [Google Scholar] [CrossRef]
  72. Shirkhanloo, H.; Khaleghi Abbasabadi, M.; Hosseini, F.; Faghihi Zarandi, A. Nanographene oxide modified phenyl methanethiol nanomagnetic composite for rapid separation of aluminum in wastewaters, foods, and vegetable samples by microwave dispersive magnetic micro solid-phase extraction. Food Chem. 2021, 347, 129042. [Google Scholar] [CrossRef]
  73. Nyaba, L.; Nomngongo, P.N. Determination of trace metals in vegetables and water samples using dispersive ultrasound-assisted cloud point-dispersive µ-solid phase extraction coupled with inductively coupled plasma optical emission spectrometry. Food Chem. 2020, 322, 126749. [Google Scholar] [CrossRef] [PubMed]
  74. Ulrich, S. Solid-phase microextraction in biomedical analysis. J. Chromatogr. A 2000, 902, 167–194. [Google Scholar] [CrossRef]
  75. Souza-Silva, E.A.; Jiang, R.F.; Rodriguez-Lafuente, A.; Gionfriddo, E.; Pawliszyn, J. A critical review of the state of the art of solid-phase microextraction of complex matrices I. Environmental analysis. TrAC Trends Anal. Chem. 2015, 71, 224–235. [Google Scholar] [CrossRef]
  76. Roszkowska, A.; Miękus, N.; Bączek, T. Application of solid-phase microextraction in current biomedical research. J. Sep. Sci. 2019, 42, 285–302. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  77. Silva, C.L.; Perestrelo, R.; Capelinha, F.; Tomás, H.; Câmara, J.S. An integrative approach based on GC–qMS and NMR metabolomics data as a comprehensive strategy to search potential breast cancer biomarkers. Metabolomics 2021, 17, 72. [Google Scholar] [CrossRef]
  78. Zhang, Y.; Zhao, Y.-G.; Chen, W.-S.; Cheng, H.-L.; Zeng, X.-Q.; Zhu, Y. Three-dimensional ionic liquid-ferrite functionalized graphene oxide nanocomposite for pipette-tip solid phase extraction of 16 polycyclic aromatic hydrocarbons in human blood sample. J. Chromatogr. A 2018, 1552, 1–9. [Google Scholar] [CrossRef]
  79. Lari, A.; Esmaeili, N.; Ghafari, H. Ionic liquid functionlized on multiwall carbon nanotubes for nickel and lead determination in human serum and urine samples by micro solid-phase extraction. Anal. Methods Environ. Chem. J. 2021, 4, 72–85. [Google Scholar] [CrossRef]
  80. Lioupi, A.; Kabir, A.; Furton, K.G.; Samanidou, V. Fabric phase sorptive extraction for the isolation of five common antidepressants from human urine prior to HPLC-DAD analysis. J. Chromatogr. B Anal. Technol. Biomed. Life Sci. 2019, 1118–1119, 171–179. [Google Scholar] [CrossRef]
  81. Musteata, F.M. Recent progress in in-vivo sampling and analysis. TrAC Trends Anal. Chem. 2013, 45, 154–168. [Google Scholar] [CrossRef]
  82. Cudjoe, E.; Bojko, B.; de Lannoy, I.; Saldivia, V.; Pawliszyn, J. Solid-phase microextraction: A complementary in vivo sampling method to microdialysis. Angew. Chem. Int. Ed. Engl. 2013, 52, 12124–12126. [Google Scholar] [CrossRef]
  83. Samanidou, V.; Filippou, O.; Marinou, E.; Kabir, A.; Furton, K.G. Sol-gel-graphene-based fabric-phase sorptive extraction for cow and human breast milk sample cleanup for screening bisphenol A and residual dental restorative material before analysis by HPLC with diode array detection. J. Sep. Sci. 2017, 40, 2612–2619. [Google Scholar] [CrossRef] [PubMed]
  84. Kabir, A.; Furton, K.G.; Tinari, N.; Grossi, L.; Innosa, D.; Macerola, D.; Tartaglia, A.; Di Donato, V.; D’Ovidio, C.; Locatelli, M. Fabric phase sorptive extraction-high performance liquid chromatography-photo diode array detection method for simultaneous monitoring of three inflammatory bowel disease treatment drugs in whole blood, plasma and urine. J. Chromatogr. B Anal. Technol. Biomed. Life Sci. 2018, 1084, 53–63. [Google Scholar] [CrossRef] [PubMed]
  85. Locatelli, M.; Tinari, N.; Grassadonia, A.; Tartaglia, A.; Macerola, D.; Piccolantonio, S.; Sperandio, E.; D’Ovidio, C.; Carradori, S.; Ulusoy, H.I.; et al. FPSE-HPLC-DAD method for the quantification of anticancer drugs in human whole blood, plasma, and urine. J. Chromatogr. B Anal. Technol. Biomed. Life Sci. 2018, 1095, 204–213. [Google Scholar] [CrossRef] [PubMed]
  86. Taraboletti, A.; Goudarzi, M.; Kabir, A.; Moon, B.H.; Laiakis, E.C.; Lacombe, J.; Ake, P.; Shoishiro, S.; Brenner, D.; Fornace, A.J., Jr.; et al. Fabric Phase Sorptive Extraction-A Metabolomic Preprocessing Approach for Ionizing Radiation Exposure Assessment. J. Proteome Res. 2019, 18, 3020–3031. [Google Scholar] [CrossRef]
  87. Manousi, N.; Plastiras, O.E.; Deliyanni, E.A.; Zachariadis, G.A. Green Bioanalytical Applications of Graphene Oxide for the Extraction of Small Organic Molecules. Molecules 2021, 26, 2790. [Google Scholar] [CrossRef]
  88. Asgharinezhad, A.A.; Ebrahimzadeh, H. Poly(2-aminobenzothiazole)-coated graphene oxide/magnetite nanoparticles composite as an efficient sorbent for determination of non-steroidal anti-inflammatory drugs in urine sample. J. Chromatogr. A 2016, 1435, 18–29. [Google Scholar] [CrossRef]
  89. Lamei, N.; Ezoddin, M.; Ardestani, M.S.; Abdi, K. Dispersion of magnetic graphene oxide nanoparticles coated with a deep eutectic solvent using ultrasound assistance for preconcentration of methadone in biological and water samples followed by GC–FID and GC–MS. Anal. Bioanal. Chem. 2017, 409, 6113–6121. [Google Scholar] [CrossRef]
