Next Article in Journal
Ginsenoside Rh1 Inhibits Angiotensin II-Induced Vascular Smooth Muscle Cell Migration and Proliferation through Suppression of the ROS-Mediated ERK1/2/p90RSK/KLF4 Signaling Pathway
Previous Article in Journal
Hyperoxia Induces Ferroptosis and Impairs Lung Development in Neonatal Mice
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Assessment of Methodological Pipelines for the Determination of Isothiocyanates Derived from Natural Sources

by
Sotiris Kyriakou
1,
Dimitrios T. Trafalis
2,
Maria V. Deligiorgi
2,
Rodrigo Franco
3,4,
Aglaia Pappa
5 and
Mihalis I. Panayiotidis
1,*
1
Department of Cancer Genetics, Therapeutics & Ultrastructural Pathology, The Cyprus Institute of Neurology & Genetics, Ayios Dometios, Nicosia 2371, Cyprus
2
Laboratory of Pharmacology, Medical School, National & Kapodistrian University of Athens, 11527 Athens, Greece
3
Redox Biology Centre, University of Nebraska-Lincoln, Lincoln, NE 68583, USA
4
Department of Veterinary Medicine & Biomedical Sciences, University of Nebraska-Lincoln, Lincoln, NE 68583, USA
5
Department of Molecular Biology & Genetics, Democritus University of Thrace, 68100 Alexandroupolis, Greece
*
Author to whom correspondence should be addressed.
Antioxidants 2022, 11(4), 642; https://doi.org/10.3390/antiox11040642
Submission received: 31 January 2022 / Revised: 17 March 2022 / Accepted: 22 March 2022 / Published: 27 March 2022

Abstract

:
Isothiocyanates are biologically active secondary metabolites liberated via enzymatic hydrolysis of their sulfur enriched precursors, glucosinolates, upon tissue plant disruption. The importance of this class of compounds lies in their capacity to induce anti-cancer, anti-microbial, anti-inflammatory, neuroprotective, and other bioactive properties. As such, their isolation from natural sources is of utmost importance. In this review article, an extensive examination of the various parameters (hydrolysis, extraction, and quantification) affecting the isolation of isothiocyanates from naturally-derived sources is presented. Overall, the effective isolation/extraction and quantification of isothiocyanate is strongly associated with their chemical and physicochemical properties, such as polarity-solubility as well as thermal and acidic stability. Furthermore, the successful activation of myrosinase appears to be a major factor affecting the conversion of glucosinolates into active isothiocyanates.

1. Introduction

Naturally occurring plant-derived chemical compounds, known as phytochemicals, have been the subject of intense research due to their important role in health promotion and disease prevention [1,2,3,4,5,6,7,8,9,10]. To these ends, numerous epidemiological studies have documented a low consumption of fruits and vegetables conception with a dramatic increase in the rate of cancer development and progression, among other diseases (e.g., cardiovascular disease, metabolic syndrome, etc.) [11,12,13,14,15,16,17,18,19,20,21]. Finally, the role of diet in the prevention of neurodegenerative diseases (e.g., Dementia, Alzheimer’s, Parkinson’s, etc.) has been extensively reviewed by emphasizing on the role of various phytochemicals (e.g., omega-3 fatty acids, flavonoids, vitamins, etc.) by exerting a plethora of disease-preventing biological properties [22,23,24,25].
Isothiocyanates (ITCs) are a class of phytochemicals with exceptional biological and nutritional activity. These are aliphatic or aromatic phytochemicals produced in high abundance in cruciferous plants, including broccoli, kale, cauliflower, cabbage, Brussel sprouts, wasabi roots, and watercress. They all have the same chemical backbone, R-N=C=S (where R- can be either an aliphatic or aromatic group; Figure 1), whereas their content varies across species, variety, and growing conditions (Table 1) [26,27].
So far, more than 154 different glucosinolates (GSLs) have been identified and these can be categorised into 4 classes: aliphatic, aromatic, indolyl, and glycosylated [40,41,42,43,44]. The breakdown product of the hydrolysis of GSLs includes a variety of nitrogenous compounds such as ITCs, thiocyanates, nitriles, and epithionitriles, depending on the reaction conditions (i.e., pH, temperature and reducing agents like ferrous ions, ascorbate, etc.) [45,46,47,48,49,50,51,52,53,54]. For instance, thiohydroximate-O-sulphate, at neutral pH, is rearranged into ITCs whereas at low pH it can be rearranged into nitriles (Figure 2) [55,56,57,58,59,60,61]. Finally, the presence of epithiospecifier protein-like (ESP) factors can drive the reaction towards the formation of nitriles as well. In addition, ESPs have the capacity to convert alkenyl-contained intermediates into epithionitriles which are considerably less stable products [59,61,62,63]. In all cases, the rearrangement of thiohydroximate-O-sulphate into its active metabolites is accompanied by loss of a sulphate group.
The importance of this class of molecules relies on their anticancer properties by downgrading the factors that are responsible for the cellular detoxification [50,64]. More specifically, ITCs can form an adduct with glutathione (GSH) during the mercapturic acid pathway, causing GSH depletion [65,66,67]. As a result, the levels of reactive oxygen species (ROS) are elevated, leading to down-stream activation of caspases and other apoptotic markers, thereby promoting the induction of apoptosis [68,69,70,71]. It has been postulated, throughout the literature, that ITCs can also induce cell cycle growth arrest by down- and/or up-regulating various cycle-dependent kinases (CDKs) and tumour suppressor genes, respectively [72,73,74]. Moreover, their anti-angiogenic capacity has also been suggested, primarily through their ability to down-regulate both vascular endothelial as well as epithelial growth factors [75,76,77]. Finally, they have been shown to affect the epigenome through their capacity to regulate the expression levels of histone deacetylases, acetyltransferases, and methyltransferases, among other epigenetic events (Figure 3) [77,78,79]. In another study, nanoformulation of ITCs facilitates their solubility and stability, leading to an improved bioavailability and anticancer capacity [80]. The nanoformulation of ITCs was acomplished via their encapsulation with either organic (micelle, liposome, dendrimer, polymeric nanoparticles, solid lipoid nanoparticle, and carbon nantube) or inorganic (gold, iron oxide or silicon) nanocarriers [80].
In conclusion, this review article aims to provide an integrated optimum methodological approach towards the isolation and quantification of different classes of ITCs. To this end, the proposed experimental pipeline provides fast and easy access to the isolation and quantification of major ITCs (SFN, IBN, AITC, BITC, and PEITC) from naturally-derived sources.

2. Determination of ITC Content

A general protocol that includes the experimental pipeline for the determination of ITCs content from naturally-derived sources is divided in four pillars, namely (i) sample storage and isolation, (ii) hydrolysis of GSLs, (iii) extraction of ITCs, and eventually (iv) their quantification (Figure 4).

2.1. Sample Processing and Storage

Sample preparation prior to analyte identification and quantification, using analytical techniques, is a very crucial step in preventing against analyte loss and/or its decomposition [81,82]. More specifically, cellular decomposition of cruciferous plants (by chopping, boiling, etc.) leads to activation of myrosinase and hence the hydrolysis of GSLs. Therefore, it is recommended that small cruciferous plants or parts of the plant are freeze-dried (at −20 °C) and stored under solid carbon dioxide conditions (−80 °C), whereas larger plants should be immediately frozen under liquid nitrogen to avoid the activation of myrosinase [81,82,83,84].

2.2. Hydrolysis of GSLs

2.2.1. Effect of pH

Hydrolysis of GSLs comprises a major role in determining the content of ITCs in cruciferous vegetables (Table 2).
This can be attributed to the fact that different parameters (such as temperature and pH) can affect the hydrolysis of GSLs, leading to their partial or complete conversion into ITCs and/or other nitrogenous-contained metabolites [109]. In previous studies, it has been demonstrated that different pH environments can affect the (i) formation of different intermediates (e.g., ITCs, nitriles), (ii) purity, and (iii) stability of the final product [60,110]. For example, Vaughn S.F and Berhow M.A. (2005) have shown that hydrolysis of the seeds from the plants Eruca sativa var. sativa (Aragula) and Erusimum allionii (Siberian wallflower) (both being sources of erucin) leads to formation of different products/mixtures (regardless of both plants containing the same GSL precursor), under a range of pH values [85]. Specifically, upon exposure of Eruca sativa var. sativa to phosphate buffer (at pH 7.0), a mixture of erucin and erucin nitrile was formed, while the GSL precursor of Erusimum allionii, under the same conditions, formed erucin, cheirolin, eysolin, and sulforaphane (SFN) [85]. On the other hand, at basic conditions (pH 10.0) both plants produced erucin, while at acidic conditions (pH 1.0), Eruca sativa var. sativa’s GSL content was completely converted into erucin nitrile [85] and Erusimum allionii’s into a mixture of erucin nitrile, sulforaphane nitrile, and erysolin nitrile [86,110]. Moreover, during the same study, exposure of Brassica hirta (L.) Moench (white mustard), Lesquerella fendleri (Lesquerella), and Mattiola longipetala (night-scented stock) at pH 7.0 and 10.0 led to complete hydrolysis of their respective GSLs into 4-hydroxybenzyl ITC, iberin (IBN), and SFN, respectively [85]. Similarly, exposure of Capsella bursa-pastoris (L.) Medik (Shepher’s purse), Lobularia maritima (L.) Desv. (Sweet alyssum), and Erysimum cheiri (L.) Crantz (English wallflower) at pH 7.0 and 10.0, 3-butenyl ITC, cheroline, and lesquerellin were produced as the maim hydrolysis products, whereas none of the GSL precursors were hydrolysed at pH 1.0 [85].
In another study, when hydrolysis of gluconasturtiin (obtained from watercress), under a range of pH values (e.g., 5.0, 7.0, and 9.0) occurred, it was shown that the optimum pH for its conversion into phenethyl ITC (PEITC) was at 7.0 in phosphate buffer [87,88]. In similar studies, hydrolysis of glucoraphanin (derived from Brassica oleracea var. italica; broccoli) using a citrate/phosphate buffer (at pH 6.5) led to complete conversion into SFN [89]. Furthermore, in studies conducted by other groups, different species of Brassica oleracea were utilised (as a source of glucoraphanin) in order to examine the effect of different pH values (e.g., 5.0, 5.5, 6.0, 6.5, 7.0, 7.5, and 8.0) into the bio-production of SFN [88,89,90]. The outcome of these studies demonstrated that although hydrolysis of glucoraphanin was achieved, under all pH values, the content of SFN varied as it was pH-dependent [88,90,91]. Finally, the authors have concluded that pH 7.0 is the optimum condition for myrosinase activity during hydrolysis of sinigrin from Armoracia rusticana (horseradish), a finding consistent with the relevant bibliography [92,93].
Finally, in another study, hydrolysis of glucotropaeolin (obtained either from Carica papaya (L.) (papaya) or Tropaeolum majus (L.) tropaeolaceae (garden nasturtium) or Lepidium sativum (L.) (garden cress)) yielded benzyl ITC (BEITC), at pH 6.5–7.0, whereas at pH 1.0 there was formation of benzyl nitrile with traces of BEITC [94,95,96,97]. Coscueta et al. (2020) observed that production of PEITC occurred at almost quantitative levels when Nasturtium officinale (watercress) was exposed at pH 7.0. At acidic (pH 5.0) or basic (pH 10.0) conditions, the activity of myrosinase was decreased, leading to the formation of either a mixture of PEITC/phenethyl nitrile or phenethyl nitrile, respectively [98]. These findings are in agreement with the studies by Hancshen et al. (2017) and Hanschen and Schreiner (2017), documenting that the hydrolysis of ITCs is enhanced upon exposure to neutral or slightly basic conditions [56,58], as under these experimental conditions the kinetic properties of myrosinase are accelerated, thus driving the reaction towards the formation of ITCs [42,99,100]. On the contrary, under acidic conditions, the reaction’s kinetics can change as this either reduces the activity of myrosinase (i.e., the hydrolysis rate is decreased) or drives the reaction towards complete conversion of GSLs into their respective nitriles. Interestingly, the majority of studies have shown that in slightly acidic pH (1.0 < pH < 7.0), GSL hydrolysis leads to the formation of a mixture of both ITCs and nitriles [61,101].

2.2.2. Effect of Temperature

Another factor of great importance, for the conversion of GSLs into ITCs, is the temperature of the hydrolysis reaction, as it can affect the activity of myrosinase and/or the stability of ITCs [111]. Earlier studies have suggested that myrosinase, extracted from broccoli, is stable at temperatures below 40 °C, whereas at 50 °C the activity decreased by 70% within 5 min of exposure [112]. In contrast, another study has identified high levels of SFN extracts in broccoli after treating the vegetable at 60 °C for 10 min, thereby demonstrating that the myrosinase can be remained active even at 60 °C [113]. In any case, various studies have aimed to characterize the relationship between myrosinase activity and temperature towards conversation of GSLs into ITCs in the Brassica family [114,115,116].
On this note, the heat stability of white (Brassica oleracea var capitate) and red (Brassica oleracea var capitata f. rubra) cabbages was examined at 70 °C for 30 min, demonstrating reduced production of AITC due to thermal degradation of myrosinase [117,118]. In addition, other studies have shown that 30 °C and 50 °C were optimum temperatures for myrosinase activity in broccoli and Brussels sprouts, respectively [111,119]. Furthermore, hydrolysis of 2-propyl and 3-butyl GSLs into their respective ITCs, in Arabidopsis thaliana (L.) Heynth, was higher at 37 °C when compared to other temperature ranges [120]. In another study, the authors examined the stability of the respective enzymes obtained from yellow, brown, and black mustard seeds, showing that their stability tolerance was limited up to 60 °C (in all three species) and that the activity of myrosinase was significantly decreased above 60 °C [112]. In particular, the myrosinase activity from yellow mustard was only stable up to 50 °C, while it lost about 79% of its activity at 60 °C. Moreover, the myrosinase activity in black and brown mustards, at 70 °C, was significantly reduced to 41% and 65%, respectively, whereas in yellow mustard it was completely abolished [112]. The fact that some enzymatic activity has been recorded at temperatures above 85 °C (as small traces of glucosinolate degradation products) could be attributed to the presence of thermostable desulphatase enzymes, which are activated at high temperatures and can catalyse the hydrolysis of thioglucosidic bonds [112]. Finally, these results were further supported by another study indicating that the inactivation of myrosinase in mustard species can occur in just 10 min of exposure at 60 °C [121,122] (Table 3).
Although temperature is a major factor that can affect the formation of ITCs from the hydrolysis of GSLs, it is less stringent than pH. It is suggested that this might be attributed to the different optimum temperatures of the different ESPs of various plants [122,123]. To this end, the presence of ESPs in Arabidopsis thaliana (L.) Heynh increases nitrile formation (at lower temperatures (e.g., 0–20 °C)), while myrosinase activity is reduced [123,124,125]. These findings are in contradiction with the results of other studies, indicating that, in Arabidopsis thaliana (L.) Heynh, a temperature of 37 °C possesses an inverse effect where the production of ITCs and nitriles is at higher rates. In addition, at 37 °C, the epithionitrile levels are reduced, implying loss of ESP activity [61,126].
In conclusion, myrosinase is a temperature sensitive enzyme and so thermal treatments (including sample heat processing and blanching) can cause a decrease in its activity and consequently in the production of ITCs. In addition, increased formation of ITCs is achieved if ESP activity is not affected and the ratio of glucosinolate-aglucon to ESP increases, thereby leading to lower EPT and higher ITCs formation [127].