  90. Taghvimi, A.; Hamishehkar, H.; Ebrahimi, M. Magnetic nano graphene oxide as solid phase extraction adsorbent coupled with liquid chromatography to determine pseudoephedrine in urine samples. J. Chromatogr. B-Anal. Technol. Biomed. Life Sci. 2016, 1009, 66–72. [Google Scholar] [CrossRef]
  91. Barati, A.; Kazemi, E.; Dadfarnia, S.; Shabani, A.M.H. Synthesis/characterization of molecular imprinted polymer based on magnetic chitosan/graphene oxide for selective separation/preconcentration of fluoxetine from environmental and biological samples. J. Ind. Eng. Chem. 2017, 46, 212–221. [Google Scholar] [CrossRef]
  92. Peng, J.; Tian, H.R.; Du, Q.Z.; Hui, X.H.; He, H. A regenerable sorbent composed of a zeolite imidazolate framework (ZIF-8), Fe3O4 and graphene oxide for enrichment of atorvastatin and simvastatin prior to their determination by HPLC. Mikrochim. Acta 2018, 185, 141. [Google Scholar] [CrossRef]
  93. Zhang, J.; Liu, D.; Meng, X.; Shi, Y.; Wang, R.; Xiao, D.; He, H. Solid phase extraction based on porous magnetic graphene oxide/beta-cyclodextrine composite coupled with high performance liquid chromatography for determination of antiepileptic drugs in plasma samples. J. Chromatogr. A 2017, 1524, 49–56. [Google Scholar] [CrossRef] [PubMed]
  94. Yuvali, D.; Narin, I.; Soylak, M.; Yilmaz, E. Green synthesis of magnetic carbon nanodot/graphene oxide hybrid material (Fe3O4@C-nanodot@GO) for magnetic solid phase extraction of ibuprofen in human blood samples prior to HPLC-DAD determination. J. Pharm. Biomed. Anal. 2020, 179, 113001. [Google Scholar] [CrossRef] [PubMed]
  95. Mabrouk, M.M.; Soliman, S.M.; El-Agizy, H.M.; Mansour, F.R. Ultrasound-assisted dispersive liquid–liquid microextraction for determination of three gliflozins in human plasma by HPLC/DAD. J. Chromatogr. B 2020, 1136, 121932. [Google Scholar] [CrossRef] [PubMed]
  96. Ghoraba, Z.; Aibaghi, B.; Soleymanpour, A. Ultrasound-assisted dispersive liquid-liquid microextraction followed by ion mobility spectrometry for the simultaneous determination of bendiocarb and azinphos-ethyl in water, soil, food and beverage samples. Ecotoxicol. Environ. Saf. 2018, 165, 459–466. [Google Scholar] [CrossRef]
  97. Kaykhaii, M.; Yavari, E.; Sargazi, G.; Ebrahimi, A.K. Highly Sensitive Determination of Bisphenol A in Bottled Water Samples by HPLC after Its Extraction by a Novel Th-MOF Pipette-Tip Micro-SPE. J. Chromatogr. Sci. 2020, 58, 373–382. [Google Scholar] [CrossRef]
  98. Alipanahpour Dil, E.; Asfaram, A.; Goudarzi, A.; Zabihi, E.; Javadian, H. Biocompatible chitosan-zinc oxide nanocomposite based dispersive micro-solid phase extraction coupled with HPLC-UV for the determination of rosmarinic acid in the extracts of medical plants and water sample. Int. J. Biol. Macromol. 2020, 154, 528–537. [Google Scholar] [CrossRef]
  99. Sereshti, H.; Toloutehrani, A.; Nodeh, H.R. Determination of cholecalciferol (vitamin D3) in bovine milk by dispersive micro-solid phase extraction based on the magnetic three-dimensional graphene-sporopollenin sorbent. J. Chromatogr. B 2020, 1136, 121907. [Google Scholar] [CrossRef]
  100. Mousavi, K.Z.; Yamini, Y.; Karimi, B.; Seidi, S.; Khorasani, M.; Ghaemmaghami, M.; Vali, H. Imidazolium-based mesoporous organosilicas with bridging organic groups for microextraction by packed sorbent of phenoxy acid herbicides, polycyclic aromatic hydrocarbons and chlorophenols. Microchim. Acta 2019, 186, 239. [Google Scholar] [CrossRef]
  101. Teixeira, R.A.; Dinali, L.A.F.; Silva, C.F.; de Oliveira, H.L.; da Silva, A.T.M.; Nascimento, C.S.; Borges, K.B. Microextraction by packed molecularly imprinted polymer followed by ultra-high performance liquid chromatography for determination of fipronil and fluazuron residues in drinking water and veterinary clinic wastewater. Microchem. J. 2021, 168, 106405. [Google Scholar] [CrossRef]
  102. Dinali, L.A.F.; de Oliveira, H.L.; Teixeira, L.S.; de Souza Borges, W.; Borges, K.B. Mesoporous molecularly imprinted polymer core@shell hybrid silica nanoparticles as adsorbent in microextraction by packed sorbent for multiresidue determination of pesticides in apple juice. Food Chem. 2021, 345, 128745. [Google Scholar] [CrossRef]
  103. Aresta, A.; Cotugno, P.; Zambonin, C. Determination of ciprofloxacin, enrofloxacin, and marbofloxacin in bovine urine, serum, and milk by microextraction by a packed sorbent coupled to ultra-high performance liquid chromatography. Anal. Lett. 2019, 52, 790–802. [Google Scholar] [CrossRef]
  104. Florez, D.H.Â.; de Oliveira, H.L.; Borges, K.B. Polythiophene as highly efficient sorbent for microextraction in packed sorbent for determination of steroids from bovine milk samples. Microchem. J. 2020, 153, 104521. [Google Scholar] [CrossRef]
  105. Jiang, Y.; Qin, Z.; Song, X.; Piao, H.; Li, J.; Wang, X.; Song, D.; Ma, P.; Sun, Y. Facile preparation of metal organic framework-based laboratory semi-automatic micro-extraction syringe packed column for analysis of parabens in vegetable oil samples. Microchem. J. 2020, 158, 105200. [Google Scholar] [CrossRef]
  106. Paris, A.; Gaillard, J.L.; Ledauphin, J. Rapid extraction of polycyclic aromatic hydrocarbons in apple: Ultrasound-assisted solvent extraction followed by microextraction by packed sorbent. Food Anal. Methods 2019, 12, 2194–2204. [Google Scholar] [CrossRef]
  107. Teixeira, L.S.; Silva, C.F.; de Oliveira, H.L.; Dinali, L.A.F.; Nascimento, C.S.; Borges, K.B. Microextraction by packed molecularly imprinted polymer to selectively determine caffeine in soft and energy drinks. Microchem. J. 2020, 158, 105252. [Google Scholar] [CrossRef]
  108. Casado, N.; Perestrelo, R.; Silva, C.L.; Sierra, I.; Câmara, J.S. Comparison of high-throughput microextraction techniques, MEPS and μ-SPEed, for the determination of polyphenols in baby food by ultrahigh pressure liquid chromatography. Food Chem. 2019, 292, 14–23. [Google Scholar] [CrossRef]
  109. Kalogiouri, N.P.; Manousi, N.; Rosenberg, E.; Zachariadis, G.A.; Paraskevopoulou, A.; Samanidou, V. Exploring the volatile metabolome of conventional and organic walnut oils by solid-phase microextraction and analysis by GC-MS combined with chemometrics. Food Chem. 2021, 363, 130331. [Google Scholar] [CrossRef]
  110. Zhao, C.C.; Eun, J.B. Characterization of volatile compounds and physicochemical properties of hongeo using headspace solid-phase microextraction and gas chromatography-mass spectrometry during fermentation. Food Biosci. 2021, 44, 101379. [Google Scholar] [CrossRef]
  111. Majithia, D.; Metrani, R.; Dhowlaghar, N.; Crosby, K.M.; Patil, B.S. Assessment and classification of volatile profiles in melon breeding lines using headspace solid-phase microextraction coupled with gas chromatography-mass spectrometry. Plants 2021, 10, 2166. [Google Scholar] [CrossRef]
  112. Thomas, C.F.; Zeh, E.; Dörfel, S.; Zhang, Y.; Hinrichs, J. Studying dynamic aroma release by headspace-solid phase microextraction-gas chromatography-ion mobility spectrometry (HS-SPME-GC-IMS): Method optimization, validation, and application. Anal. Bioanal. Chem. 2021, 413, 2577–2586. [Google Scholar] [CrossRef]
  113. Zianni, R.; Mentana, A.; Campaniello, M.; Chiappinelli, A.; Tomaiuolo, M.; Chiaravalle, A.E.; Marchesani, G. An investigation using a validated method based on HS-SPME-GC-MS detection for the determination of 2-dodecylcyclobutanone and 2-tetradecylcyclobutanone in X-ray irradiated dairy products. LWT 2022, 153, 112466. [Google Scholar] [CrossRef]
  114. He, X.; Majid, B.; Zhang, H.; Liu, W.; Limmer, M.A.; Burken, J.G.; Shi, H. Green analysis: Rapid-throughput analysis of volatile contaminants in plants by freeze-thaw-equilibration sample preparation and SPME-GC-MS analysis. J. Agric. Food. Chem. 2021, 69, 5428–5434. [Google Scholar] [CrossRef] [PubMed]
  115. Passos, C.P.; Petronilho, S.; Serodio, A.F.; Neto, A.C.M.; Torres, D.; Rudnitskaya, A.; Nunes, C.; Kukurova, K.; Ciesarova, Z.; Rocha, S.M.; et al. HS-SPME Gas Chromatography Approach for Underivatized Acrylamide Determination in Biscuits. Foods 2021, 10, 2183. [Google Scholar] [CrossRef] [PubMed]
  116. Pang, L.; Yang, P.; Pang, R.; Lu, X.; Xiao, J.; Li, S.; Zhang, H.; Zhao, J. Ionogel-based ionic liquid coating for solid-phase microextraction of organophosphorus pesticides from wine and juice samples. Food Anal. Methods 2018, 11, 270–281. [Google Scholar] [CrossRef]
  117. Perestrelo, R.; Silva, C.L.; Algarra, M.; Câmara, J.S. Evaluation of the occurrence of phthalates in plastic materials used in food packaging. Appl. Sci. 2021, 11, 2130. [Google Scholar] [CrossRef]
  118. Chen, Y.; Zhang, Y.; Xu, L. A rapid method for analyzing synthetic phenolic antioxidants in food grade lubricant samples based on headspace solid-phase microextracion coupled with gas chromatography-mass spectrometer. Food Anal. Methods 2021, 14, 2524–2533. [Google Scholar] [CrossRef]
  119. Mejía-Carmona, K.; Lanças, F.M. Modified graphene-silica as a sorbent for in-tube solid-phase microextraction coupled to liquid chromatography-tandem mass spectrometry. Determination of xanthines in coffee beverages. J. Chromatogr. A 2020, 1621, 461089. [Google Scholar] [CrossRef]
  120. Peng, L.Q.; Zhang, Y.; Yan, T.C.; Gu, Y.X.; Zi, X.; Cao, J. Carbonized biosorbent assisted matrix solid-phase dispersion microextraction for active compounds from functional food. Food Chem. 2021, 365, 130545. [Google Scholar] [CrossRef]
  121. Gomez-Mejia, E.; Mikkelsen, L.H.; Rosales-Conrado, N.; Leon-Gonzalez, M.E.; Madrid, Y. A combined approach based on matrix solid-phase dispersion extraction assisted by titanium dioxide nanoparticles and liquid chromatography to determine polyphenols from grape residues. J. Chromatogr. A 2021, 1644, 462128. [Google Scholar] [CrossRef]
  122. Segatto, M.L.; Zanotti, K.; Zuin, V.G. Microwave-assisted extraction and matrix solid-phase dispersion as green analytical chemistry sample preparation techniques for the valorisation of mango processing waste. Curr. Res. Chem. Biol. 2021, 1, 100007. [Google Scholar] [CrossRef]
  123. Qian, Z.; Wu, Z.; Li, C.; Tan, G.; Hu, H.; Li, W. A green liquid chromatography method for rapid determination of ergosterol in edible fungi based on matrix solid-phase dispersion extraction and a core-shell column. Anal. Methods 2020, 12, 3327–3343. [Google Scholar] [CrossRef] [PubMed]
  124. Martín-Girela, I.; Albero, B.; Tiwari, B.K.; Miguel, E.; Aznar, R. Screening of contaminants of emerging concern in microalgae food supplements. Separations 2020, 7, 28. [Google Scholar] [CrossRef]
  125. Souza, M.R.R.; Jesus, R.A.; Costa, J.A.S.; Barreto, A.S.; Navickiene, S.; Mesquita, M.E. Applicability of metal–organic framework materials in the evaluation of pesticide residues in egg samples of chicken (Gallus gallus domesticus). J. Consum. Prot. Food Saf. 2021, 16, 83–91. [Google Scholar] [CrossRef]