2.2.3. Effect of EDTA and Ascorbic Acid

In addition to pH and temperature, the expression levels of ESPs and the presence of Fe2+ are secondary factors which can also affect the hydrolysis of GSLs and the production of ITCs. Previous studies have shown that acidic conditions in combination with high levels of ESPs and ferrous ions (Fe2+) can lead to hydrolysis reactions into yielding high levels of nitriles and thus undesired toxicity. In addition, activation of ESPs requires Fe2+, therefore decreasing the content of ferrous ion which could suppress the formation of nitriles and thus facilitate the production of ITCs. Various studies have suggested that the addition of ascorbic acid can stimulate the action of myrosinase by facilitating the hydrolysis of GSLs [128,129,130,131]. To this end, various attempts have been made in identifying the role of ferrous ions and ascorbic acid on myrosinase activity [132]. On another note, the addition of 10 mM of ethylenediaminetetraacetic acid (EDTA) (at pH 6.8) can lead to a high ITC conversion rate due to its chelating capability in binding and coordinating metal ions (magnesium; Mg2+, ferrous; Fe2+, ferric; Fe3+, etc.), thereby diminishing the presence of metal ions in enzymolytic solutions [128]. In addition, the formation of nitriles was shown to be reduced by decreasing the content of Fe2+, in Brussel sprouts, thus enhancing the production of ITCs [133,134]. Overall, the optimum EDTA concentration needed for complete inhibition of ESPs (hence maximal production of ITCs) is of utmost importance in modulating myrosinase activity [133,134].
On the other hand, ascorbic acid has been previously reported to function as a cofactor of myrosinase, thus its availability and concentration can affect its activity [135,136]. Specifically, excessive levels of ascorbic acid can inhibit myrosinase activity (thus less ITCs formation), whereas reduced levels might not lead to optimum levels of myrosinase activity [137]. To this end, it was suggested that the optimum concentration of ascorbic acid, for maximum myrosinase activity, was about 5% w/v. This can be attributed to ascorbic acid acting as an uncompetitive activator of myrosinase, although the exact mechanism of such activation is not entirely elucidated [136,137].

3. Extraction of Isothiocyanates

Due to their major biological properties, the effective extraction and isolation of ITCs is of utmost importance. This is because an optimum extraction method will allow the maximum recovery and thus a valid determination of ITCs content in cruciferous plants. However, a common extraction procedure for all ITCs cannot be achieved since each ITC has different physical properties, including polarity, volatility, and stability. For this purpose, different procedures have been applied, including liquid–liquid, solid phase, supercritical gas assisted, and ultrasonic extractions. In this section the extraction of the most utilized and abundant ITCs will be discussed, including SFN, IBN, AITC, BITC, and PEITC.

3.1. Extraction of Sulforaphane

The enzymatic hydrolysis of glucoraphane generates a variety of compounds (e.g., glucose, sulphates, ITCs, thiocyanates, nitriles, etc.) which interfere with the separation and determination of SFN [138,139,140]. Therefore, it is necessary to establish a simple and convenient method for the selective extraction and separation of SFN from various cruciferous vegetable sources, including broccoli. In most studies, the isolation of SFN is accompanied with the hydrolysis of its precursor GSL (i.e., glucoraphane) by treating the plant (or parts of the plant) with water, at pH 6.0–7.0, under various temperatures (e.g., 25–37 °C), followed by a further extraction step with a solvent of medium polarity such as dichloromethane [29,128,140,141,142,143,144,145,146,147,148,149,150], ethyl acetate [151,152], or chloroform [153,154]. In a few studies, liquid–liquid extraction was employed in order to extract SFN from various plant sources such as broccoli seeds after being defatted with light petroleum [155,156]. Then, hydrolysis of glucoraphane was performed in a solution mixture of dichloromethane: phosphate buffer containing phosphoric acid (H3PO4) (0.05 mol/mL) at pH 7.0 for 8 hr at 25 °C [155,157]. Afterwards, SFN extraction was performed by utilizing dichloromethane at room temperature and reconstituting the concentrated extracts in acetone prior to determination of the SFN content [155,156]. In some cases, it has been reported that a step prior to hydrolysis is added where glucoraphane is isolated and incubated with 2-(N-morpholino) ethanesulfonic instead of acidic water [148,158]. In a recent study, SFN isolation (from broccoli) was attempted by solid-phase extraction, with the aim to examine the different factors that can potentially affect the extraction process [139]. Initially, the separation capacity of three different cartridges (with different stationary phase), namely C18, silica, and amino, were evaluated, with the authors concluding that the ideal column was silica after elution with a range of solvents of different polarity (e.g., water, acetonitrile, dichloromethane, 0.1 M acetic acid, ethyl acetate, and hexane). This was attributed to the fact that SFN is a molecule of weak polarity and thus can easily be absorbed by a matrix of a similar polarity (e.g., silica) as opposed to C18 and amino cartridges [136]. Furthermore, the selection of a variety of different washing and elution solvents was evaluated indicating that ethyl acetate was the most appropriate one. This finding is in agreement with other studies, regarding SFN isolation from broccoli, using Solid Phase Extraction (SPE) [140,145,154,157,159,160] in combination with a solvent of medium polarity (e.g., dichloromethane) for achieving the highest rate of SFN recovery [139,161]. This finding comes into contradiction, with other studies claiming that the use of methanol as an elution solvent can lead to isolation of SFN with higher purity levels and recovery rates [162].
The last parameter to be investigated was the volume of the elution solvent. Various authors have suggested that increasing the elution volume of dichloromethane (from 2 to 6 mL) results in an increased amount of extracted SFN. However, in other studies, when the volume of dichloromethane exceeded 4 mL, the amount of extracted SFN remained almost constant [139,161]. Moreover, when 5 mL of methanol was utilized as an elution solvent, an SFN content of 94% purity was obtained [162]. Finally, the amount of elution solvents relies on many factors, including the quantity and viscosity of the analyte as well as the size of the cartridge.

3.2. Extraction of Iberin

The hydrolysis of glucoiberin yields IBN, which is in high abundance in the Brassicaceae and Capparaceae species and differs from SFN by one methylene unit in the aliphatic chain [163,164]. Despite its evidenced health benefits [165,166,167,168,169], very little is known about the extraction and purification of IBN, mainly due its chemical instability [170,171]. In a recent study, horseradish roots were used (as a source of IBN) as a raw material which was extracted overnight by maceration with ethyl acetate [34]. Afterwards, the non-soluble material was extracted with methanol, followed by ethyl acetate, and then washed with water. Finally, the highly lipophilic organic phase was sequentially chromatographed on various C18 cartridges [34]. Analysis of the fraction with ultra-violate (UV) absorbance at 242 nm, using high resolution electron spray ionization, demonstrated the presence of a single compound with a molecular formula of C5H9NOS2 and mass to charge ratio (m/z) equal to 164.0202 (in the positive mode) indicative of the successful extraction, isolation and purification of IBN [34]. In another study, an alternative way of IBN extraction was suggested, including an 8-hr shacking incubation of Lesquerella fendleri (L.) and Physaria fendleri seeds in dichloromethane (acidified with 0.1 M HCl) at room temperature followed by re-extraction of aqueous extracts with dichloromethane. Then, the dichloromethane extracts were sequentially washed with water and hexanes (in order to separate the hydrophobic and hydrophilic components), while total IBN content was determined with Gas Chromatography-Flame Ionization Detection (GC-FID) [85]. A post-incubation extraction procedure was previously suggested where a naturally-derived source of IBN was shaken-incubated, at room temperature, in de-ionized distilled water and then anhydrous magnesium sulfate and sodium chloride solutions were added [172]. Then, the mixture was extracted with dichloromethane and the resulting dichloromethane extracts were washed with hexane and water [172]. The yield of IBN being extracted by incubation with dichloromethane (acidified with 0.1 M HCl) was 90%, whereas the respective yield from the post-incubation extraction was only 48.6% [85,172]. The low yield of the post-incubation protocol can be credited to the instability of IBN in an aqueous environment [170].
Although IBN exerts structural similarities with SFN, it appears to be less stable in common solvents, thereby preventing its extraction using standard convectional extraction protocols [170,173,174]. For example, IBN is less stable in an aqueous environment rather than in methanol and ethanol. IBN’s instability, in alcohols, is due to its rearrangement into thiocarbamate or disubstituted thiourea when exposed to water or alcohol, respectively [53,170,173,175]. However, IBN remains relatively stable in acetonitrile. On the other hand, when raising the temperature of the column thermostat from 20 °C to 30 °C to 40 °C there appears to be a linear relation between temperature and instability of IBN. In addition, pH appears to be another crucial player that affects the stability of IBN as under alkaline conditions it can be decomposed, whereas an acidic environment considerably improves its stability [173,174,176,177]. Overall, given that the extraction and isolation of IBN is a trivial process (due to its instability), a possible approach to overcome this is by the substitution of water and alcohols with acetonitrile and the use of non-thermal techniques [170,171,178,179].

3.3. Extraction of Allyl Isothiocyanate

Among the various alkenyl ITCs, AITC (derived from the hydrolysis of sinigrin) is the most predominant one, characterized by its pungent odour and volatility. It is derived from cruciferous vegetables including Brassica juncea (L.) (Indian mustard), Wasabia jamonica (L.) matsum (wasabi), and Armoracia rusticana (horseradish) [180,181,182]. It has been shown to exert anticancer [183,184,185,186,187,188], anti-microbial [189,190,191], and anti-fungal activities [192,193,194,195], while also protecting against neurological disorders [196,197].
During the past decade, the extraction of AITC has been performed by either steam distillation or solvent extraction. In a study by Wu et al. (2009), the authors examined various solvents with different polarities such as cooking oil, petroleum, dichloromethane, diethyl ether, and ethyl acetate in an attempt to extract AITC from the hydrolysed dry roots of Armoracia rusticana [35]. The authors suggested that among these different extraction solvents, dichloromethane was the most effective one. In addition, a water extraction step was also added, according to which the dry (powdered) roots of Armoracia rusticana were mixed with acidified water (pH 4.0) for 24 hr at 20–40 °C. This step appears to maximize the extraction rate before final extraction with dichloromethane [35], a finding which is in agreement with the work of others [198,199,200,201]. Overall, hydrodistillation with dichloromethane is a more effective extraction method as AITC content is considerably higher than water extraction alone [35]. However, its major drawback is that all heat-liable volatile compounds (including AITC) can be decomposed and hydrolysed, thus leading to loss of sample quality [202,203]. Additionally, the utilization of heat (either for extraction purposes or for solvent evaporation) can induce loss of volatile compounds, thus minimizing the efficacy of the extraction process [202,203,204]. Finally, prolonged evaporation, under reduced pressure, can form traces of organic solvents in the final sample, thereby masking the abundance of the analyte [205].
In order to overcome loss of volatiles, an alternative extraction technology has been employed, called supercritical fluid extraction (SFE), according to which the extractant is separated directly from a mixture matrix using fluids above their critical values (supercritical fluids) [206,207,208,209]. SFE can be carried out at moderate temperatures, thus preventing thermal disintegration and loss of volatile compounds in addition to having no traces of organic solvents found in the final mixture [204,205]. Extraction of AITC was performed using supercritical CO2, with ethanol as a co-solvent, but because of the instability of ITCs in alcohols the study was discontinued [210]. Several attempts were made by other groups (e.g., by employing supercritical CO2 without any co-solvent or supercritical CO2 with either hexane, diethyl ether, ethyl acetate, and acetone) in order to minimize the instability of AITC. It was confirmed that the decomposition of AITC was prevented when hexane used as a co-solvent [35,37,201]. Li et al. (2010) identified moisture, temperature, and pressure of the gas as the critical factors to be considered for optimal SPE extraction of AITC [37]. Specifically, the authors reported that increasing the percentage of water improved the rate of AITC’s hydrolysis and hence its yield production and extraction [37]. However, excess of moisture can lead to degradation of the produced AITC into allyl allydithiocarbamate, which can further degrade into diallyl tetra- and penta-sulfide [37,199]. However, the excess of water on the SPE system can induce destruction on the porous morphology of the freeze-dried plant [37,202]. Therefore, it was concluded that the optimal moisture content of AITC (from wasabi), for SPE, should be maintained at 125% [37]. On the other hand, the same study has suggested that the temperature should be kept constant at 35 °C, since at this point the highest yield was noticed (408 mg AITC/100 g of wasabi roots) whereas at higher temperatures, 55 °C, the corresponding yield was lower (245 mg AITC/100 g of wasabi roots) [37]. This effect is likely due to the increase of AITC’s solubility as a result of the increased density of the supercritical CO2 [37]. Moreover, at lower temperatures no results were observed, probably due to the inactivation of myrosinase [37]. A similar effect was also reported in the case of the pressure of supercritical CO2 as the yield of extraction was higher under a pressure of 20 MPa and a temperature of 35 °C (104 mg of AITC/100 g of wasabi roots) when compared to a pressure of 25 MPa and a temperature of 35 °C (89 mg of AITC/100 g of wasabi roots) [37]. These findings are in agreement with those of other groups utilizing horseradish as the natural source of AITC (instead of wasabi) [211,212]. Eventually, Hu et al. (2009) claimed that “despite the fact that the SPE-CO2 extraction requires higher investment in equipment, is more efficient and less energy consuming and produces AITC of better quality” [35].
In an attempt to minimize the risk of AITC’s degradation, the extraction and isolation of sinigrin rather than AITC itself was utilized. In such studies, extraction with boiling water [48,209] or an aqueous–organic solvent system [213] was employed. To these ends, Powel et al. (2005) employed an optimized solid liquid protocol, taking into consideration the solvent composition, temperature, and pH [214]. In addition, Mohn et al. (2007) optimized critical parameters of pressurized liquid extraction, including solvent composition, particle size, temperature, and number of required extraction steps [215]. Furthermore, other research groups have focused on the extraction of sinigrin using ultrasonic-stimulated treatment [216]. This extraction technique relies on the disruption of cell wall promoting reduction in the particle-size and thus the mass transfer of the cell is enhanced due to the collapse of the cavitation bubble [217]. This technique is widely used in the separation and extraction of bioactive compounds, derived from natural sources, in the industrial scale [218,219,220]. Its major advantages involve improved extraction productivity, reduction of solvents, and overall protection of volatile and unstable phytochemicals [220,221,222]. In the case of sinigrin, it is recommended that the optimum condition for its maximal recovery is 57% ethanol, at 81 °C for 60 min, as under these conditions its yield increases by 70.67% when compared to other convectional extraction methodologies [216].

3.4. Extraction of Benzyl Isothiocyanate

BITC is the simplest form of benzylic ITC and it can be formed upon hydrolysis of its GSL precursor called Glucotropaeolin [223]. It is of high abundance in papaya seeds [39], garden [224], as well as Indian cress [225], and despite its simplicity in its chemical structure, its biological importance has been demonstrated in several studies [194,226,227,228,229,230,231,232]. Due to its numerous biological applications, its extraction has been challenging (like all ITCs) as it is thermally unstable and prone to degradation upon exposure to aqueous or alcohol environments [233,234,235]. In a study by Nakamura et al. (2007), the authors attempted to extract BITC from papaya seeds utilizing a liquid–liquid extraction protocol at room temperature [96]. The results of their methodology suggested that upon myrosinase inactivation the concentration of BITC was 213 μg/100 g of papaya seeds, substantially higher than during avoidance of the enzyme’s inactivation (68.7 mg/100 g of papaya seeds) [96]. Alternatively, solid–liquid extraction methodologies utilizing Salvadora persica (L.) as a source of BITC have been employed [236]. In such methodologies, the authors of various studies have avoided the use of hydroxylated solvents (alcohols) as BITC can form non-separable thiocarbamate conformers. Additionally, an increase in temperature also accelerates the formation of these thiocarbamate species [236,237]. For this purpose, three different classes of solvents were employed, namely chloroform (halogenated solvent), acetone (carbonyl contained solvent), and ethanol (hydroxylated solvent) [236]. The outcome of this study revealed that, among the three solvents, chloroform was the most effective as it gave the highest yield (2.1–2.4% w/w), suggesting that halogenated solvents are most efficient for BITC extraction [236]. These findings are in agreement with those of others, utilizing dichloromethane as an extraction analyte [237,238]. As expected, extraction of air dry or fresh plant material in ethanol or acetone leads to low recovery yields of 0.06–0.07% w/w or 1.12% w/w of BITC, respectively [236]. In the case of acetone, the authors suggested that the observed low recovery yield of BITC may be attributed to the water miscibility of acetone, which enhances the contact between BITC and polar components present in the plant [236,237,239]. Furthermore, hot continuous Soxhlet extraction with chloroform decreased the BITC content (0.88–0.92% w/w), possibly due to the volatile nature of BITC or its thermal degradation [202,240].
Even though there are numerous reports documenting that ITCs can be degraded under thermal conditions, a study by Nakamura et al. (2019) managed to extract BITC (of 80% purity) via heat distillation without the participation of any organic solvent [122]. Interestingly, in this study, there was no evidence of thiocarbamates formation (as a result of nucleophilic addition of water’s hydroxide in the ITC unit). In addition, the authors attempted to increase the stability of BITC, in an aqueous environment, through conjugation with glutathione (GSH), N-acetyl cysteine (NAC), or L-cysteine [122]. The results suggested that the decomposition of BITC is water was prevented by the addition of L-cysteine, but not with GSH nor NAC [122]. However, these findings contradict another study where authors examined the stability of benzylic-type ITCs in hydrodistillation conditions [241,242,243]. To this end, De Nicola et al. (2012) showed that upon exposure of BITC in an aqueous environment of greater than 90 °C, it is converted into benzylamine after 7 hr of treatment [241].