  126. Liang, T.; Gao, L.; Qin, D.; Chen, L. Determination of sulfonylurea herbicides in grain samples by matrix solid-phase dispersion with mesoporous structured molecularly imprinted polymer. Food Anal. Methods 2019, 12, 1938–1948. [Google Scholar] [CrossRef]
  127. Narimani-Sabegh, S.; Noroozian, E. Magnetic solid-phase extraction and determination of ultra-trace amounts of antimony in aqueous solutions using maghemite nanoparticles. Food Chem. 2019, 287, 382–389. [Google Scholar] [CrossRef]
  128. Boon, Y.H.; Mohamad Zain, N.N.; Mohamad, S.; Osman, H.; Raoov, M. Magnetic poly(beta-cyclodextrin-ionic liquid) nanocomposites for micro-solid phase extraction of selected polycyclic aromatic hydrocarbons in rice samples prior to GC-FID analysis. Food Chem. 2019, 278, 322–332. [Google Scholar] [CrossRef]
  129. Chen, J.Y.; Cao, S.R.; Xi, C.X.; Chen, Y.; Li, X.L.; Zhang, L.; Wang, G.M.; Chen, Y.L.; Chen, Z.Q. A novel magnetic β-cyclodextrin modified graphene oxide adsorbent with high recognition capability for 5 plant growth regulators. Food Chem. 2018, 239, 911–919. [Google Scholar] [CrossRef]
  130. Rahimi, A.; Zanjanchi, M.A.; Bakhtiari, S.; Dehsaraei, M. Selective determination of caffeine in foods with 3D-graphene based ultrasound-assisted magnetic solid phase extraction. Food Chem. 2018, 262, 206–214. [Google Scholar] [CrossRef]
  131. Mesa, R.; Kabir, A.; Samanidou, V.; Furton, K.G. Simultaneous determination of selected estrogenic endocrine disrupting chemicals and bisphenol A residues in whole milk using fabric phase sorptive extraction coupled to HPLC-UV detection and LC-MS/MS. J. Sep. Sci. 2019, 42, 598–608. [Google Scholar] [CrossRef]
  132. Ubeda, S.; Aznar, M.; Nerín, C.; Kabir, A. Fabric phase sorptive extraction for specific migration analysis of oligomers from biopolymers. Talanta 2021, 233, 122603. [Google Scholar] [CrossRef]
  133. Guedes-Alonso, R.; Sosa-Ferrera, Z.; Santana-Rodríguez, J.J.; Kabir, A.; Furton, K.G. Fabric phase sorptive extraction of selected steroid hormone residues in commercial raw milk followed by ultra-high-performance liquid chromatography–tandem mass spectrometry. Foods 2021, 10, 343. [Google Scholar] [CrossRef] [PubMed]
  134. Agadellis, E.; Tartaglia, A.; Locatelli, M.; Kabir, A.; Furton, K.G.; Samanidou, V. Mixed-mode fabric phase sorptive extraction of multiple tetracycline residues from milk samples prior to high performance liquid chromatography-ultraviolet analysis. Microchem. J. 2020, 159, 105437. [Google Scholar] [CrossRef]
  135. Campone, L.; Celano, R.; Piccinelli, A.L.; Pagano, I.; Cicero, N.; Sanzo, R.D.; Carabetta, S.; Russo, M.; Rastrelli, L. Ultrasound assisted dispersive liquid-liquid microextraction for fast and accurate analysis of chloramphenicol in honey. Food Res. Int. 2019, 115, 572–579. [Google Scholar] [CrossRef] [PubMed]
  136. Qiao, L.Z.; Sun, R.T.; Yu, C.M.; Tao, Y.; Yan, Y. Novel hydrophobic deep eutectic solvents for ultrasound-assisted dispersive liquid-liquid microextraction of trace non-steroidal anti-inflammatory drugs in water and milk samples. Microchem. J. 2021, 170, 106686. [Google Scholar] [CrossRef]
  137. Bernardi, G.; Kemmerich, M.; Adaime, M.B.; Prestes, O.D.; Zanella, R. Miniaturized QuEChERS method for determination of 97 pesticide residues in wine by ultra-high performance liquid chromatography coupled with tandem mass spectrometry. Anal. Methods 2020, 12, 2682–2692. [Google Scholar] [CrossRef] [PubMed]
  138. Kamal El-Deen, A.; Shimizu, K. Modified μ-QuEChERS coupled to diethyl carbonate-based liquid microextraction for PAHs determination in coffee, tea, and water prior to GC–MS analysis: An insight to reducing the impact of caffeine on the GC–MS measurement. J. Chromatogr. B Anal. Technol. Biomed. Life Sci. 2021, 1171, 122555. [Google Scholar] [CrossRef]
  139. Casado, N.; Perestrelo, R.; Silva, C.L.; Sierra, I.; Câmara, J.S. An improved and miniaturized analytical strategy based on μ-QuEChERS for isolation of polyphenols. A powerful approach for quality control of baby foods. Microchem. J. 2018, 139, 110–118. [Google Scholar] [CrossRef]
  140. Abreu, D.C.P.; Botrel, B.M.C.; Bazana, M.J.F.; e Rosa, P.V.; Sales, P.F.; Marques, M.d.S.; Saczk, A.A. Development and comparative analysis of single-drop and solid-phase microextraction techniques in the residual determination of 2-phenoxyethanol in fish. Food Chem. 2019, 270, 487–493. [Google Scholar] [CrossRef]
  141. Saraji, M.; Javadian, S. Single-drop microextraction combined with gas chromatography-electron capture detection for the determination of acrylamide in food samples. Food Chem. 2019, 274, 55–60. [Google Scholar] [CrossRef]
  142. Jain, A.; Soni, S.; Verma, K.K. Combined liquid phase microextraction and fiber-optics-based cuvetteless micro-spectrophotometry for sensitive determination of ammonia in water and food samples by the indophenol reaction. Food Chem. 2021, 340, 128156. [Google Scholar] [CrossRef]
  143. Ma, Z.; Zhao, T.; Cui, S.; Zhao, X.; Fan, Y.; Song, J. Determination of ethyl carbamate in wine by matrix modification-assisted headspace single-drop microextraction and gas chromatography—mass spectrometry technique. Food Chem. 2021, 373, 131573. [Google Scholar] [CrossRef] [PubMed]
  144. Qi, T.; Xu, M.; Yao, Y.; Chen, W.; Xu, M.; Tang, S.; Shen, W.; Kong, D.; Cai, X.; Shi, H.; et al. Gold nanoprism/Tollens’ reagent complex as plasmonic sensor in headspace single-drop microextraction for colorimetric detection of formaldehyde in food samples using smartphone readout. Talanta 2020, 220, 121388. [Google Scholar] [CrossRef] [PubMed]