3.5. Extraction of Phenethyl Isothiocyanate

Watercress (Nasturtium officinale) is an enriched source of gluconasturtiin, the GSL precursor of PEITC [31]. It has been previously documented that PEITC is a powerful antioxidant [242,243,244,245,246,247,248,249,250,251], cancer chemo-preventive [252,253,254,255,256,257,258,259,260] and antimicrobial [261,262,263,264,265] agent. Several studies have demonstrated that the formation of PEITC is easily affected by various factors like temperature and pH. In addition, the cooking process of watercress itself inactivates myrosinase, which is susceptible to thermal denaturation, hence preventing the conversion of gluconasturtiin into PEITC [87,266].
The extraction of PEITC utilizing a simple ultrasound extraction methodology has been proposed by Fusari et al. (2019), according to which defatted rocked seeds were sonicated at 40 kHz, 600 W for 5 min in the presence of 0.1 M bicarbonate buffer (pH 8.1). The formed solution was stirred at 37 °C for 2 hrs to ensure complete conversion of GSLs into their respective ITCs. Finally, the hydrolysis solution was then extracted (from acetonitrile/dichloromethane) prior to High Performance Liquid Chromatography couple to a diode array detector (HPLC-DAD) analysis and separation [267]. A similar liquid–liquid extraction methodology using safflower oil was used for the extraction of non-polar ITCs, including PEITC, from natural sources [268]. More specifically, the authors homogenized pulverized powder of watercress, in deionized water, to allow hydrolysis of GSLs. To this solution, safflower oil (containing antioxidants like tocopherol and carotenoids) was added in order to extract PEITC from the aqueous hydrolytic mixture. Interestingly, the safflower oil and the antioxidants did not interfere with PEITC, and they protected the analyte from its oxidation and degradation [268]. Additionally, the hydrophobic nature of the oil allows the maximum extraction of PEITC without the need for further re-extraction (as it usually happened during standard liquid–liquid extraction) [268]. In another conventional solvent extraction model, utilized by Rodrigez et al. (2016), hydrolysis of gluconasturtiin was performed by incubating freeze-dried watercress humidified with 125% of distilled water, at 35 °C for 60 min, under atmospheric pressure [269]. The aqueous hydrolysis mixture, containing PEITC, was then extracted with hexane pre-warmed at 40 °C [269]. The isolated extracts contained 20.5 ± 2.7 μmol of PEITC/g of freeze-dried watercress as indicated by GC-MS coupled to UV [269]. The same extraction methodology was utilized by Ji et al. (2003) to determine the total PEITC content in human plasma and urine samples [270].
The use of SFE is widely applied for the extraction of bioactive phytochemicals from natural sources, including the recovery of PEITC from watercress. Initially, the raw freeze-dried material was incubated at 25–35 °C, under 25 MPa pressure, for 30–120 min to ensure the hydrolysis of the GSL precursors. The extraction of ITCs was achieved via supercritical CO2 in the presence or absence of ethanol at 35 °C and 25 MPa [269]. Overall, when supercritical CO2 extraction was utilized in the absence of any solvent, selectively isolated PEITC of high purity and content (31.7 ± 1.6 μmol PEITC/g of freeze-dried material) was noted. On the contrary, when a CO2: ethanol mixture was used, the concentration of the recovered PEITC decreased to 17.2 ± 2.7 μmol PEITC/g of freeze-dried material [269]. The difference in the recovered PEITC content can be attributed to the fact that ITCs are generally unstable under conditions including hydroxylated solvents (e.g., ethanol) [269]. Overall, extraction with supercritical CO2 (with or without ethanol) is a considerably more effective extraction methodology for PEITC (from a watercress source) as it has the capacity for recover higher quantities and purity levels. However, the expensive and specific instrumentation required for the implementation of such methodologies constitutes a major disadvantage.
Alternatively, Coscueta et al. (2020) exploited, for the first time, the extractive capacity of aqueous micellar systems composed of two non-ionic surfactants for the extraction of PEITC from watercress in an attempt to employ economic alternatives to flammable, toxic, and expensive organic solvents [98]. Non-ionic surfactants are amphiphiles which have been previously utilized for the extraction and purification of bioactive compounds [271,272,273,274]. Examples of non-ionic surfactants are Triton (X-100 and X-114), Brij (−30, −56 and −97), Genapol (X-080), and to a lesser extent, the Tergitols (15-S-X). A major advantage of using micelles is that their transparency in the 240–280 nm region allows the monitoring of aromatic or conjugated systems more easily. In addition, non-ionic surfactants have the capacity for developing interactions with wither hydrophobic or hydrophilic parts of the different molecules, allowing their extraction and purification more effectively [98,275,276,277,278]. Specifically, upon extraction of PEITC (from watercress) with either alcohol ethoxylate non-ionic surfactants (e.g., Genapol X-080 and Tergitol 15-S-7) or a range of organic solvents of decreasing polarity (e.g., n-hexanes and a mixture of acetonitrile/chloroform; 10:7), results suggested that its extraction with non-ionic surfactants was the most efficient one. The PEITC content obtained from the extraction with Tergitol 15-S-7 and acetonitrile/chloroform did not differ significantly from n-hexane, whereas in the case of Genapol X-080, the respective PEITC content was decreased [98]. Overall, it was suggested that the optimum conditions for PEITC should be 2.0% m/m of non-ionic surfactant at 25 °C and a pH range from neutral to slightly basic, conditions at which the concentration of the analyte was found to be 2887 and 2971 µg of PEITC/g of freeze-dried watercress with Genapol X-080 and Tergitol 15-S-7, respectively [98]. Finally, Table 4 summarizes the various methodologies used in the extraction of various ITCs from their natural sources.

4. Quantification of the Extracted ITCs Content

The isolation and quantification of ITCs from various natural sources can be a challenging process. This is mainly because each plant produces different ITCs and their extraction procedure varies depending on their physicochemical nature (polarity, stability). Furthermore, it appears that the hydrolysis of GSLs can also be a highly demanding procedure as alterations in the hydrolytic conditions (temperature and pH) can induce denaturation and inactivation of myrosinase, hence limited ITCs production. Additionally, such effect on myrosinase can promote the chemical rearrangement of GSLs into nitriles, thus posing a toxic profile. Another characteristic factor affecting the quantification of ITCs is their volatility, which makes them vulnerable to degradation under thermal conditions. As a result, several procedures have been utilised for the extraction of specific ITCs. Overall, multiple protocols have been employed, allowing the quantification of ITCs either as a mixture or the major extraction product. These methodological pipelines are summarized in Figure 5.

4.1. Chemistry and Reactivity of ITCs

The chemistry of ITCs relies mainly on the electrophilicity of their -N=C=S moiety, allowing them to react with several nucleophilic centres including oxygen, nitrogen, and sulfur, thereby forming dithiocarbamates, thioureas, and dithiocarbamates, respectively [279,280]. An example is the Edman degradation during which PEITC reacts with the amino group of amino acids, under alkaline conditions, forming an intermediate which is then rearranged into substituted phenethylthiohydantoin products [281,282]. Furthermore, the reactivity of ITCs is also facilitating their metabolism. To this end, ITCs absorbed in the liver then enter the mercapturic acid pathway in which the carbon atom of the –N=C=S moiety reacts primarily with the cysteine thiol of GSH in a reaction catalyzed by glutathione S-transferase (GST) [283]. The resulting conjugate (e.g., glutathione dithiocarbamate) undergoes further enzymatic modifications yielding sequentially a cysteinyl glycine (in a reaction facilitated by γ- glutamyl transpeptidase) and N-acetyl cysteine (NAC) (by a N-acetyl transferase) conjugates, which are excreted in urine [284,285,286,287].

4.2. Cyclo-Condensation Assay

A study by Fabian et al. (1967) was the first one that attempted to quantify ITCs [288]. This was achieved by reacting propyl ITC with 2,3-dimercaptopropanol in aqueous solution of moderated basicity [288]. The continuous ultra-violet (UV) monitoring of the reaction suggested the (rapid) formation of a product carrying the dithiocarbamate moiety with characteristic maximal absorption at 270 nm. Further incubation of this dithiocarbamate intermediate led to the development of a high intensity spectra with a single isosbestic point at 287 nm [288]. Later, in a study performed by Zhang et al. (1992), the authors isolated, purified, and characterized the isolated product, documenting the presence of a cyclic thione called 4-hydroxymethyl-1,3-dithiolane-2-thione [289]. The mechanism of the reaction suggested that the centre carbon atom of the isothiocyanate group of ITCs undergoes nucleophilic attack by a thiol group of a vicinal dithiol system, leading initially in the formation of the unstable dithiocarbamate molecule [290]. Intramolecular cyclization and consequent degradation of dithiocarbamate into a five membered 1,3-dithiolane-2-thione was shown to be promoted by a second nucleophilic attack of the unreacted thiol group (adjacent thiol group), upon prolonged standing [290]. Despite the fact that intramolecular cyclization between ITCs and vicinal amino or thiol groups had previously been demonstrated, a cyclo-condensation with vicinal dithiols had not been reported before thereby suggesting a new chemical reaction of ITCs [290,291]. This cyclo-condensation reaction was subsequently shown to be useful in generating primary amines from alkyl and aryl ITCs, but it was also realized that this reaction can be applied into the quantification of ITCs by measuring the absorbance of the formed cyclic product [291,292]. Further optimization of the cyclo-condensation reaction was performed by screening several vicinal dithiol substrates, demonstrating that reaction of ITCs with 1,2-benzenedithiol (BDT) can lead to the rapid formation of a very stable 1,3-benzodithiole-2-thione with high molar extinction coefficient at the long wavelength (ε = 23,000 M−1 cm−1 at 365 nm) (Figure 6) [289].
The cyclo-condensation reaction can be performed by all aliphatic and aromatic ITCs, except for the tertiary ones (e.g., tert-butyl ITC) [289]. Several attempts have been made to promote the reaction, such as increasing the concentration of BDT to 50 mM (rather than 4 mM) or by introducing prolong incubation periods (15 hr at 65 °C) [289]. Despite several optimization steps, only 20% of tert-butyl ITC was shown to be able to react. The low reactivity of the tertiary ITCs can be attributed to a potential steric hindrance as the tertiary carbon atom which can block the side where thiol can react to the central carbon of -N=C=S. In addition, the electronegativity of the adjacent nitrogen might be reduced due to the presence of the three methyl groups (in the case of tert-butyl ITC), preventing the elimination of amine [289]. The quantification of ITCs from blood, urine, plant extracts, etc., via the cyclo-condensation assay, does not require their prior modification. In the past, labelling of the ITCs with radioactive 14C or deuterium was essential for their quantification [267,285,293,294,295,296]. With respect to the sensitivity of the assay, it has been previously reported that the limit of cyclic adduct detection can be as low as 10 pmol, according to simple reverse phase HPLC area integration, using a C18 column and isocratic mobile phase consisting of 80% methanol, 20% water, and a photodiode array detection unit [297]. It has also been suggested that the sensitivity could be lowered at least another 10-fold by utilizing a solid phase extraction of the ITC coupled 1,3-benzodithiole-2-thione via a reverse phase cartridge (Sep-Pak) [298]. Since the cyclisation reaction occurs under relatively basic conditions, the substitution of phosphate buffer with a borate one (500 mM at pH 9.25) can increase the stability and the sensitivity of the detection of ITCs in urine samples. Alternatively, the use of aprotic polar solvents (e.g., acetone, acetonitrile, dimethylsulfoxide, and dimethylformamide) can facilitate the solubility of the reaction solution in high protein-containing samples [298].
Finally, the main disadvantage of the cyclo-condensation assay is that it does not allow the discrimination of each individual ITC or dithiocarbamate molecule in a mixture, since both molecules form the same cyclic product. As a result, the urinary levels of ITC equivalents (including ITCs and dithiocarbamates) were inversely associated with carcinogen-induced DNA damage as well as for the risk of developing stomach, lung, colon, and breast carcinomas [267,296,297]. Furthermore, the chance of having a positive false during the detection is likely, as a number of chemical entities can react with BDT in the same way as ITCs do [287,294].

4.3. Quantification via ITCs Chemical Derivatization

In many cases, the quantification of ITCs appears to be a challenging procedure due to their limited stability in either aqueous or methanolic media solutions [170,299,300,301]. As a result, in many studies, the quantification of ITCs was performed either by measuring the concentration of the degradation products or by derivatizing the ITC extracts prior to quantification. The chemical derivatization of ITCs results in the formation of considerably more stable compounds, allowing their quantification by analytical instrumentation.
In the case of PEITC, a study by Negrusz et al. (1998) used the degradation product phenethylamine as a marker to quantify the content of PEITC found in dog’s plasma by using gas chromatography coupled to mass spectrometry with chemical ionization [302]. In another study, the determination and quantification of PEITC in human plasma and urine was performed by derivatizing these samples with ammonia [270,303]. Chemical derivatization of samples is an approach that results from the major disadvantage of the cyclo-condensation assay which does not allow the distinction between PEITC from other species like phenethyl-N-acetyl cysteine, dithiocarbamates, etc. [270,303]. In addition, the derivatization allows the recovery of higher amounts of ITC as well since there is no loss, due to volatility, during the evaporation process [270,303]. Even though mercapturate metabolites (e.g., phenethyl-ITC-N-acetyl cysteine) can also be converted to phenethyl thiourea, after treatment with ammonia, these can be removed during the hexane-based extraction phase [270]. Therefore, any false positives when determining PEITC content are potentially prevented. For instance, unlike the cyclo-condensation reaction, the resulting thiourea product is unique for each individual ITC and so interference from any other ITCs (which might present in the sample) is prevented [270,303]. The same approach was followed by another study when determining the content of PEITC in human blood samples [303,304].
Finally, others have postulated that the accurate quantification of ITCs (via reverse-phase Liquid Chromatography on C18 cartridges) of low polarity (e.g., hepty-ITC or SFN) is prevented due to their aqueous precipitation. To overcome this effect, the authors of another study derivatized the extracted ITCs with mercaptoethanol, thus allowing the formation of more polar and consequently more water-soluble compounds [305]. This derivatization facilitated the analysis of their corresponding ITCs using an aqueous or water-containing mobile phase, in addition to increasing the overall sensitivity of the assay’s detection and quantification limits [305].

4.4. Quantification via Analytical Instrumentation

4.4.1. Attenuated Total Reflectance Infrared Fourier Transform (ATR-FT-IR) Spectroscopy

As it is mentioned above, the instability of ITCs, as well as their precipitation in aqueous mobile phases, led many researchers to derivatize ITCs in an attempt to minimize these factors. However, the derivatization of ITCs introduces extra steps into the experimental methodology of quantification, with major impacts including cost effective reagents as well as the possibility of minimizing the yield of product formation. Therefore, Revelou et al. (2017), for the first time, utilized attenuated total reflectance infrared Fourier transform (ATR-FT-IR) spectroscopy and partial least-squares for the determination of total isothiocyanate content in broccoli [306]. For this purpose, each spectrum was recorded using a zinc–selenide (ZnSe) 45 flat plate against a ZnSe background and the spectra smoothing was performed by applying a Savitsky–Golay algorithm [306]. The increment in the concentration of the analyte standard was accompanied with an increment in the absorption at spectral region; 2150–2020 cm−1, whereas the doublet peak at 2120 and 2058 cm−1 of the same spectra has been correlated with asymmetric stretching of N=C=S functionality, since blank samples did not detect such signals [307]. Analysis of the broccoli extracts suggests the appearance of the characteristic isothiocyanate bond stretching (by means of doublet peak) verified the above findings. Those findings proposed the utilization of ATR-FTIR methodology as an alternative quick, reproducible, and accurate approach for the determination and quantification of total isothiocyanate content in Brassica vegetables [306]. Despite the encouraging results that were obtained, the authors did not characterize this technique in terms of selectivity. Since the quantification of ITCs via ATR-FTIR relies on the signal’s intensity, it should be expected, like cyclo-condensation, that the same absorbance would be for all ITCs. Therefore, this method can only be used and applied in cases where only one ITC is omitted.