  145. Tiwari, S.; Deb, M.K. Modified silver nanoparticles-enhanced single drop microextraction of tartrazine in food samples coupled with diffuse reflectance Fourier transform infrared spectroscopic analysis. Anal. Methods 2019, 11, 3552–3562. [Google Scholar] [CrossRef]
  146. Neri, T.S.; Rocha, D.P.; Munoz, R.A.A.; Coelho, N.M.M.; Batista, A.D. Highly sensitive procedure for determination of Cu(II) by GF AAS using single-drop microextraction. Microchem. J. 2019, 147, 894–898. [Google Scholar] [CrossRef]
  147. Tavakoli, M.; Jamali, M.R.; Nezhadali, A. Ultrasound-Assisted Dispersive Liquid–Liquid Microextraction (DLLME) Based on Solidification of Floating Organic Drop Using a Deep Eutectic Solvent for Simultaneous Preconcentration and Determination of Nickel and Cobalt in Food and Water Samples. Anal. Lett. 2021, 54, 2863–2873. [Google Scholar] [CrossRef]
  148. Shirani, M.; Akbari-adergani, B.; Shahdadi, F.; Faraji, M.; Akbari, A. A Hydrophobic Deep Eutectic Solvent-Based Ultrasound-Assisted Dispersive Liquid–Liquid Microextraction for Determination of β-Lactam Antibiotics Residues in Food Samples. Food Anal. Methods 2021, 15, 391–400. [Google Scholar] [CrossRef]
  149. Mardani, A.; Torbati, M.; Farajzadeh, M.A.; Mohebbi, A.; Alizadeh, A.A.; Afshar Mogaddam, M.R. Development of temperature-assisted solidification of floating organic droplet-based dispersive liquid–liquid microextraction performed during centrifugation for extraction of organochlorine pesticide residues in cocoa powder prior to GC-ECD. Chem. Pap. 2021, 75, 1691–1700. [Google Scholar] [CrossRef]
  150. Pataro, G.; Carullo, D.; Ferrari, G. Effect of PEF pre-treatment and extraction temperature on the recovery of carotenoids from tomato wastes. Chem. Eng. Trans. 2019, 75, 139–144. [Google Scholar] [CrossRef]
  151. Dong, Z.Y.; Wang, H.H.; Li, M.Y.; Liu, W.; Zhang, T.H. Optimization of high-intensity pulsed electric field-assisted extraction of procyanidins from Vitis amurensis seeds using response surface methodology. E3S Web Conf. 2020, 189, 02029. [Google Scholar] [CrossRef]
  152. Mahn, A.; Comett, R.; Segura-Ponce, L.A.; Díaz-Álvarez, R.E. Effect of pulsed electric field-assisted extraction on recovery of sulforaphane from broccoli florets. J. Food Process Eng. 2021, e13837. [Google Scholar] [CrossRef]
  153. Santos, P.H.; Kammers, J.C.; Silva, A.P.; Oliveira, J.V.; Hense, H. Antioxidant and antibacterial compounds from feijoa leaf extracts obtained by pressurized liquid extraction and supercritical fluid extraction. Food Chem. 2021, 344, 128620. [Google Scholar] [CrossRef] [PubMed]
  154. Santos, O.V.; Lorenzo, N.D.; Souza, A.L.G.; Costa, C.E.F.; Conceição, L.R.V.; Lannes, S.C.d.S.; Teixeira-Costa, B.E. CO2 supercritical fluid extraction of pulp and nut oils from Terminalia catappa fruits: Thermogravimetric behavior, spectroscopic and fatty acid profiles. Food Res. Int. 2021, 139, 109814. [Google Scholar] [CrossRef] [PubMed]
  155. Fornereto Soldan, A.C.; Arvelos, S.; Watanabe, É.O.; Hori, C.E. Supercritical fluid extraction of oleoresin from Capsicum annuum industrial waste. J. Clean. Prod. 2021, 297, 126593. [Google Scholar] [CrossRef]
  156. Arturo-Perdomo, D.; Mora, J.P.J.; Ibáñez, E.; Cifuentes, A.; Hurtado-Benavides, A.; Montero, L. Extraction and Characterization of the Polar Lipid Fraction of Blackberry and Passion Fruit Seeds Oils Using Supercritical Fluid Extraction. Food Anal. Methods 2021, 14, 2026–2037. [Google Scholar] [CrossRef]
  157. Jha, A.K.; Sit, N. Comparison of response surface methodology (RSM) and artificial neural network (ANN) modelling for supercritical fluid extraction of phytochemicals from Terminalia chebula pulp and optimization using RSM coupled with desirability function (DF) and genetic. Ind. Crops Prod. 2021, 170, 113769. [Google Scholar] [CrossRef]
  158. Wang, Y.; Ye, Y.; Wang, L.; Yin, W.; Liang, J. Antioxidant activity and subcritical water extraction of anthocyanin from raspberry process optimization by response surface methodology. Food Bioscience 2021, 44, 101394. [Google Scholar] [CrossRef]
  159. Pangestuti, R.; Haq, M.; Rahmadi, P.; Chun, B.-s. Nutritional Value and Biofunctionalities of Two Edible Green Seaweeds (Ulva lactuca and Caulerpa racemosa) from Indonesia by Subcritical Water Hydrolysis. Mar. Drugs 2021, 19, 578. [Google Scholar] [CrossRef]
  160. Hwang, H.J.; Kim, H.J.; Ko, M.J.; Chung, M.S. Recovery of hesperidin and narirutin from waste Citrus unshiu peel using subcritical water extraction aided by pulsed electric field treatment. Food Sci. Biotechnol. 2021, 30, 217–226. [Google Scholar] [CrossRef]
  161. Pinto, D.; Vieira, E.F.; Peixoto, A.F.; Freire, C.; Freitas, V.; Costa, P.; Delerue-Matos, C.; Rodrigues, F. Optimizing the extraction of phenolic antioxidants from chestnut shells by subcritical water extraction using response surface methodology. Food Chem. 2021, 334, 127521. [Google Scholar] [CrossRef]
  162. Jamaludin, R.; Kim, D.S.; Salleh, L.M.; Lim, S.B. Kinetic study of subcritical water extraction of scopoletin, alizarin, and rutin from morinda citrifolia. Foods 2021, 10, 2260. [Google Scholar] [CrossRef]