4.4.2. Chromatographic Approaches

The use of analytical instrumentation (e.g., liquid (LC) and/or gas chromatography (GC)) coupled to either a UV or PDA (photodiode detector array) detection unit or tandem mass spectrometry (MS/MS) has been widely used in the isolation and quantification of phytochemicals, including ITCs. Initially, the quantification of ITCs was performed on gas chromatography coupled to mass spectrometry. Marton et al. (2013) quantified the total ITC content (from mustard seeds) by employing a GC-MS/MS-based approach, according to which each analyte was eluted through a preheated column and then ionized via electron ionization [38]. The authors concluded that this methodology was more time-cost efficient, simple, reproducible, and sensitive [38]. In another study, the quantification of SFN from broccoli was attempted by using GC-MS instrumentation [308]. In this case, the authors stated that this methodology caused approximately 80% degradation of SFN into 3-butenyl ITC [308]. To minimize the thermal degradation (down to 5%), the authors utilized a combination of a 1.5 mm direct inlet liner in conjugation with fast initial injection flow. These conditions improved the on-column injection, thermal degradation, and also prolonged the column deterioration. These findings are in agreement with a study by Jin et al. (1999) which confirmed thermal degradation of SFN at 50 °C and 100 °C (temperatures that GC instrumentation operates) [309]. To this end, an alternative way of quantifying the content of ITCs (in Chinese herbs) was proposed, during which a cyclo-condensation assay was utilized, followed by the monitoring of the quantity of the resulting 1,3-benzodithiole-2-thione molecule by GC-MS and selective ion recording, to increase the sensitivity of detection [310]. The authors concluded that the GC-MS method for the analysis of ITC content in vegetables and herbs is an optimum one since 1,3-benzodithiole-2-thione is highly stable, thus allowing for robust data of high sensitivity [310]. Eventually, numerous reports suggested the use of GC-XX as an optimal instrumentation for the determination and quantification of ITC content in vegetables [302,311,312,313,314]. However, the thermal decomposition of ITCs during the analysis is an important factor that should be taken into consideration during quantification.
In many studies, the quantification of extracted ITCs has been performed by utilizing HPLC instrumentation coupled to several detection sources (e.g., MS, PDA, or UV). Zheng et al. (2014) attempted to quantify the PEITC content in human plasma, taking into consideration all previously published methodologies together with all the associated limitations (e.g., duration, scale, sensitivity, and selectivity) [315]. For the first time, the authors utilised atmospheric pressure chemical ionization (APCI) in order to detect and quantify PEITC. The authors recommended the use of this ionization mode to be considerably more effective for the ionisation of PEITC when compared to the electrospray ionisation using the LC-MS/MS platform [315]. This is because the protonation of phenethyl ITC under electron spray ionisation (ESI) occurs in the liquid phase, whereas the respective protonation under APCI occurs in the gaseous phase. Furthermore, when considering the instability of PEITC, simple protein precipitation (with acetonitrile) was utilized prior to exposure of the sample at low temperature under acidic conditions. The quantitation was then performed by using multiple reaction monitoring (MRM) mode for the transition: PEITC, [M + H]+ m/z 164.0→m/z 130.0 [312]. Eventually, the authors concluded that this methodology comprised a simple, sensitive, and valid approach for the quantification of PEITC in human plasma samples. A later study by Yu et al. (2020) suggested a less expensive but yet effective methodology for the determination of SFN content by utilizing a UHPLC-MS/MS-based approach [316]. The overall analysis relied on the MRM mode, according to which the SFN content was quantified using the following transition [M+H]+ m/z 178.3 →m/z 114.2 [316]. Similar approaches were employed by others in an attempt to quantify SFN content from various natural sources [30,162,317,318].

5. Conclusions

This review summarises the experimental pipeline followed in order to extract and quantify major ITCs from naturally-derived sources. The entire process (which varies across the different classes of ITCs) involves the conversion of GSLs into ITCs prior to their extraction. The major limitation(s) of this pipeline is that there is not an established universal extraction and quantification protocol which can be applied to all ITCs. Therefore, their quantitative extraction (from a single natural source) can be a challenging process. In addition, the chemical instability of the ITCs moiety makes them prone to destabilisation when exposed to various solvents and/or high temperature. Finally, several challenges are observed during the entire process of isolation and quantification of ITCs, even if the same protocol is followed. To an extent, this can be potentially attributed to the different cultivation and growing conditions of the primary natural sources.
Given the crucial role of ITCs in modulating several cellular cascades, it is of utmost importance to be able to determine their content in various naturally-derived sources. In future studies, the development of a single universal pipeline should be a priority in order to allow the quantitative extraction of ITCs from naturally occurring sources. This, in turn, would potentially allow further exploitation of ITCs in drug development, among other applications.

Author Contributions

Conceptualization, S.K. and M.I.P.; resources, M.I.P.; writing—original draft preparation, S.K.; writing—review and editing, S.K., D.T.T., M.V.D., R.F., A.P. and M.I.P.; supervision, M.I.P.; funding acquisition, M.I.P. All authors have read and agreed to the published version of the manuscript.

Funding

This work was supported by a grant provided by the Cyprus Institute of Neurology and Genetics (Telethon Cyprus), Nicosia, Cyprus (MIP).

Conflicts of Interest

The authors declare no conflict of interest.

Abbreviations

AITCallyl isothiocyanate
APCIatmospheric pressure chemical ionization
ATR-FT-IRattenuated total reflectance infrared Fourier transform spectroscopy
BDT1,2-benzenedithiol
BITCbenzyl isothiocyanate
CDKcyclic depended kinase
DADdiode-array detection
EDTAethylenediaminetetraacetic acid
ESIelectron spray ionisation
ESPepithiospecifier protein-like factors
FIDflame ionization detection
GCgas chromatography
GLsglucosinolates
GSHglutathione
GSTglutathione S-transferase
HPLChigh performance liquid chromatography
IBNiberin
ITCisothiocyanates
MRMmultiple reaction monitoring
MSmass spectrometry
MS/MStandem mass spectrometry
m/zmass to charge ratio
NACN-acetyl cysteine
PDAphotodiode-array detection
PEITCphenethyl isothiocyanate
ROSreactive oxygen species
SFEsupercritical fluid extraction
SFNsulforaphane
SPEsolid phase extraction
UVultra-violate