  163. Rodrigues, L.A.; Matias, A.A.; Paiva, A. Recovery of antioxidant protein hydrolysates from shellfish waste streams using subcritical water extraction. Food Bioprod. Process. 2021, 130, 154–163. [Google Scholar] [CrossRef]
  164. Zhang, F.; Zhang, L.; Chen, J.; Du, X.; Lu, Z.; Wang, X.; Yi, Y.; Shan, Y.; Liu, B.; Zhou, Y.; et al. Systematic evaluation of a series of pectic polysaccharides extracted from apple pomace by regulation of subcritical water conditions. Food Chem. 2022, 368, 130833. [Google Scholar] [CrossRef] [PubMed]
  165. Darvishnejad, M.; Ebrahimzadeh, H. Graphitic carbon nitride-reinforced polymer ionic liquid nanocomposite: A novel mixed-mode sorbent for microextraction in packed syringe. Int. J. Environ. Anal. Chem. 2020, 1–14. [Google Scholar] [CrossRef]
  166. Saraji, M.; Jafari, M.T.; Amooshahi, M.M. Sol-gel/nanoclay composite as a sorbent for microextraction in packed syringe combined with corona discharge ionization ion mobility spectrometry for the determination of diazinon in water samples. J. Sep. Sci. 2018, 41, 493–500. [Google Scholar] [CrossRef]
  167. Moradi, E.; Mehrani, Z.; Ebrahimzadeh, H. Gelatin/sodium triphosphate hydrogel electrospun nanofiber mat as a novel nanosorbent for microextraction in packed syringe of La3+ and Tb3+ ions prior to their determination by ICP-OES. React. Funct. Polym. 2020, 153, 104627. [Google Scholar] [CrossRef]
  168. Taghani, A.; Goudarzi, N.; Bagherian, G.A.; Arab Chamjangali, M.; Amin, A.H. Application of nanoperlite as a new natural sorbent in the preconcentration of three organophosphorus pesticides by microextraction in packed syringe coupled with gas chromatography and mass spectrometry. J. Sep. Sci. 2018, 41, 2245–2252. [Google Scholar] [CrossRef]
  169. Matin, P.; Ayazi, Z.; Jamshidi-Ghaleh, K. Montmorillonite reinforced polystyrene nanocomposite supported on cellulose as a novel layered sorbent for microextraction by packed sorbent for determination of fluoxetine followed by spectrofluorimetry based on multivariate optimisation. Int. J. Environ. Anal. Chem. 2020, 1–16. [Google Scholar] [CrossRef]
  170. Arcoleo, A.; Bianchi, F.; Careri, M. A sensitive microextraction by packed sorbent-gas chromatography-mass spectrometry method for the assessment of polycyclic aromatic hydrocarbons contamination in Antarctic surface snow. Chemosphere 2021, 282, 131082. [Google Scholar] [CrossRef]
  171. Amiri, A.; Chahkandi, M.; Targhoo, A. Synthesis of nano-hydroxyapatite sorbent for microextraction in packed syringe of phthalate esters in water samples. Anal. Chim. Acta 2017, 950, 64–70. [Google Scholar] [CrossRef]
  172. Vera, J.; Fernandes, V.C.; Correia-Sá, L.; Mansilha, C.; Delerue-Matos, C.; Domingues, V.F. Occurrence of Selected Known or Suspected Endocrine-Disrupting Pesticides in Portuguese Surface Waters Using SPME-GC-IT/MS. Separations 2021, 8, 81. [Google Scholar] [CrossRef]
  173. Saliu, F.; Montano, S.; Hoeksema, B.W.; Lasagni, M.; Galli, P. A non-lethal SPME-LC/MS method for the analysis of plastic-associated contaminants in coral reef invertebrates. Anal. Methods 2020, 12, 1935–1942. [Google Scholar] [CrossRef]
  174. Terzaghi, E.; Falakdin, P.; Fattore, E.; Di Guardo, A. Estimating temporal and spatial levels of PAHs in air using rain samples and SPME analysis: Feasibility evaluation in an urban scenario. Sci. Total Environ. 2021, 762, 144184. [Google Scholar] [CrossRef] [PubMed]
  175. El-Sheikh, A.H.; Al-Quse, R.W.; El-Barghouthi, M.I.; Al-Masri, F.S. Derivatization of 2-chlorophenol with 4-amino-anti-pyrine: A novel method for improving the selectivity of molecularly imprinted solid phase extraction of 2-chlorophenol from water. Talanta 2010, 83, 667–673. [Google Scholar] [CrossRef] [PubMed]
  176. Anirudhan, T.S.; Alexander, S. Multiwalled carbon nanotube based molecular imprinted polymer for trace determination of 2,4-dichlorophenoxyaceticacid in natural water samples using a potentiometric method. Appl. Surf. Sci. 2014, 303, 180–186. [Google Scholar] [CrossRef]
  177. Shaikh, H.; Memon, N.; Bhanger, M.I.; Nizamani, S.M.; Denizli, A. Core-shell molecularly imprinted polymer-based solid-phase microextraction fiber for ultra trace analysis of endosulfan I and II in real aqueous matrix through gas chromatography-micro electron capture detector. J. Chromatogr. A 2014, 1337, 179–187. [Google Scholar] [CrossRef] [PubMed]
  178. Gao, X.; Pan, M.; Fang, G.; Jing, W.; He, S.; Wang, S. An ionic liquid modified dummy molecularly imprinted polymer as a solid-phase extraction material for the simultaneous determination of nine organochlorine pesticides in environmental and food samples. Anal. Methods 2013, 5, 6128–6134. [Google Scholar] [CrossRef]
  179. Zhao, F.; Wang, S.; She, Y.; Zhang, C.; Zheng, L.; Jin, M.; Shao, H.; Jin, F.; Du, X.; Wang, J. Subcritical water extraction combined with molecular imprinting technology for sample preparation in the detection of triazine herbicides. J. Chromatogr. A 2017, 1515, 17–22. [Google Scholar] [CrossRef]
  180. Stoeckelhuber, M.; Müller, C.; Vetter, F.; Mingo, V.; Lötters, S.; Wagner, N.; Bracher, F. Determination of Pesticides Adsorbed on Arthropods and Gastropods by a Micro-QuEChERS Approach and GC–MS/MS. Chromatographia 2017, 80, 825–829. [Google Scholar] [CrossRef]
  181. Kurth, D.; Krauss, M.; Schulze, T.; Brack, W. Measuring the internal concentration of volatile organic compounds in small organisms using micro-QuEChERS coupled to LVI-GC-MS/MS. Anal. Bioanal. Chem. 2017, 409, 6041–6052. [Google Scholar] [CrossRef]