References

  1. Rodriguez-Casado, A. The Health Potential of Fruits and Vegetables Phytochemicals: Notable Examples. Crit. Rev. Food Sci. Nutr. 2016, 56, 1097–1107. [Google Scholar] [CrossRef] [PubMed]
  2. Alasalvar, C.; Bolling, B.W. Review of Nut Phytochemicals, Fat-Soluble Bioactives, Antioxidant Components and Health Effects. Br. J. Nutr. 2015, 113, S68–S78. [Google Scholar] [CrossRef] [PubMed]
  3. Chang, S.K.; Alasalvar, C.; Shahidi, F. Superfruits: Phytochemicals, Antioxidant Efficacies, and Health Effects—A Comprehensive Review. Crit. Rev. Food Sci. Nutr. 2019, 59, 1580–1604. [Google Scholar] [CrossRef] [PubMed]
  4. Zhang, Y.-J.; Gan, R.-Y.; Li, S.; Zhou, Y.; Li, A.-N.; Xu, D.-P.; Li, H.-B. Antioxidant Phytochemicals for the Prevention and Treatment of Chronic Diseases. Molecules 2015, 20, 21138–21156. [Google Scholar] [CrossRef]
  5. Nabavi, S.M.; Daglia, M.; Braidy, N.; Nabavi, S.F. Natural Products, Micronutrients, and Nutraceuticals for the Treatment of Depression: A Short Review. Nutr. Neurosci. 2017, 20, 180–194. [Google Scholar] [CrossRef]
  6. Xu, D.-P.; Li, Y.; Meng, X.; Zhou, T.; Zhou, Y.; Zheng, J.; Zhang, J.-J.; Li, H.-B. Natural Antioxidants in Foods and Medicinal Plants: Extraction, Assessment and Resources. Int. J. Mol. Sci. 2017, 18, 96. [Google Scholar] [CrossRef]
  7. Arulselvan, P.; Fard, M.T.; Tan, W.S.; Gothai, S.; Fakurazi, S.; Norhaizan, M.E.; Kumar, S.S. Role of Antioxidants and Natural Products in Inflammation. Oxid. Med. Cell. Longev. 2016, 2016, 5276130. [Google Scholar] [CrossRef] [Green Version]
  8. Baby, B.; Antony, P.; Vijayan, R. Antioxidant and Anticancer Properties of Berries. Crit. Rev. Food Sci. Nutr. 2018, 58, 2491–2507. [Google Scholar] [CrossRef]
  9. Azzini, E.; Giacometti, J.; Russo, G.L. Antioxidant Phytochemicals at the Pharma-Nutrition Interface. Oxidative Med. Cell. Longev. 2017, 2017, 6986143. [Google Scholar] [CrossRef] [Green Version]
  10. Lu, B.; Zhao, Y. Photooxidation of Phytochemicals in Food and Control: A Review. Ann. N. Y. Acad. Sci. 2017, 1398, 72–82. [Google Scholar] [CrossRef]
  11. Satija, A.; Hu, F.B. Plant-Based Diets and Cardiovascular Health. Trends Cardiovasc. Med. 2018, 28, 437–441. [Google Scholar] [CrossRef] [PubMed]
  12. Miller, V.; Mente, A.; Dehghan, M.; Rangarajan, S.; Zhang, X.; Swaminathan, S.; Dagenais, G.; Gupta, R.; Mohan, V.; Lear, S.; et al. Fruit, Vegetable, and Legume Intake, and Cardiovascular Disease and Deaths in 18 Countries (PURE): A Prospective Cohort Study. Lancet 2017, 390, 2037–2049. [Google Scholar] [CrossRef] [Green Version]
  13. Vieira, A.R.; Abar, L.; Vingeliene, S.; Chan, D.S.M.; Aune, D.; Navarro-Rosenblatt, D.; Stevens, C.; Greenwood, D.; Norat, T. Fruits, Vegetables and Lung Cancer Risk: A Systematic Review and Meta-Analysis. Ann. Oncol. Off. J. Eur. Soc. Med. Oncol. 2016, 27, 81–96. [Google Scholar] [CrossRef] [PubMed]
  14. Aune, D.; Giovannucci, E.; Boffetta, P.; Fadnes, L.T.; Keum, N.; Norat, T.; Greenwood, D.C.; Riboli, E.; Vatten, L.J.; Tonstad, S. Fruit and Vegetable Intake and the Risk of Cardiovascular Disease, Total Cancer and All-Cause Mortality-a Systematic Review and Dose-Response Meta-Analysis of Prospective Studies. Int. J. Epidemiol. 2017, 46, 1029–1056. [Google Scholar] [CrossRef] [PubMed]
  15. Guo, T.; Liu, C.; Gao, Z.; He, Y. Study on Effects and Mechanisms of Phytochemicals in Vegetables and Fruits in Preventing and Treating Lung Cancer. Chin. J. Lung Cancer 2017, 20, 841–846. [Google Scholar] [CrossRef]
  16. Sayem, A.S.M.; Arya, A.; Karimian, H.; Krishnasamy, N.; Ashok Hasamnis, A.; Hossain, C.F. Action of Phytochemicals on Insulin Signaling Pathways Accelerating Glucose Transporter (GLUT4) Protein Translocation. Molecules 2018, 23, 258. [Google Scholar] [CrossRef] [Green Version]
  17. Cicero, A.F.G.; Colletti, A. Role of Phytochemicals in the Management of Metabolic Syndrome. Phytomedicine 2016, 23, 1134–1144. [Google Scholar] [CrossRef]
  18. Ansari, P.; Flatt, P.R.; Harriott, P.; Abdel-Wahab, Y.H.A. Evaluation of the antidiabetic and insulin releasing effects of A. Squamosa including isolation and characterization of active phytochemicals. Plants 2020, 9, 1348. [Google Scholar] [CrossRef]
  19. Bacanli, M.; Dilsiz, S.A.; Başaran, N.; Başaran, A.A. Effects of Phytochemicals against Diabetes. Adv. Food Nutr. Res. 2019, 89, 209–238. [Google Scholar] [CrossRef]
  20. Onaolapo, A.Y.; Onaolapo, O.J. Nutraceuticals and diet-based phytochemicals in Type 2 diabetes mellitus: From whole food to components with defined roles and mechanisms. Curr. Diabetes Rev. 2019, 16, 12–25. [Google Scholar] [CrossRef]
  21. Zhao, C.; Yang, C.; Wai, S.T.C.; Zhang, Y.; Portillo, M.P.; Paoli, P.; Wu, Y.; San Cheang, W.; Liu, B.; Carpéné, C.; et al. Regulation of Glucose Metabolism by Bioactive Phytochemicals for the Management of Type 2 Diabetes Mellitus. Crit. Rev. Food Sci. Nutr. 2019, 59, 830–847. [Google Scholar] [CrossRef]
  22. La Rosa, F.; Clerici, M.; Ratto, D.; Occhinegro, A.; Licito, A.; Romeo, M.; Di Iorio, C.; Rossi, P. The Gut-Brain Axis in Alzheimer’s Disease and Omega-3. A Critical Overview of Clinical Trials. Nutrients 2018, 10, 1267. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  23. Spagnuolo, C.; Moccia, S.; Russo, G.L. Anti-Inflammatory Effects of Flavonoids in Neurodegenerative Disorders. Eur. J. Med. Chem. 2018, 153, 105–115. [Google Scholar] [CrossRef]
  24. Uddin, M.S.; Kabir, M.T.; Niaz, K.; Jeandet, P.; Clément, C.; Mathew, B.; Rauf, A.; Rengasamy, K.R.R.; Sobarzo-Sánchez, E.; Ashraf, G.M.; et al. Molecular Insight into the Therapeutic Promise of Flavonoids against Alzheimer’s Disease. Molecules 2020, 25, 1267. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  25. Zhao, X.; Zhang, M.; Li, C.; Jiang, X.; Su, Y.; Zhang, Y. Benefits of Vitamins in the Treatment of Parkinson’s Disease. Oxid. Med. Cell. Longev. 2019, 2019, 9426867. [Google Scholar] [CrossRef] [PubMed]
  26. Palliyaguru, D.L.; Yuan, J.-M.; Kensler, T.W.; Fahey, J.W. Isothiocyanates: Translating the Power of Plants to People. Mol. Nutr. Food Res. 2018, 62, e1700965. [Google Scholar] [CrossRef] [PubMed]
  27. Assony, S.J. Chapter 28-The Chemistry of Isothiocyanates. In Organic Sulfur Compounds; Kharasch, N., Ed.; Pergamon: Oxfrord, UK, 1961; pp. 326–338. [Google Scholar] [CrossRef]
  28. Ghawi, S.K.; Methven, L.; Niranjan, K. The Potential to Intensify Sulforaphane Formation in Cooked Broccoli (Brassica oleracea var. italica) Using Mustard Seeds (Sinapis alba). Food Chem. 2013, 138, 1734–1741. [Google Scholar] [CrossRef] [PubMed]
  29. Li, Z.; Liu, Y.; Fang, Z.; Yang, L.; Zhuang, M.; Zhang, Y.; Zhao, W.; Sun, P. Variation of Sulforaphane Levels in Broccoli (Brassica Oleracea var. Italica) during Flower Development and the Role of Gene AOP2. J. Liq. Chromatogr. Relat. Technol. 2014, 37, 1199–1211. [Google Scholar] [CrossRef]
  30. Vanduchova, A.; Anzenbacher, P.; Anzenbacherova, E. Isothiocyanate from Broccoli, Sulforaphane, and Its Properties. J. Med. Food 2019, 22, 121–126. [Google Scholar] [CrossRef]
  31. Gupta, P.; Wright, S.E.; Kim, S.-H.; Srivastava, S.K. Phenethyl Isothiocyanate: A Comprehensive Review of Anti-Cancer Mechanisms. Biochim. Biophys. Acta 2014, 1846, 405–424. [Google Scholar] [CrossRef] [Green Version]
  32. Liang, H.; Yuan, Q.P.; Dong, H.R.; Liu, Y.M. Determination of Sulforaphane in Broccoli and Cabbage by High-Performance Liquid Chromatography. J. Food Compos. Anal. 2006, 19, 473–476. [Google Scholar] [CrossRef]
  33. Koo, S.Y.; Cha, K.H.; Song, D.-G.; Lee, D.-U.; Pan, C.-H. Increased Sulforaphane Concentration in Brussels Sprout Following High Hydrostatic Pressure Treatment. J. Korean Soc. Appl. Biol. Chem. 2012, 55, 685–687. [Google Scholar] [CrossRef]
  34. Jakobsen, T.H.; Bragason, S.K.; Phipps, R.K.; Christensen, L.D.; van Gennip, M.; Alhede, M.; Skindersoe, M.; Larsen, T.O.; Høiby, N.; Bjarnsholt, T.; et al. Food as a Source for Quorum Sensing Inhibitors: Iberin from Horseradish Revealed as a Quorum Sensing Inhibitor of Pseudomonas Aeruginosa. Appl. Environ. Microbiol. 2012, 78, 2410–2421. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  35. Wu, H.; Zhang, G.-A.; Zeng, S.; Lin, K. Extraction of Allyl Isothiocyanate from Horseradish (Armoracia Rusticana) and Its Fumigant Insecticidal Activity on Four Stored-Product Pests of Paddy. Pest Manag. Sci. 2009, 65, 1003–1008. [Google Scholar] [CrossRef]
  36. Dixon, M.J.; Shaw, P.J. Watercress and Water Quality: The Effect of Phenethyl Isothiocyanate on the Mating Behaviour of Gammarus pulex. Int. J. Zool. 2011, 2011, 328749. [Google Scholar] [CrossRef] [Green Version]
  37. Li, L.; Lee, W.; Lee, W.J.; Auh, J.H.; Kim, S.S.; Yoon, J. Extraction of Allyl Isothiocyanate from Wasabi (Wasabia Japonica Matsum) Using Supercritical Carbon Dioxide. Food Sci. Biotechnol. 2010, 19, 405–410. [Google Scholar] [CrossRef]
  38. Márton, M.; Lavric, V. A Simple Method for the Quantification of Isothiocyanates from Mustard. U.P.B. Sci. Bull. Ser. B 2013, 75, 63–72. [Google Scholar]
  39. Ma, Y.; Wen, Y.; Chen, J.; Zhang, Y.; Zhang, H.; Sui, J.; Yi, G.; He, X. Rapid and sensitive analysis of benzyl isothiocyanate in peel, pulp, and seeds of Carica papaya linn. by headspace gas chromatography-mass spectrometry. SN Appl. Sci. 2021, 374, 1–9. [Google Scholar] [CrossRef]
  40. Jeschke, V.; Gershenzon, J.; Vassão, D.G. A Mode of Action of Glucosinolate-Derived Isothiocyanates: Detoxification Depletes Glutathione and Cysteine Levels with Ramifications on Protein Metabolism in Spodoptera littoralis. Insect Biochem. Mol. Biol. 2016, 71, 37–48. [Google Scholar] [CrossRef]
  41. Wittstock, U.; Agerbirk, N.; Stauber, E.J.; Olsen, C.E.; Hippler, M.; Mitchell-Olds, T.; Gershenzon, J.; Vogel, H. Successful Herbivore Attack Due to Metabolic Diversion of a Plant Chemical Defense. Proc. Natl. Acad. Sci. USA 2004, 101, 4859–4864. [Google Scholar] [CrossRef] [Green Version]
  42. Bhat, R.; Vyas, D. Myrosinase: Insights on Structural, Catalytic, Regulatory, and Environmental Interactions. Crit. Rev. Biotechnol. 2019, 39, 508–523. [Google Scholar] [CrossRef] [PubMed]
  43. Chhajed, S.; Misra, B.B.; Tello, N.; Chen, S. Chemodiversity of the Glucosinolate-Myrosinase System at the Single Cell Type Resolution. Front. Plant Sci. 2019, 10, 618. [Google Scholar] [CrossRef] [PubMed]
  44. Peñas, E.; Frias, J.; Martínez-Villaluenga, C.; Vidal-Valverde, C. Bioactive compounds, myrosinase activity and antioxidant capacity of white cabbages grown in different locations of Spain. J. Agric. Food Chem. 2011, 59, 3772–3779. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  45. Bellostas, N.; Sørensen, J.; Sørensen, H. Qualitative and Quantitative Evaluation of Glucosinolates in Cruciferous Plants during Their Life Cycles. Agroindustria 2004, 3, 5–10. [Google Scholar]
  46. Kim, M.J.; Chiu, Y.-C.; Ku, K.-M. Glucosinolates, Carotenoids, and Vitamins E and K Variation from Selected Kale and Collard Cultivars. J. Food Qual. 2017, 2017, 5123572. [Google Scholar] [CrossRef] [Green Version]
  47. Alvarez, S.; He, Y.; Chen, S. Comparative Investigations of the Glucosinolate-Myrosinase System in Arabidopsis Suspension Cells and Hypocotyls. Plant Cell Physiol. 2008, 49, 324–333. [Google Scholar] [CrossRef] [Green Version]
  48. Wang, J.; Yu, H.; Zhao, Z.; Sheng, X.; Shen, Y.; Gu., H. Natural Variation of Glucosinolates and Their Breakdown Products in Broccoli (Brassica oleracea var. italica) Seeds. J. Agric. Food Chem. 2019, 67, 12528–12537. [Google Scholar] [CrossRef]
  49. Sánchez-Pujante, P.J.; Borja-Martínez, M.; Pedreño, M.Á.; Almagro, L. Biosynthesis and Bioactivity of Glucosinolates and Their Production in Plant in Vitro Cultures. Planta 2017, 246, 19–32. [Google Scholar] [CrossRef]
  50. Guerrero-Alonso, A.; Antunez-Mojica, M.; Medina-Franco, J.L. Chemoinformatic Analysis of Isothiocyanates: Their Impact in Nature and Medicine. Mol. Inf. 2021, 40, 2411–2502. [Google Scholar] [CrossRef]
  51. Iriti, M.; Faoro, F. Chemical Diversity and Defence Metabolism: How Plants Cope with Pathogens and Ozone Pollution. Int. J. Mol. Sci. 2009, 10, 3371–3399. [Google Scholar] [CrossRef] [Green Version]
  52. Sarikamis, G.; Marquez, J.; Maccormack, R.; Bennett, R.; Roberts, J.; Mithen, R. High Glucosinolate Broccoli: A Delivery System for Sulforaphane. Mol. Breed. 2006, 18, 219–228. [Google Scholar] [CrossRef] [Green Version]
  53. Dinkova-Kostova, A.T.; Kostov, R.V. Glucosinolates and Isothiocyanates in Health and Disease. Trends Mol. Med. 2012, 18, 337–347. [Google Scholar] [CrossRef] [PubMed]
  54. Radojcic Redovnikovic, I.; Glivetic, T.; Delonga, K.; Vorkapic-Furac, J. Glucosinolates and Their Potential Role in Plant. Period. Biol. 2008, 110, 297–309. [Google Scholar]
  55. Lee, J.W.; Kim, I.H.; Woyengo, T.A. Toxicity of Canola-Derived Glucosinolate Degradation Products in Pigs—A Review. Animals 2020, 10, 2337. [Google Scholar] [CrossRef]
  56. Hanschen, F.S.; Klopsch, R.; Oliviero, T.; Schreiner, M.; Verkerk, R.; Dekker, M. Optimizing Isothiocyanate Formation during Enzymatic Glucosinolate Breakdown by Adjusting pH Value, Temperature and Dilution in Brassica Vegetables and Arabidopsis thaliana. Sci. Rep. 2017, 7, 40807. [Google Scholar] [CrossRef]
  57. Williams, D.J.; Critchley, C.; Pun, S.; Chaliha, M.; O’Hare, T.J. Differing Mechanisms of Simple Nitrile Formation on Glucosinolate Degradation in Lepidium sativum and Nasturtium officinale Seeds. Phytochemistry 2009, 70, 1401–1409. [Google Scholar] [CrossRef]
  58. Hanschen, F.S.; Schreiner, M. Isothiocyanates, Nitriles, and Epithionitriles from Glucosinolates Are Affected by Genotype and Developmental Stage in Brassica Oleracea Varieties. Front. Plant Sci. 2017, 8, 1095. [Google Scholar] [CrossRef] [Green Version]
  59. Prieto, M.A.; López, C.J.; Simal-Gandara, J. Glucosinolates: Molecular Structure, Breakdown, Genetic, Bioavailability, Properties and Healthy and Adverse Effects. Adv. Food Nutr. Res. 2019, 90, 305–350. [Google Scholar] [CrossRef]
  60. Cole, R.A. Isothiocyanates, Nitriles and Thiocyanates as Products of Autolysis of Glucosinolates in Cruciferae. Phytochemistry 1976, 15, 759–762. [Google Scholar] [CrossRef]
  61. Lambrix, V.; Reichelt, M.; Mitchell-Olds, T.; Kliebenstein, D.J.; Gershenzon, J. The Arabidopsis Epithiospecifier Protein Promotes the Hydrolysis of Glucosinolates to Nitriles and Influences Trichoplusia ni Herbivory. Plant Cell 2001, 13, 2793–2807. [Google Scholar] [CrossRef] [Green Version]
  62. Kissen, R.; Bones, A.M. Nitrile-Specifier Proteins Involved in Glucosinolate Hydrolysis in Arabidopsis thaliana. J. Biol. Chem. 2009, 284, 12057–12070. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  63. Kissen, R.; Hyldbakk, E.; Wang, C.-W.V.; Sørmo, C.G.; Rossiter, J.T.; Bones, A.M. Ecotype Dependent Expression and Alternative Splicing of Epithiospecifier Protein (ESP) in Arabidopsis thaliana. Plant Mol. Biol. 2012, 78, 361–375. [Google Scholar] [CrossRef] [PubMed]
  64. Yang, G.; Gao, T.Y.; Shu, X.O.; Cai, Q.; Li, G.L.; Ji, B.T.; Rothman, D.; Dyba, M.; Xiang, Y.B.; Chung, F.L.; et al. Isothiocyanate exposure, glutathione S-transferase polymorphisms, and colorectal cancer risk. Am. J. Clin. Nutr. 2009, 91, 704–711. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  65. Wu, X.; Zhu, Y.; Yan, H.; Liu, B.; Li, Y.; Zhou, Q.; Xu, K. Isothiocyanates Induce Oxidative Stress and Suppress the Metastasis Potential of Human Non-Small Cell Lung Cancer Cells. BMC Cancer 2010, 10, 269. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  66. Zhu, Y.; Liu, S.; Yan, S.; Wang, J.; Zhang, L.; Li, X.; Wen, L.; Wu, J. Phenylethyl Isothiocyanate Induces Oxidative Damage of Porcine Kidney Cells Mediated by Reactive Oxygen Species. J. Biochem. Mol. Toxicol. 2020, 34, e22428. [Google Scholar] [CrossRef]
  67. Seow, A.; Yuan, J.-M.; Sun, C.-L.; Van Den Berg, D.; Lee, H.-P.; Yu, M.C. Dietary Isothiocyanates, Glutathione S -Transferase Polymorphisms and Colorectal Cancer Risk in the Singapore Chinese Health Study. Carcinogenesis 2002, 23, 2055–2061. [Google Scholar] [CrossRef] [Green Version]
  68. Nakamura, Y. Chemoprevention by Isothiocyanates: Molecular Basis of Apoptosis Induction. Forum Nutr. 2009, 61, 170–181. [Google Scholar] [CrossRef]
  69. Huang, L.; Cai, C.; Dang, W.; Lu, J.; Hu, G.; Gu, J. Propyl Isothiocyanate Induces Apoptosis in Gastric Cancer Cells by Oxidative Stress via Glutathione Depletion. Oncol. Lett. 2019, 18, 5490–5498. [Google Scholar] [CrossRef]
  70. Han, K.W.W.; Po, W.W.; Sohn, U.D.; Kim, H.-J. Benzyl Isothiocyanate Induces Apoptosis via Reactive Oxygen Species-Initiated Mitochondrial Dysfunction and DR4 and DR5 Death Receptor Activation in Gastric Adenocarcinoma Cells. Biomolecules 2019, 9, 839. [Google Scholar] [CrossRef] [Green Version]
  71. Hu, J.; Straub, J.; Xiao, D.; Singh, S.V.; Yang, H.-S.; Sonenberg, N.; Vatsyayan, J. Phenethyl Isothiocyanate, a Cancer Chemopreventive Constituent of Cruciferous Vegetables, Inhibits Cap-Dependent Translation by Regulating the Level and Phosphorylation of 4E-BP1. Cancer Res. 2007, 67, 3569–3573. [Google Scholar] [CrossRef] [Green Version]
  72. Tang, L.; Zhang, Y. Dietary Isothiocyanates Inhibit the Growth of Human Bladder Carcinoma Cells. J. Nutr. 2004, 134, 2004–2010. [Google Scholar] [CrossRef] [PubMed]
  73. Zhang, Y.; Tang, L.; Gonzalez, V. Selected Isothiocyanates Rapidly Induce Growth Inhibition of Cancer Cells. Mol. Cancer Ther. 2003, 2, 1045–1052. [Google Scholar] [PubMed]
  74. Mantso, T.; Anestopoulos, I.; Lamprianidou, E.; Kotsianidis, I.; Pappa, A.; Panayiotidis, M.I. Isothiocyanate-Induced Cell Cycle Arrest in a Novel In Vitro Exposure Protocol of Human Malignant Melanoma (A375) Cells. Anticancer Res. 2019, 39, 591–596. [Google Scholar] [CrossRef] [PubMed]
  75. Cavell, B.E.; Syed Alwi, S.S.; Donlevy, A.; Packham, G. Anti-Angiogenic Effects of Dietary Isothiocyanates: Mechanisms of Action and Implications for Human Health. Biochem. Pharmacol. 2011, 81, 327–336. [Google Scholar] [CrossRef] [Green Version]
  76. Keum, Y.-S.; Jeong, W.-S.; Kong, A.-N.T. Chemopreventive Functions of Isothiocyanates. Drug News Perspect. 2005, 18, 445–451. [Google Scholar] [CrossRef]
  77. Sundaram, M.K.; Preetha, R.; Haque, S.; Akhter, N.; Khan, S.; Ahmed, S.; Hussain, A. Dietary Isothiocyanates Inhibit Cancer Progression by Modulation of Epigenome. Semin. Cancer Biol. 2021, in press, S1044-579X(20)30281-9. [Google Scholar] [CrossRef]
  78. Mitsiogianni, M.; Amery, T.; Franco, R.; Zoumpourlis, V.; Pappa, A.; Panayiotidis, M.I. From chemoprevention to epigenetic regulation: The role of isothiocyanates in skin cancer prevention. Pharmacol. Ther. 2018, 190, 187–201. [Google Scholar] [CrossRef]
  79. Novío, S.; Núñez-Iglesias, M.J.; Freire-Garabal, M. Chapter 7-Isothiocyanates, Epigenetics, and Cancer Prevention. In Translational Epigenetics; Bishayee, A., Bhatia, D., Eds.; Academic Press: Cambridge, MA, USA, 2019; Volume 8, pp. 149–168. [Google Scholar] [CrossRef]
  80. Wang, Q.; Bao, Y. Nanodelivery of natural isothiocyanates as a cancer therapeutic. Free Radic. Biol. Med. 2021, 167, 125–140. [Google Scholar] [CrossRef]
  81. Oliviero, T.; Verkerk, R.; Dekker, M. Isothiocyanates from Brassica Vegetables—Effects of Processing, Cooking, Mastication, and Digestion. Mol. Nutr. Food Res. 2018, 62, 1701069. [Google Scholar] [CrossRef] [Green Version]
  82. Song, L.; Thornalley, P.J. Effect of Storage, Processing and Cooking on Glucosinolate Content of Brassica Vegetables. Food Chem. Toxicol. 2007, 45, 216–224. [Google Scholar] [CrossRef]
  83. Rungapamestry, V.; Duncan, A.J.; Fuller, Z.; Ratcliffe, B. Effect of Cooking Brassica Vegetables on the Subsequent Hydrolysis and Metabolic Fate of Glucosinolates. Proc. Nutr. Soc. 2007, 66, 69–81. [Google Scholar] [CrossRef] [PubMed]
  84. Kosson, R.; Horbowicz, M. Effect of Long-Term Storage on Some Nutritive Components and Isothiocyanates Content in Roots of Two Horseradish Types. Veg. Crop. Res. Bull. 2008, 69, 155–164. [Google Scholar] [CrossRef] [Green Version]
  85. Vaughn, S.F.; Berhow, M.A. Glucosinolate Hydrolysis Products from Various Plant Sources: pH Effects, Isolation, and Purification. Ind. Crops Prod. 2005, 21, 193–202. [Google Scholar] [CrossRef]
  86. Luang-In, V.; Rossiter, J.T. Stability Studies of Isothiocyanates and Nitriles in Aqueous Media. Songklanakarin J. Sci. Technol. 2015, 37, 625–630. [Google Scholar]
  87. Surugau, N.; Aripin, N. Effects of Temperature and pH on Myrosinase Activity and Gluconasturtiin Hydrolysis Products in Watercress. Trans. Sci. Technol. 2016, 3, 449–454. [Google Scholar]
  88. Li, Z. Development and Verification of Sulforaphane Extraction Method in Cabbage (Brassica oleracea L. Var. capitata) and Broccoli (Brassica oleracea L. Var. italica Planch.). J. Med. Plants Res. 2012, 6, 4796–4803. [Google Scholar] [CrossRef]
  89. Uda, Y.; Kurata, T.; Arakawa, N. Effects of pH and Ferrous Ion on the Degradation of Glucosinolates by Myrosinase. Agric. Biol. Chem. 1986, 50, 2735–2740. [Google Scholar] [CrossRef]
  90. Li, Z.; Liu, Y.; Fang, Z.; Yang, L.; Zhuang, M.; Zhang, Y.; Lv, H. Natural Sulforaphane from Broccoli Seeds against Influenza A Virus Replication in MDCK Cells. Nat. Prod. Commun. 2019, 14, 1–8. [Google Scholar] [CrossRef] [Green Version]
  91. Liu, X.; Wang, Y.; Hoeflinger, J.L.; Neme, B.P.; Jeffery, E.H.; Miller, M.J. Dietary Broccoli Alters Rat Cecal Microbiota to Improve Glucoraphanin Hydrolysis to Bioactive Isothiocyanates. Nutrients 2017, 9, 262. [Google Scholar] [CrossRef] [Green Version]
  92. Dufour, V.; Stahl, M.O.; Baysse, C. The antibacterial properties of isothiocyanates. Microbiology. 2015, 161, 229–243. [Google Scholar] [CrossRef] [Green Version]
  93. Li, X.; Kushad, M.M. Purification and Characterization of Myrosinase from Horseradish (Armoracia rusticana) Roots. Plant Physiol. Biochem. 2005, 43, 503–511. [Google Scholar] [CrossRef] [PubMed]
  94. Kassie, F.; Rabot, S.; Uhl, M.; Huber, W.; Qin, H.M.; Helma, C.; Schulte-Hermann, R.; Knasmüller, S. Chemoprotective Effects of Garden Cress (Lepidium sativum) and Its Constituents towards 2-Amino-3-Methyl-Imidazo [4,5-f] Quinoline (IQ)-Induced Genotoxic Effects and Colonic Preneoplastic Lesions. Carcinogenesis 2002, 23, 1155–1161. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  95. Yokoyama, S.-I.; Kodera, M.; Hirai, A.; Nakada, M.; Ueno, Y.; Osawa, T. Benzyl isothiocyanate produced by garden cress (Lepidium sativum) prevents accumulation of hepatic lipids. J. Nutr. Sci. Vitaminol. 2020, 66, 481–487. [Google Scholar] [CrossRef] [PubMed]
  96. Nakamura, Y.; Yoshimoto, M.; Murata, Y.; Shimoishi, Y.; Asai, Y.; Park, E.Y.; Sato, K.; Nakamura, Y. Papaya Seed Represents a Rich Source of Biologically Active Isothiocyanate. J. Agric. Food Chem. 2007, 55, 4407–4413. [Google Scholar] [CrossRef]
  97. Guzmán-Pérez, V.; Bumke-Vogt, C.; Schreiner, M.; Mewis, I.; Borchert, A.; Pfeiffer, A.F.H. Benzylglucosinolate Derived Isothiocyanate from Tropaeolum majus Reduces Gluconeogenic Gene and Protein Expression in Human Cells. PLoS ONE 2016, 11, e0162397. [Google Scholar] [CrossRef]
  98. Coscueta, E.R.; Reis, C.A.; Pintado, M. Phenylethyl Isothiocyanate Extracted from Watercress By-Products with Aqueous Micellar Systems: Development and Optimisation. Antioxidants 2020, 9, 698. [Google Scholar] [CrossRef]
  99. Prakash, O.; Rai, A.K.; Singh, J.; Singh, P.M. Partial purification and kinetic properties of myrosinase from cauliflower (Brassica oleracea var. botrytis). Indian J. Agric. Biochem. 2013, 26, 190–194. [Google Scholar]
  100. Román, J.; Castillo, A.; Cottet, L.; Mahn, A. Kinetic and Structural Study of Broccoli Myrosinase and Its Interaction with Different Glucosinolates. Food Chem. 2018, 254, 87–94. [Google Scholar] [CrossRef]
  101. Monsterrat, E. Mechanisms Underlying Biological Effects of Cruciferous Glucosinolate-Derived Isothiocyanates/Indoles: A Focus on Metabolic Syndrome. Front. Nutr. 2020, 7, 111. [Google Scholar] [CrossRef]
  102. Omri Hichri, A.; Mosbah, H.; Majouli, K.; Besbes Hlila, M.; Ben Jannet, H.; Flamini, G.; Aouni, M.; Selmi, B. Chemical Composition and Biological Activities of Eruca vesicaria Subsp. longirostris Essential Oils. Pharm. Biol. 2016, 54, 2236–2243. [Google Scholar] [CrossRef] [Green Version]
  103. Miyazawa, M.; Kawata, J. Identification of the Main Aroma Compounds in Dried Seeds of Brassica Hirta. J. Nat. Med. 2006, 60, 89–92. [Google Scholar] [CrossRef]
  104. Matusheski, N.V.; Jeffery, E.H. Comparison of the Bioactivity of Two Glucoraphanin Hydrolysis Products Found in Broccoli, Sulforaphane and Sulforaphane Nitrile. J. Agric. Food Chem. 2001, 49, 5743–5749. [Google Scholar] [CrossRef] [PubMed]
  105. Van Ommen Kloeke, A.E.E.; Jager, T.; van Gestel, C.A.M.; Ellers, J.; Pomeren, M.; van Krommenhoek, T.; Styrishave, B.; Hansen, M.; Roelofs, D. Time-Related Survival Effects of Two Gluconasturtiin Hydrolysis Products on the Terrestrial Isopod Porcellio Scaber. Chemosphere 2012, 89, 1084–1090. [Google Scholar] [CrossRef] [PubMed]
  106. Kyriakou, S.; Tragkola, V.; Alghol, H.; Antestopoulos, I.; Amery, T.; Stewart, K.; Winyard, G.P.; Trafalis, T.D.; Franco, R.; Pappa, A.; et al. Evaluation of Bioctive Properties of Lipophilic Fractions of Edible and Non-Edible Parts of Nasturtium officinale (Watercress) in a Model of Human Malignant Melanoma Cells. Pharmaceuticals 2022, 15, 141. [Google Scholar] [CrossRef]
  107. Wielanek, M.; Urbanek, H. Enhanced Glucotropaeolin Production in Hairy Root Cultures of Tropaeolum Majus L. by Combining Elicitation and Precursor Feeding. Plant Cell. Tissue Organ Cult. 2006, 86, 177–186. [Google Scholar] [CrossRef]
  108. Herzallah, S.; Lledó, M.L.; Holley, R. Influence of NaCl and NaNO3 on Sinigrin Hydrolysis by Foodborne Bacteria. J. Food Prot. 2011, 74, 2162–2168. [Google Scholar] [CrossRef]
  109. Rouzaud, G.; Young, S.A.; Duncan, A.J. Hydrolysis of glucosinolates to isothiocyanates after ingestion of raw or microwaved cabbage by human volunteers. Cancer Epidemiol. Biomark. Prev. 2004, 13, 125–131. [Google Scholar] [CrossRef] [Green Version]
  110. Connolly, E.L.; Sim, M.; Travica, N.; Marx, W.; Beasy, G.; Lynch, G.S.; Bondonno, C.P.; Lewis, J.R.; Hodgson, J.M.; Blekkenhorst, L.C. Glucosinolates from Cruciferous Vegetables and Their Potential Role in Chronic Disease: Investigating the Preclinical and Clinical Evidence. Front. Pharmacol. 2021, 12, 2964. [Google Scholar] [CrossRef]
  111. Tarar, A.; Alyami, E.M.; Peng, C.-A. Eradication of Myrosinase-Tethered Cancer Cells by Allyl Isothiocyanate Derived from Enzymatic Hydrolysis of Sinigrin. Pharmaceutics 2022, 14, 144. [Google Scholar] [CrossRef]
  112. Okunade, O.A.; Ghawi, S.K.; Methven, L.; Niranjan, K. Thermal and Pressure Stability of Myrosinase Enzymes from Black Mustard (Brassica nigra L. W.D.J. Koch. Var. nigra), Brown Mustard (Brassica juncea L. Czern. Var. juncea) and Yellow Mustard (Sinapsis alba L. Subsp. maire) Seeds. Food Chem. 2015, 187, 485–490. [Google Scholar] [CrossRef]
  113. Pérez, C.; Barrientos, H.; Roman, J.; Mahn, A. Optimization of a Blanching Step to Maximize Sulforaphane Synthesis in Broccoli Florets. Food Chem. 2014, 145, 264–271. [Google Scholar] [CrossRef] [PubMed]
  114. Van Eylen, D.; Oey, I.; Hendrickx, M.; Van Loey, A. Kinetics of the Stability of Broccoli (Brassica oleracea Cv italica) Myrosinase and Isothiocyanates in Broccoli Juice during Pressure/Temperature Treatments. J. Agric. Food Chem. 2007, 55, 2163–2170. [Google Scholar] [CrossRef] [PubMed]
  115. Mahn, A.; Angulo, A.; Cabañas, F. Purification and Characterization of Broccoli (Brassica oleracea Var italica) Myrosinase (β-Thioglucosidase Glucohydrolase). J. Agric. Food Chem. 2014, 62, 11666–11671. [Google Scholar] [CrossRef] [PubMed]
  116. Oliviero, T.; Verkerk, R.; Van Boekel, M.A.J.S.; Dekker, M. Effect of Water Content and Temperature on Inactivation Kinetics of Myrosinase in Broccoli (Brassica oleracea Var. italica). Food Chem. 2014, 163, 197–201. [Google Scholar] [CrossRef]
  117. Akpolat, H.; Barringer, S. The Effect of pH and Temperature on Cabbage Volatiles During Storage. J. Food Sci. 2015, 80, S1878–S1884. [Google Scholar] [CrossRef]
  118. Barba, F.J.; Nikmaram, N.; Roohinejad, S.; Khelfa, A.; Zhu, Z.; Koubaa, M. Bioavailability of Glucosinolates and their Breakdown Products: Impact of processing. Front. Nutr. 2016, 3, 24. [Google Scholar] [CrossRef] [Green Version]
  119. Kong, X.Y.; Kissen, R.; Bones, A.M. Characterization of Recombinant Nitrile-Specifier Proteins (NSPs) of Arabidopsis thaliana: Dependency on Fe(II) Ions and the Effect of Glucosinolate Substrate and Reaction Conditions. Phytochemistry 2012, 84, 7–17. [Google Scholar] [CrossRef]
  120. Van Dam, N.M.; Tytgat, T.O.G.; Kirkegaard, J.A. Root and Shoot Glucosinolates: A Comparison of Their Diversity, Function and Interactions in Natural and Managed Ecosystems. Phytochem. Rev. 2009, 8, 171–186. [Google Scholar] [CrossRef] [Green Version]