  182. Mehravar, A.; Feizbakhsh, A.; Sarafi, A.H.M.; Konoz, E.; Faraji, H. Deep eutectic solvent-based headspace single-drop microextraction of polycyclic aromatic hydrocarbons in aqueous samples. J. Chromatogr. A 2020, 1632, 461618. [Google Scholar] [CrossRef]
  183. Kiszkiel-Taudul, I.; Starczewska, B. Single drop microextraction coupled with liquid chromatography-tandem mass spectrometry (SDME-LC-MS/MS) for determination of ranitidine in water samples. Microchem. J. 2019, 145, 936–941. [Google Scholar] [CrossRef]
  184. Nunes, L.S.; Korn, M.G.A.; Lemos, V.A. A novel direct-immersion single-drop microextraction combined with digital colorimetry applied to the determination of vanadium in water. Talanta 2021, 224, 121893. [Google Scholar] [CrossRef] [PubMed]
  185. Werner, J. Low Density Ionic Liquid-Based Ultrasound-Assisted Dispersive Liquid–Liquid Microextraction for the Preconcentration of Trace Aromatic Amines in Waters. J. Anal. Chem. 2021, 76, 1182–1188. [Google Scholar] [CrossRef]
  186. Yang, S.; Liu, H.; Hu, K.; Deng, Q.; Wen, X. Investigation of thermospray flame furnace atomic absorption spectrometric determination of cadmium combined with ultrasound-assisted dispersive liquid-liquid microextraction. Int. J. Environ. Anal. Chem. 2022, 102, 443–455. [Google Scholar] [CrossRef]
  187. Ali, J.; Tuzen, M.; Citak, D.; Uluozlu, O.D.; Mendil, D.; Kazi, T.G.; Afridi, H.I. Separation and preconcentration of trivalent chromium in environmental waters by using deep eutectic solvent with ultrasound-assisted based dispersive liquid-liquid microextraction method. J. Mol. Liq. 2019, 291, 111299. [Google Scholar] [CrossRef]
  188. Shojaei, S.; Shojaei, S.; Nouri, A.; Baharinikoo, L. Application of chemometrics for modeling and optimization of ultrasound-assisted dispersive liquid–liquid microextraction for the simultaneous determination of dyes. npj Clean Water 2021, 4, 23. [Google Scholar] [CrossRef]
  189. Xizhi, S.; Sun, A.-l.; Wang, Q.-h.; Hengel, M.; Shibamoto, T. Rapid Multi-Residue Analysis of Herbicides with Endocrine-Disrupting Properties in Environmental Water Samples Using Ultrasound-Assisted Dispersive Liquid–Liquid Microextraction and Gas Chromatography–Mass Spectrometry. Chromatographia 2018, 81, 1071–1083. [Google Scholar] [CrossRef]
  190. Wang, X.M.; Du, T.T.; Wang, J.; Kou, H.X.; Du, X.Z. Determination of polybrominated biphenyls in environmental water samples by ultrasound-assisted dispersive liquid-liquid microextraction followed by high-performance liquid chromatography. Microchem. J. 2019, 148, 85–91. [Google Scholar] [CrossRef]
  191. Xinya, L.; Liu, C.; Qian, H.; Qu, Y.; Zhang, S.; Lu, R.; Gao, H.; Zhou, W. Ultrasound-assisted dispersive liquid-liquid microextraction based on a hydrophobic deep eutectic solvent for the preconcentration of pyrethroid insecticides prior to determination by high-performance liquid chromatography. Microchem. J. 2019, 146, 614–621. [Google Scholar] [CrossRef]
  192. Rahimi Moghadam, M.; Zargar, B.; Rastegarzadeh, S. Determination of Tetracycline Using Ultrasound-Assisted Dispersive Liquid-Liquid Microextraction Based on Solidification of Floating Organic Droplet Followed by HPLC-UV System. J. AOAC Int. 2021, 104, 999–1004. [Google Scholar] [CrossRef]
  193. Fayaz, S.M.; Abdoli, M.A.; Baghdadi, M.; Karbasi, A. Ag removal from e-waste using supercritical fluid: Improving efficiency and selectivity. Int. J. Environ. Stud. 2021, 78, 459–473. [Google Scholar] [CrossRef]
  194. Falsafi, Z.; Raofie, F.; Kazemi, H.; Ariya, P.A. Simultaneous extraction and fractionation of petroleum biomarkers from tar balls and crude oils using a two-step sequential supercritical fluid extraction. Mar. Pollut. Bull. 2020, 159, 111484. [Google Scholar] [CrossRef] [PubMed]
  195. Meskar, M.; Sartaj, M.; Infante Sedano, J.A. Assessment and comparison of PHCs removal from three types of soils (sand, silt loam and clay) using supercritical fluid extraction. Environ. Technol. 2019, 40, 3040–3053. [Google Scholar] [CrossRef] [PubMed]
  196. Tita, G.J.; Navarrete, A.; Martin, A.; Cocero, M.J. Model assisted supercritical fluid extraction and fractionation of added-value products from tobacco scrap. J. Supercrit. Fluids 2021, 167, 105046. [Google Scholar] [CrossRef]
  197. Lie, J.; Tanda, S.; Liu, J.-C. Subcritical Water Extraction of Valuable Metals from Spent Lithium-Ion Batteries. Molecules 2020, 25, 2166. [Google Scholar] [CrossRef]
  198. Taki, G.; Islam, M.N.; Park, S.-J.; Park, J.-H. Optimization of operating parameters to remove and recover crude oil from contaminated soil using subcritical water extraction process. Environ. Eng. Res. 2018, 23, 175–180. [Google Scholar] [CrossRef]
  199. Kang, S.J.; Sun, Y.H.; Qiao, M.Y.; Li, S.L.; Deng, S.H.; Guo, W.; Li, J.S.; He, W.T. The enhancement on oil shale extraction of FeCl3 catalyst in subcritical water. Energy 2022, 238, 121763. [Google Scholar] [CrossRef]
  200. Zohar, M.; Matzrafi, M.; Abu-Nassar, J.; Khoury, O.; Gaur, R.Z.; Posmanik, R. Subcritical water extraction as a circular economy approach to recover energy and agrochemicals from sewage sludge. J. Environ. Manag. 2021, 285, 112111. [Google Scholar] [CrossRef]
Figure 1. Different steps involved in sample preparation.
Figure 1. Different steps involved in sample preparation.