  121. Van Eylen, D.; Hendrickx, M.; Van Loey, A. Temperature and Pressure Stability of Mustard Seed (Sinapis alba L.) Myrosinase. Food Chem. 2006, 97, 263–271. [Google Scholar] [CrossRef]
  122. Nakamura, T.; Murata, Y.; Nakamura, Y. Characterization of Benzyl Isothiocyanate Extracted from Mashed Green Papaya by Distillation. Food Chem. 2019, 299, 125118. [Google Scholar] [CrossRef]
  123. Wittstock, U.; Meier, K.; Dörr, F.; Ravindran, B.M. NSP-Dependent Simple Nitrile Formation Dominates upon Breakdown of Major Aliphatic Glucosinolates in Roots, Seeds, and Seedlings of Arabidopsis thaliana Columbia-0. Front. Plant Sci. 2016, 7, 1821. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  124. Burow, M.; Losansky, A.; Müller, R.; Plock, A.; Kliebenstein, D.J.; Wittstock, U. The genetic basis of constitutive and herbivore-induced ESP-independent nitrile formation in Arabidopsis. Plant Physiol. 2009, 149, 561–574. [Google Scholar] [CrossRef] [Green Version]
  125. Eisenschmidt-Bönn, D.; Schneegans, N.; Backenköhler, A.; Wittstock, U.; Brandt, W. Structural Diversification during Glucosinolate Breakdown: Mechanisms of Thiocyanate, Epithionitrile and Simple Nitrile Formation. Plant J. 2019, 99, 329–343. [Google Scholar] [CrossRef] [PubMed]
  126. de Torres Zabala, M.; Grant, M.; Bones, A.M.; Bennett, R.; Lim, Y.S.; Kissen, R.; Rossiter, J.T. Characterisation of Recombinant Epithiospecifier Protein and Its Over-Expression in Arabidopsis thaliana. Phytochemistry 2005, 66, 859–867. [Google Scholar] [CrossRef]
  127. Burow, M.; Markert, J.; Gershenzon, J.; Wittstock, U. Comparative Biochemical Characterization of Nitrile-Forming Proteins from Plants and Insects That Alter Myrosinase-Catalysed Hydrolysis of Glucosinolates. FEBS J. 2006, 273, 2432–2446. [Google Scholar] [CrossRef] [PubMed]
  128. Guo, Q.; Guo, L.; Wang, Z.; Zhuang, Y.; Gu, Z. Response Surface Optimization and Identification of Isothiocyanates Produced from Broccoli Sprouts. Food Chem. 2013, 141, 1580–1586. [Google Scholar] [CrossRef]
  129. Doheny-Adams, T.; Redeker, K.; Kittipol, V.; Bancroft, I.; Hartley, S.E. Development of an Efficient Glucosinolate Extraction Method. Plant Methods 2017, 13, 17. [Google Scholar] [CrossRef] [Green Version]
  130. Mohd Zul, S.; Surugau, N. Effects of Ascorbic Acid and Ferum Ions Concentration on the Hydrolysis of Glucosinolate and Myrosinase Activity in the Watercress (Nasturtium officinale sp.). J. Teknol. 2016, 78, 133–138. [Google Scholar] [CrossRef] [Green Version]
  131. Novotny, C.; Schulzova, V.; Krmela, A.; Hajslova, J.; Svobodova, K.; Koudela, M. Ascorbic Acid and Glucosinolate Levels in New Czech Cabbage Cultivars: Effect of Production System and Fungal Infection. Molecules 2018, 23, 1855. [Google Scholar] [CrossRef] [Green Version]
  132. Liang, H.; Yuan, Q.; Xiao, Q. Effects of Metal Ions on Myrosinase Activity and the Formation of Sulforaphane in Broccoli Seed. J. Mol. Catal. B Enzym. 2006, 43, 19–22. [Google Scholar] [CrossRef]
  133. Gu, Z.; GUO, Q.; GU, Y. Factors Influencing Glucoraphanin and Sulforaphane Formation in Brassica Plants: A Review. J. Integr. Agric. 2012, 11, 1804–1816. [Google Scholar] [CrossRef]
  134. Gu, Y.; Guo, Q.; Zhang, L.; Chen, Z.; Han, Y.; Gu, Z. Physiological and Biochemical Metabolism of Germinating Broccoli Seeds and Sprouts. J. Agric. Food Chem. 2012, 60, 209–213. [Google Scholar] [CrossRef]
  135. Ludikhuyze, L.; Rodrigo, L.; Hendrickx, M. The Activity of Myrosinase from Broccoli (Brassica oleracea L. cv Italica): Influence of Intrinsic and Extrinsic Factors. J Food Prot. 2000, 63, 400–403. [Google Scholar] [CrossRef] [PubMed]
  136. Bones, A.M.; Rossiter, J.T. The Enzymic and Chemically Induced Decomposition of Glucosinolates. Phytochemistry 2006, 67, 1053–1067. [Google Scholar] [CrossRef] [PubMed]
  137. Kleinwachter, M.; Selmar, D. A novel approach for reliable activity determination of ascorbic acid depending myrosinases. Biochem. Biophys. Methods 2004, 59, 253–265. [Google Scholar] [CrossRef]
  138. Tian, S.; Liu, X.; Lei, P.; Zhang, X.; Shan, Y. Microbiota: A Mediator to Transform Glucosinolate Precursors in Cruciferous Vegetables to the Active Isothiocyanates. J. Sci. Food Agric. 2018, 98, 1255–1260. [Google Scholar] [CrossRef]
  139. Han, D.; Row, K.H. Separation and Purification of Sulforaphane from Broccoli by Solid Phase Extraction. Int. J. Mol. Sci. 2011, 12, 1854–1861. [Google Scholar] [CrossRef] [Green Version]
  140. Ares, A.M.; Bernal, J.; Martín, M.T.; Bernal, J.L.; Nozal, M.J. Optimized Formation, Extraction, and Determination of Sulforaphane in Broccoli by Liquid Chromatography with Diode Array Detection. Food Anal. Methods 2014, 7, 730–740. [Google Scholar] [CrossRef]
  141. Campas-Baypoli, O.N.; Sánchez-Machado, D.I.; Bueno-Solano, C.; Ramírez-Wong, B.; López-Cervantes, J. HPLC Method Validation for Measurement of Sulforaphane Level in Broccoli By-Products. Biomed. Chromatogr. 2010, 24, 387–392. [Google Scholar] [CrossRef]
  142. Suresh, S.; Waly, M.I.; Rahman, M.S. Broccoli (Brassica oleracea) as a Preventive Biomaterial for Cancer. In Bioactive Components, Diet and Medical Treatment in Cancer Prevention; Waly, M.I., Rahman, M.S., Eds.; Springer International Publishing: Cham, Switzerland, 2018; pp. 75–87. [Google Scholar] [CrossRef]
  143. Abdulah, R.; Faried, A.; Kobayashi, K.; Yamazaki, C.; Suradji, E.W.; Ito, K.; Suzuki, K.; Murakami, M.; Kuwano, H.; Koyama, H. Selenium Enrichment of Broccoli Sprout Extract Increases Chemosensitivity and Apoptosis of LNCaP Prostate Cancer Cells. BMC Cancer 2009, 9, 414. [Google Scholar] [CrossRef] [Green Version]
  144. Martínez-Hernández, G.B.; Artés-Hernández, F.; Gómez, P.A.; Artés, F. Induced Changes in Bioactive Compounds of Kailan-Hybrid Broccoli after Innovative Processing and Storage. J. Funct. Foods 2013, 5, 133–143. [Google Scholar] [CrossRef]
  145. Guo, R.; Yuan, G.; Wang, Q. Effect of Sucrose and Mannitol on the Accumulation of Health-Promoting Compounds and the Activity of Metabolic Enzymes in Broccoli Sprouts. Sci. Hortic. 2011, 128, 159–165. [Google Scholar] [CrossRef]
  146. Arora, R.; Vig, A.; Arora, S. Glucosinolates: Transposing Trends of Identification Methods from Paper Chromatography to Microchip Analysis. Int. J. Life Sc. Bt Pharma Res. 2014, 3, 42–61. [Google Scholar]
  147. Hanschen, F.S.; Brüggemann, N.; Brodehl, A.; Mewis, I.; Schreiner, M.; Rohn, S.; Kroh, L.W. Characterization of Products from the Reaction of Glucosinolate-Derived Isothiocyanates with Cysteine and Lysine Derivatives Formed in Either Model Systems or Broccoli Sprouts. J. Agric. Food Chem. 2012, 60, 7735–7745. [Google Scholar] [CrossRef] [PubMed]
  148. Guo, R.F.; Yuan, G.F.; Wang, Q.M. Effect of NaCl Treatments on Glucosinolate Metabolism in Broccoli Sprouts. J. Zhejiang Univ. Sci. B 2013, 14, 124–131. [Google Scholar] [CrossRef] [PubMed]
  149. Cuomo, V.; Luciano, F.; Meca, G.; Ritieni, A.; Mañes, J. Bioaccessibility of Glucoraphanin from Broccoli Using an in Vitro Gastrointestinal Digestion Model. CyTA-J. Food. 2014, 13, 361–365. [Google Scholar] [CrossRef] [Green Version]
  150. Vergara, F.; Wenzler, M.; Hansen, B.G.; Kliebenstein, D.J.; Halkier, B.A.; Gershenzon, J.; Schneider, B. Determination of the Absolute Configuration of the Glucosinolate Methyl Sulfoxide Group Reveals a Stereospecific Biosynthesis of the Side Chain. Phytochemistry 2008, 69, 2737–2742. [Google Scholar] [CrossRef]
  151. Van Eylen, D.; Bellostas, N.; Strobel, B.W.; Oey, I.; Hendrickx, M.; Van Loey, A.; Sørensen, H.; Sørensen, J.C. Influence of Pressure/Temperature Treatments on Glucosinolate Conversion in Broccoli (Brassica oleraceae L. Cv italica) Heads. Food Chem. 2009, 112, 646–653. [Google Scholar] [CrossRef]
  152. Liang, H.; Li, C.; Yuan, Q.; Vriesekoop, F. Separation and Purification of Sulforaphane from Broccoli Seeds by Solid Phase Extraction and Preparative High-Performance Liquid Chromatography. J. Agric. Food Chem. 2007, 55, 8047–8053. [Google Scholar] [CrossRef]
  153. Farag, M.A.; Motal, A.A.A. Sulforaphane Composition, Cytotoxic and Antioxidant Activity of Crucifer Vegetables. J. Adv. Res. 2010, 1, 65–70. [Google Scholar] [CrossRef] [Green Version]
  154. Ares, A.M.; Nozal, M.J.; Bernal, J. Extraction, Chemical Characterization and Biological Activity Determination of Broccoli Health Promoting Compounds. J. Chromatogr. A. 2013, 1313, 78–95. [Google Scholar] [CrossRef] [PubMed]
  155. Song, L.; Iori, R.; Thornalley, P.J. Purification of Major Glucosinolates from Brassicaceae Seeds and Preparation of Isothiocyanate and Amine Metabolites. J. Sci. Food Agric. 2006, 86, 1271–1280. [Google Scholar] [CrossRef]
  156. Sangkret, S.; Pongmalai, P.; Devahastin, S.; Chiewchan, N. Enhanced Production of Sulforaphane by Exogenous Glucoraphanin Hydrolysis Catalyzed by Myrosinase Extracted from Chinese Flowering Cabbage (Brassica rapa var. parachinensis). Sci. Rep. 2019, 9, 9882. [Google Scholar] [CrossRef]
  157. Zhang, B.; Wang, X.; Yang, Y.; Zhang, X. Extraction and Identification of Isothiocyanates from Broccolini Seeds. Nat. Prod. Commun. 2011, 6, 65–66. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  158. Miao, H.; Wang, J.; Cai, C.; Chang, J.; Zhao, Y.; Wang, Q. Accumulation of Glucosinolates in Broccoli. In Glucosinolates; Mérillon, J.-M., Ramawat, K.G., Eds.; Springer International Publishing: Cham, Switzerland, 2016; pp. 1–30. [Google Scholar] [CrossRef]
  159. Campas-Baypoli, O.N.; Bueno-Solano, C.; Martínez-Ibarra, D.M.; Camacho-Gil, F.; Villa-Lerma, A.G.; Rodríguez-Núñez, J.R.; Lóez-Cervantes, J.; Sánchez-Machado, D.I. Sulforaphane (1-isothiocyanato-4-(methylsulfinyl)-butane) content in cruciferous vegetables. Arch. Latinoam. Nutr. 2009, 59, 95–100. [Google Scholar] [PubMed]
  160. Totušek, J.; Tříska, J.; Lefnerová, D.; Strohalm, J.; Vrchotová, N.; Zendulka, O.; Průchová, J.; Chaloupková, J.; Novotná, P.; Houška, M. Contents of Sulforaphane and Total Isothiocyanates, Antimutagenic Activity, and Inhibition of Clastogenicity in Pulp Juices from Cruciferous Plants. Czech J. Food Sci. 2011, 29, 548–556. [Google Scholar] [CrossRef] [Green Version]
  161. Langston-Cox, A.; Anderson, D.; Creek, D.J.; Palmer, K.; Wallace, E.M.; Marshall, S.A. Measuring Sulforaphane and Its Metabolites in Human Plasma: A High Throughput Method. Molecules. 2020, 25, 829. [Google Scholar] [CrossRef] [Green Version]
  162. Hafezian, S.M.; Azizi, S.N.; Biparva, P.; Bekhradnia, A. High-Efficiency Purification of Sulforaphane from the Broccoli Extract by Nanostructured SBA-15 Silica Using Solid-Phase Extraction Method. J. Chromatogr. B. 2019, 1108, 1–10. [Google Scholar] [CrossRef]
  163. Blažević, I.; Montaut, S.; Burčul, F.; Olsen, C.E.; Burow, M.; Rollin, P.; Agerbirk, N. Glucosinolate Structural Diversity, Identification, Chemical Synthesis and Metabolism in Plants. Phytochemistry 2020, 169, 112100. [Google Scholar] [CrossRef]
  164. Fahey, J.W.; Zalcmann, A.T.; Talalay, P. The Chemical Diversity and Distribution of Glucosinolates and Isothiocyanates among Plants. Phytochemistry 2001, 56, 5–51. [Google Scholar] [CrossRef]
  165. Kassie, F.; Uhl, M.; Rabot, S.; Grasl-Kraupp, B.; Verkerk, R.; Kundi, M.; Chabicovsky, M.; Schulte-Hermann, R.; Knasmüller, S. Chemoprevention of 2-Amino-3-Methylimidazo[4,5- f ]Quinoline (IQ)-Induced Colonic and Hepatic Preneoplastic Lesions in the F344 Rat by Cruciferous Vegetables Administered Simultaneously with the Carcinogen. Carcinogenesis 2003, 24, 255–261. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  166. Jadhav, U.; Ezhilarasan, R.; Vaughn, S.F.; Berhow, M.A.; Mohanam, S. Iberin Induces Cell Cycle Arrest and Apoptosis in Human Neuroblastoma Cells. Int. J. Mol. Med. 2007, 19, 353–361. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  167. Jadhav, U.; Ezhilarasan, R.; Vaughn, S.; Berhow, M.; Mohanam, S. Dietary Isothiocyanate Iberin Inhibits Growth and Induces Apoptosis in Human Glioblastoma Cells. J. Pharmacol. Sci. 2007, 103, 247–251. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  168. Tan, S.Y.-Y.; Liu, Y.; Chua, S.L.; Vejborg, R.M.; Jakobsen, T.H.; Chew, S.C.; Li, Y.; Nielsen, T.E.; Tolker-Nielsen, T.; Yang, L.; et al. Comparative Systems Biology Analysis to Study the Mode of Action of the Isothiocyanate Compound Iberin on Pseudomonas Aeruginosa. Antimicrob. Agents Chemother. 2014, 58, 6648–6659. [Google Scholar] [CrossRef] [Green Version]
  169. Mitsiogianni, M.; Trafalis, D.; Franco, R.; Zoumpourlis, V.; Pappa, A.; Panayiotidis, M. Sulforaphane and Iberin Are Potent Epigenetic Modulators of Histone Acetylation and Methylation in Malignant Melanoma. Eur. J. Nutr. 2021, 60, 147–158. [Google Scholar] [CrossRef]
  170. Cirilli, R.; Gallo, F.R.; Multari, G.; Palazzino, G.; Mustazza, C.; Panusa, A. Study of Solvent Effect on the Stability of Isothiocyanate Iberin, a Breakdown Product of Glucoiberin. J. Food Compos. Anal. 2020, 92, 103515. [Google Scholar] [CrossRef]
  171. Deng, Q.; Zinoviadou, K.G.; Galanakis, C.M.; Orlien, V.; Grimi, N.; Vorobiev, E.; Lebovka, N.; Barba, F.J. The Effects of Conventional and Non-Conventional Processing on Glucosinolates and Its Derived Forms, Isothiocyanates: Extraction, Degradation, and Applications. Food Eng. Rev. 2015, 7, 357–381. [Google Scholar] [CrossRef]
  172. Matusheski, N.V.; Wallig, M.A.; Juvik, J.A.; Klein, B.P.; Kushad, M.M.; Jeffery, E.H. Preparative HPLC Method for the Purification of Sulforaphane and Sulforaphane Nitrile from Brassica oleracea. J. Agric. Food Chem. 2001, 49, 1867–1872. [Google Scholar] [CrossRef]
  173. Tian, G.; Tang, P.; Xie, R.; Cheng, L.; Yuan, Q.; Hu, J. The Stability and Degradation Mechanism of Sulforaphene in Solvents. Food Chem. 2016, 199, 301–306. [Google Scholar] [CrossRef]
  174. Tian, G.; Li, Y.; Cheng, L.; Yuan, Q.; Tang, P.; Kuang, P.; Hu, J. The Mechanism of Sulforaphene Degradation to Different Water Contents. Food Chem. 2016, 194, 1022–1027. [Google Scholar] [CrossRef]
  175. Conaway, C.C.; Chung, Y.Y. and F. Isothiocyanates as Cancer Chemopreventive Agents: Their Biological Activities and Metabolism in Rodents and Humans. Curr. Drug Metabol. 2002, 3, 233–255. [Google Scholar] [CrossRef] [PubMed]
  176. Zhang, H.; Sun, S.; Chen, Q.S.; Chai, Y.F.; Sun, D.X.; Zhang, G.Q. Analysis of Sulforaphane by HPLC-MS/MS in Vitro and in Vivo: Chemical Stability, Metabolic Rate and Metabolites. Asian J. Chem. 2015, 27, 1045–1048. [Google Scholar] [CrossRef]