Molecules 27 02953 g001
Figure 2. Different sorbent-based GreETs used in several fields of analysis. Legend: DI-SPME: solid-phase microextraction in direct immersion mode; FPSE: fabric-phase solvent extraction; GreETs: green extraction techniques; HS-SPME: solid-phase microextraction in headspace mode; MEPS: microextraction in packed sorbent; MSAμE: multisphere adsorptive microextraction; MSPD: matrix solid-phase dispersion; MSPE: magnetic solid-phase extraction; SBSE: stir-bar sorbent extraction; SPDE: solid-phase dynamic extraction; SPE: solid-phase extraction; SPME: solid-phase microextraction; μSPE: micro-solid-phase extraction.
Figure 2. Different sorbent-based GreETs used in several fields of analysis. Legend: DI-SPME: solid-phase microextraction in direct immersion mode; FPSE: fabric-phase solvent extraction; GreETs: green extraction techniques; HS-SPME: solid-phase microextraction in headspace mode; MEPS: microextraction in packed sorbent; MSAμE: multisphere adsorptive microextraction; MSPD: matrix solid-phase dispersion; MSPE: magnetic solid-phase extraction; SBSE: stir-bar sorbent extraction; SPDE: solid-phase dynamic extraction; SPE: solid-phase extraction; SPME: solid-phase microextraction; μSPE: micro-solid-phase extraction.
Molecules 27 02953 g002
Figure 3. Advantages and schematic overview of µSPEed.
Figure 3. Advantages and schematic overview of µSPEed.
Molecules 27 02953 g003
Figure 4. Different liquid-based GreETs used in several fields of analysis. Legend: DLLME: dispersive liquid–liquid microextraction; GreETs: green extraction techniques; HF-LPME: hollow fiber liquid-phase microextraction, SDME: single-drop microextraction; SFOME: solidification of floating organic drop microextraction; UABE: ultrasound-assisted back extraction; µQuEChERS: micro-QuEChERS.
Figure 4. Different liquid-based GreETs used in several fields of analysis. Legend: DLLME: dispersive liquid–liquid microextraction; GreETs: green extraction techniques; HF-LPME: hollow fiber liquid-phase microextraction, SDME: single-drop microextraction; SFOME: solidification of floating organic drop microextraction; UABE: ultrasound-assisted back extraction; µQuEChERS: micro-QuEChERS.
Molecules 27 02953 g004
Figure 5. Schematic representation of direct immersion and headspace modes of single-drop microextraction (SDME) (a) and two- and three-phase modes of hollow fiber liquid-phase microextraction (HF-LPME) (b).
Figure 5. Schematic representation of direct immersion and headspace modes of single-drop microextraction (SDME) (a) and two- and three-phase modes of hollow fiber liquid-phase microextraction (HF-LPME) (b).
Molecules 27 02953 g005
Figure 6. Different steps involved in sample preparation. GreETs: green extraction techniques.
Figure 6. Different steps involved in sample preparation. GreETs: green extraction techniques.
Molecules 27 02953 g006
Table 1. Advantages and disadvantages of some GreETs commonly used in the analysis of biological, food, and environmental matrices.
Table 1. Advantages and disadvantages of some GreETs commonly used in the analysis of biological, food, and environmental matrices.
Extraction ProcedureAdvantagesDisadvantages
SPMEAlternative to SPE
A limited number of steps
Reduced sample amount
Reuse of the polymeric phase
Environmentally friendly
Short extraction time
Potential contamination of the SPME needle
µSPEAlternative of LLE
Simplicity of automation
Suitable for large scale
Low sorbent
Low solvent volume
Requires stirring
Possibility of low recoveries
MEPSLow solvent volume
Low sample amount
Fast and easy to use
Economical
Fully automated for online procedure
Requires a wide range of optimization steps
MSPEEnvironmentally friendly
A limited number of steps
Low amount of sorbent material
Reuse of sorbent material
Short extraction time
Requires vortex/shaker/magnetic stirrerSelection of suitable sorbent
MSDPEnvironmentally friendly
A limited number of steps
Quick
Simple
Requires anhydrous sorbents activated at high temperatures
FPSEEfficient
Fast extraction
Low volume of solvents
High preconcentration factor
Low sorbent capacity
Long sample preparation time
DLLMEEconomical
High recovery
Low sample amount
Low extraction time
Low solvent volume
Low selectivity
Requires centrifugation
SFOMEEnvironmentally friendly
High enrichment factor economical
Low volume of solvents
Simplicity of automation
Requires a wide range of optimization steps
μQuEChERSEconomical
Efficient clean-up by dSPE
Low solvent consumption
Labor intensive
Difficult to automate
Emulsion formation
SFEEnvironmentally friendly
No required solvents
Low operating temperatures (40–80 °C)
Fast and high yield
Very expensive
Complex equipment operating at high pressures
High power consumption
Legend: DLLME: dispersive liquid–liquid microextraction; FPSE: fabric-phase sorptive extraction; MEPS: microextraction in packed sorbent; MSPD: matrix solid-phase dispersion; MSPE: magnetic solid-phase extraction; SFE: supercritical fluid extraction; SFOME: solidification of floating organic drop microextraction; SPME: solid-phase microextraction; µQuEChERS: micro-QuEChERS; µSPE: micro-solid-phase extraction.
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Câmara, J.S.; Perestrelo, R.; Berenguer, C.V.; Andrade, C.F.P.; Gomes, T.M.; Olayanju, B.; Kabir, A.; M. R. Rocha, C.; Teixeira, J.A.; Pereira, J.A.M. Green Extraction Techniques as Advanced Sample Preparation Approaches in Biological, Food, and Environmental Matrices: A Review. Molecules 2022, 27, 2953. https://doi.org/10.3390/molecules27092953

AMA Style

Câmara JS, Perestrelo R, Berenguer CV, Andrade CFP, Gomes TM, Olayanju B, Kabir A, M. R. Rocha C, Teixeira JA, Pereira JAM. Green Extraction Techniques as Advanced Sample Preparation Approaches in Biological, Food, and Environmental Matrices: A Review. Molecules. 2022; 27(9):2953. https://doi.org/10.3390/molecules27092953

Chicago/Turabian Style

Câmara, José S., Rosa Perestrelo, Cristina V. Berenguer, Carolina F. P. Andrade, Telma M. Gomes, Basit Olayanju, Abuzar Kabir, Cristina M. R. Rocha, José António Teixeira, and Jorge A. M. Pereira. 2022. "Green Extraction Techniques as Advanced Sample Preparation Approaches in Biological, Food, and Environmental Matrices: A Review" Molecules 27, no. 9: 2953. https://doi.org/10.3390/molecules27092953

Article Metrics

Back to TopTop