  177. Franklin, S.J.; Dickinson, S.E.; Karlage, K.L.; Bowden, G.T.; Myrdal, P.B. Stability of Sulforaphane for Topical Formulation. Drug Dev. Ind. Pharm. 2014, 40, 494–502. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  178. Galanakis, C.M. Separation of Functional Macromolecules and Micromolecules: From Ultrafiltration to the Border of Nanofiltration. Trends Food Sci. Technol. 2015, 42, 44–63. [Google Scholar] [CrossRef]
  179. Galanakis, C.M. Phenols Recovered from Olive Mill Wastewater as Additives in Meat Products. Trends Food Sci. Technol. 2018, 79, 98–105. [Google Scholar] [CrossRef]
  180. Langeveld, M.; Tan, C.Y.; Soeters, M.R.; Virtue, S.; Watson, L.P.; Murgatroyd, P.R.; Ambler, G.K.; Vidal-Puig, S.; Chatterjee, K.V.; Vidal-Puig, A. No Metabolic Effects of Mustard Allyl-Isothiocyanate Compared with Placebo in Men. Am. J. Clin. Nutr. 2017, 106, 1197–1205. [Google Scholar] [CrossRef] [Green Version]
  181. Subedi, L.; Venkatesan, R.; Kim, S.Y. Neuroprotective and Anti-Inflammatory Activities of Allyl Isothiocyanate through Attenuation of JNK/NF-κB/TNF-α Signaling. Int. J. Mol. Sci. 2017, 18, 1423. [Google Scholar] [CrossRef] [Green Version]
  182. Yun, Y.-K.; Kim, H.-K.; Kim, J.-R.; Hwang, K.; Ahn, Y.-J. Contact and Fumigant Toxicity of Armoracia rusticana Essential Oil, Allyl Isothiocyanate and Related Compounds to Dermatophagoides Farinae. Pest Manag. Sci. 2012, 68, 788–794. [Google Scholar] [CrossRef]
  183. Zhang, Y. Allyl Isothiocyanate as a Cancer Chemopreventive Phytochemical. Mol. Nutr. Food Res. 2010, 54, 127–135. [Google Scholar] [CrossRef] [Green Version]
  184. Bo, P.; Lien, J.-C.; Chen, Y.-Y.; Yu, F.-S.; Lu, H.-F.; Yu, C.-S.; Chou, Y.-C.; Yu, C.-C.; Chung, J.-G. Allyl Isothiocyanate Induces Cell Toxicity by Multiple Pathways in Human Breast Cancer Cells. Am. J. Chin. Med. 2016, 44, 415–437. [Google Scholar] [CrossRef]
  185. Chiang, J.-H.; Tsai, F.-J.; Hsu, Y.-M.; Yin, M.-C.; Chiu, H.-Y.; Yang, J.-S. Sensitivity of Allyl Isothiocyanate to Induce Apoptosis via ER Stress and the Mitochondrial Pathway upon ROS Production in Colorectal Adenocarcinoma Cells. Oncol. Rep. 2020, 44, 1415–1424. [Google Scholar] [CrossRef] [PubMed]
  186. Sávio, A.L.V.; da Silva, G.N.; Salvadori, D.M.F. Inhibition of Bladder Cancer Cell Proliferation by Allyl Isothiocyanate (Mustard Essential Oil). Mutat. Res. 2015, 771, 29–35. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  187. Bhattacharya, A.; Li, Y.; Wade, K.L.; Paonessa, J.D.; Fahey, J.W.; Zhang, Y. Allyl Isothiocyanate-Rich Mustard Seed Powder Inhibits Bladder Cancer Growth and Muscle Invasion. Carcinogenesis 2010, 31, 2105–2110. [Google Scholar] [CrossRef] [PubMed]
  188. Qin, G.; Li, P.; Xue, Z. Effect of Allyl Isothiocyanate on the Viability and Apoptosis of the Human Cervical Cancer HeLa Cell Line in Vitro. Oncol. Lett. 2018, 15, 8756–8760. [Google Scholar] [CrossRef]
  189. Dufour, V.; Alazzam, B.; Ermel, G.; Thepaut, M.; Rossero, A.; Tresse, O.; Baysse, C. Antimicrobial Activities of Isothiocyanates against Campylobacter Jejuni Isolates. Front. Cell. Infect. Microbiol. 2012, 2, 53. [Google Scholar] [CrossRef] [Green Version]
  190. Olivier, C.; Vaughn, S.F.; Mizubuti, E.S.G.; Loria, R. Variation in Allyl Isothiocyanate Production within Brassica Species and Correlation with Fungicidal Activity. J. Chem. Ecol. 1999, 25, 2687–2701. [Google Scholar] [CrossRef]
  191. Ahmed, R. Evaluation of Antimicrobial Activity of Allyl Isothiocyanate (AITC) Adsorbed in Oyster Shell on Food-Borne Bacteria. Clean Technol. 2015, 21, 241–247. [Google Scholar] [CrossRef] [Green Version]
  192. Kara, M.; Soylu, E.M. Assessment of Glucosinolate-Derived Isothiocyanates as Potential Natural Antifungal Compounds against Citrus Sour Rot Disease Agent Geotrichum Citri-Aurantii. J. Phytopathol. 2020, 168, 279–289. [Google Scholar] [CrossRef]
  193. Wu, H.; Zhang, X.; Zhang, G.-A.; Zeng, S.-Y.; Lin, K.-C. Antifungal Vapour-phase Activity of a Combination of Allyl Isothiocyanate and Ethyl Isothiocyanate Against Botrytis cinerea and Penicillium expansum Infection on Apples. J. Phytopathol. 2011, 159, 450–455. [Google Scholar] [CrossRef]
  194. Azaiez, I.; Meca, G.; Manyes, L.; Fernández-Franzón, M. Antifungal Activity of Gaseous Allyl, Benzyl and Phenyl Isothiocyanate in Vitro and Their Use for Fumonisins Reduction in Bread. Food Control 2013, 32, 428–434. [Google Scholar] [CrossRef]
  195. Manyes, L.; Luciano, F.B.; Mañes, J.; Meca, G. In Vitro Antifungal Activity of Allyl Isothiocyanate (AITC) against Aspergillus parasiticus and Penicillium expansum and Evaluation of the AITC Estimated Daily Intake. Food Chem. Toxicol. 2015, 83, 293–299. [Google Scholar] [CrossRef] [PubMed]
  196. Mao, L.; Wang, H.D.; Wang, X.L.; Qiao, L.; Yin, H.X. Sulforaphane Attenuates Matrix Metalloproteinase-9 Expression Following Spinal Cord Injury in Mice. Ann. Clin. Lab. Sci. 2010, 40, 354–360. [Google Scholar] [PubMed]
  197. Latronico, T.; Larocca, M.; Milella, S.; Fasano, A.; Rossano, R.; Liuzzi, G.M. Neuroprotective Potential of Isothiocyanates in an in Vitro Model of Neuroinflammation. Inflammopharmacology 2020, 29, 561–571. [Google Scholar] [CrossRef] [PubMed]
  198. Dai, R.; Lim, L.T. Release of Allyl Isothiocyanate from Mustard Seed Meal Powder. J. Food Sci. 2014, 79, E47–E53. [Google Scholar] [CrossRef] [PubMed]
  199. Cools, K.; Terry, L.A. Comparative Study between Extraction Techniques and Column Separation for the Quantification of Sinigrin and Total Isothiocyanates in Mustard Seed. J. Chromatogr. B. 2012, 901, 115–118. [Google Scholar] [CrossRef]
  200. Sharna, H.K.; Ingle, S.; Singh, C.; Sarkar, B.C.; Upadhyay, A. Effect of various process treatment conditions on the allyl isothiocyanate extraction rate from mustard meal. J. Food Sci. Technol. 2012, 49, 368–372. [Google Scholar] [CrossRef] [Green Version]
  201. Tripathi, M.K.; Mishra, A. Glucosinolates in Animal Nutrition: A Review. Anim. Feed Sci. Technol. 2007, 132, 1–27. [Google Scholar] [CrossRef]
  202. Li, Y.; Teng, Z.; Chen, P.; Song, Y.; Luo, Y.; Wang, Q. Enhancement of aqueous stability of allyl isothiocyanate using nanoemulsions prepared by an emulsion inversion point method. J. Colloid Interface Sci. 2014, 15, 130–137. [Google Scholar] [CrossRef]
  203. Liu, T.T.; Yang, T.S. Stability and antimicrobial activity of allyl isothiocyanate during long-term storage in an oil-in-water emulsion. J. Food Sci. 2010, 75, 445–451. [Google Scholar] [CrossRef]
  204. Nahar, L.; Sarker, S.D. Supercritical fluid extraction in natural products analyses. Methods Mol. Biol. 2012, 864, 43–74. [Google Scholar] [CrossRef]
  205. Moyler, D.A. Extraction of Flavours and Fragrances with Compressed CO2. In Extraction of Natural Products Using Near-Critical Solvents; King, M.B., Bott, T.R., Eds.; Springer: Dordrecht, The Netherlands, 1993; pp. 140–183. [Google Scholar] [CrossRef]
  206. Kraujalis, P.; Venskutonis, P.R. Optimisation of Supercritical Carbon Dioxide Extraction of Amaranth Seeds by Response Surface Methodology and Characterization of Extracts Isolated from Different Plant Cultivars. J. Supercrit. Fluids 2013, 73, 80–86. [Google Scholar] [CrossRef]
  207. Gracia, I.; Rodríguez, J.F.; García, M.T.; Alvarez, A.; García, A. Isolation of Aroma Compounds from Sugar Cane Spirits by Supercritical CO2. J. Supercrit. Fluids 2007, 43, 37–42. [Google Scholar] [CrossRef]
  208. Wang, Q.; Shi, A.; Liu, H.; Liu, L.; Zhang, Y.; Li, N.; Gong, K.; Yu, M.; Zheng, L. Peanut By-Products Utilization Technology. In Peanuts: Processing Technology and Product Development; Wang, Q., Ed.; Academic Press: Cambridge, MA, USA, 2016; pp. 211–325. [Google Scholar] [CrossRef]
  209. Lizcano, S.C.; Dávila, J.A.; Hernández, V. Fruit Agroindustrial Wastes for Preparing Beverages for Medicinal Purposes by Supercritical Fluid Extraction Technology: Andes Berry (Rubus glaucus benth) Case. In Production and Management of Beverages; Grumezescu, A.M., Holban, A.M., Eds.; Woodhead Publishing: Sawston, UK, 2019; pp. 151–177. [Google Scholar] [CrossRef]
  210. Kim, S.-J.; Lee, M.-K.; Back, S.-S.; Chun, B.-S. Extraction and Identification of Volatile Isothiocyanates from Wasabi Using Supercritical Carbon Dioxide. Korean J. Biotechnol. Bioeng 2007, 22, 174–178. [Google Scholar]
  211. Jain, A.; Ong, V.; Jayaraman, S.; Balasubramanian, R.; Srinivasan, M.P. Supercritical fluid immobilization of horseradish peroxidase on high surface area mesoporous activated carbon. J Supercrit Fluids 2016, 107, 513–518. [Google Scholar] [CrossRef]
  212. Szmigielska, A.M.; Schoenau, J.J. Use of Anion-Exchange Membrane Extraction for the High-Performance Liquid Chromatographic Analysis of Mustard Seed Glucosinolates. J. Agric. Food Chem. 2000, 48, 5190–5194. [Google Scholar] [CrossRef]
  213. Bennett, R.N.; Carvalho, R.; Mellon, F.A.; Eagles, J.; Rosa, E.A.S. Identification and Quantification of Glucosinolates in Sprouts Derived from Seeds of Wild Eruca sativa L. (Salad Rocket) and Diplotaxis tenuifolia L. (Wild Rocket) from Diverse Geographical Locations. J. Agric. Food Chem. 2007, 55, 67–74. [Google Scholar] [CrossRef]
  214. Powell, E.E.; Hill, G.A.; Juurlink, B.H.J.; Carrier, D.J. Glucoraphanin Extraction from Cardaria Draba: Optimization of Batch Extraction. J. Chem. Technol. Biotechnol. 2005, 80, 985–991. [Google Scholar] [CrossRef]
  215. Mohn, T.; Cutting, B.; Ernst, B.; Hamburger, M. Extraction and Analysis of Intact Glucosinolates—A Validated Pressurized Liquid Extraction/Liquid Chromatography–Mass Spectrometry Protocol for Isatis tinctoria, and Qualitative Analysis of Other Cruciferous Plants. J. Chromatogr. A 2007, 1166, 142–151. [Google Scholar] [CrossRef]
  216. Wang, T.; Liang, H.; Yuan, Q. Optimization of Ultrasonic-Stimulated Solvent Extraction of Sinigrin from Indian Mustard Seed (Brassica juncea L.) Using Response Surface Methodology. Phytochem. Anal. 2011, 22, 205–213. [Google Scholar] [CrossRef]
  217. Soares Melecchi, M.I.; Péres, V.F.; Dariva, C.; Zini, C.A.; Abad, F.C.; Martinez, M.M.; Caramão, E.B. Optimization of the sonication extraction method of Hibiscus tiliaceus L. flowers. Ultrason. Sonochem. 2006, 13, 242–250. [Google Scholar] [CrossRef]
  218. Huang, W.; Xue, A.; Niu, H.; Jia, Z.; Wang, J. Optimised Ultrasonic-Assisted Extraction of Flavonoids from Folium Eucommiae and Evaluation of Antioxidant Activity in Multi-Test Systems In Vitro. Food Chem. 2009, 114, 1147–1154. [Google Scholar] [CrossRef]
  219. Da Porto, C.; Decorti, D. Ultrasound-Assisted Extraction Coupled with under Vacuum Distillation of Flavour Compounds from Spearmint (Carvone-Rich) Plants: Comparison with Conventional Hydrodistillation. Ultrason. Sonochem. 2009, 16, 795–799. [Google Scholar] [CrossRef] [PubMed]
  220. Toma, M.; Vinatoru, M.; Paniwnyk, L.; Mason, T.J. Investigation of the Effects of Ultrasound on Vegetal Tissues during Solvent Extraction. Ultrason. Sonochem. 2001, 8, 137–142. [Google Scholar] [CrossRef]
  221. Sivakumar, V.; Ravi Verma, V.; Rao, P.G.; Swaminathan, G. Studies on the Use of Power Ultrasound in Solid–Liquid Myrobalan Extraction Process. J. Clean. Prod. 2007, 15, 1813–1818. [Google Scholar] [CrossRef]
  222. Boonkird, S.; Phisalaphong, C.; Phisalaphong, M. Ultrasound-Assisted Extraction of Capsaicinoids from Capsicum Frutescens on a Lab- and Pilot-Plant Scale. Ultrason. Sonochem. 2008, 15, 1075–1079. [Google Scholar] [CrossRef]
  223. Variyar, P.S.; Banerjee, A.; Akkarakaran, J.J.; Suprasanna, P. Role of Glucosinolates in Plant Stress Tolerance. In Emerging Technologies and Management of Crop Stress Tolerance; Ahmad, P., Rasool, S., Eds.; Academic Press: San Diego, CA, USA, 2014; pp. 271–291. [Google Scholar] [CrossRef]
  224. Jensen, J.; Styrishave, B.; Gimsing, A.L.; Bruun Hansen, H.C. The Toxic Effects of Benzyl Glucosinolate and Its Hydrolysis Product, the Biofumigant Benzyl Isothiocyanate, to Folsomia fimetaria. Environ. Toxicol. Chem. 2010, 29, 359–364. [Google Scholar] [CrossRef]
  225. Platz, S.; Kühn, C.; Schiess, S.; Schreiner, M.; Mewis, I.; Kemper, M.; Pfeiffer, A.; Rohn, S. Determination of Benzyl Isothiocyanate Metabolites in Human Plasma and Urine by LC-ESI-MS/MS after Ingestion of Nasturtium (Tropaeolum majus L.). Anal. Bioanal. Chem. 2013, 405, 7427–7436. [Google Scholar] [CrossRef]
  226. Zakaria, S.; Helmy, M.W.; Salahuddin, A.; Omran, G. Chemopreventive and Antitumor Effects of Benzyl Isothiocynate on HCC Models: A Possible Role of HGF /PAkt/ STAT3 Axis and VEGF. Biomed. Pharmacother. 2018, 108, 65–75. [Google Scholar] [CrossRef]
  227. Zhu, M.; Li, W.; Guo, J.; Lu, Y.; Dong, X.; Lin, B.; Chen, Y.; Zhang, X.; Li, M. Alpha Fetoprotein Antagonises Benzyl Isothiocyanate Inhibition of the Malignant Behaviors of Hepatocellular Carcinoma Cells. Oncotarget 2016, 7, 75749–75762. [Google Scholar] [CrossRef] [Green Version]
  228. Boreddy, S.R.; Sahu, R.P.; Srivastava, S.K. Benzyl Isothiocyanate Suppresses Pancreatic Tumor Angiogenesis and Invasion by Inhibiting HIF-α/VEGF/Rho-GTPases: Pivotal Role of STAT-3. PLoS ONE 2011, 6, e25799. [Google Scholar] [CrossRef]
  229. Sahu, R.P.; Srivastava, S.K. The Role of STAT-3 in the Induction of Apoptosis in Pancreatic Cancer Cells by Benzyl Isothiocyanate. J. Natl. Cancer Inst. 2009, 101, 176–193. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  230. Kamii, E.; Isshiki, K. Antimicrobial Efficacy of Benzyl Isothiocyanate. Shokuhin Eiseigaku Zasshi. 2009, 50, 311–314. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  231. Wang, T.; Li, Y.; Bi, Y.; Zhang, M.; Zhang, T.; Zheng, X.; Dong, Y.; Huang, Y. Benzyl Isothiocyanate Fumigation Inhibits Growth, Membrane Integrity and Mycotoxin Production in Alternaria alternata. RSC Adv. 2020, 10, 1829–1837. [Google Scholar] [CrossRef] [Green Version]
  232. Pereira, C.; Calado, A.M.; Sampaio, A.C. The Effect of Benzyl Isothiocyanate on Candida albicans Growth, Cell Size, Morphogenesis, and Ultrastructure. World J. Microbiol. Biotechnol. 2020, 36, 153. [Google Scholar] [CrossRef] [PubMed]
  233. Afsharypuor, S.; Hadi, M.-E. Volatile Constituents of the Seeds, Roots and Non-Flowering Aerial Parts of Lepidium satvium L. J. Essent. Oil Res. 2006, 18, 495–496. [Google Scholar] [CrossRef]
  234. Németh, A.G.; Keserű, G.M.; Ábrányi-Balogh, P. A Novel Three-Component Reaction between Isocyanides, Alcohols or Thiols and Elemental Sulfur: A Mild, Catalyst-Free Approach towards O-Thiocarbamates and Dithiocarbamates. Beilstein J. Org. Chem. 2019, 15, 1523–1533. [Google Scholar] [CrossRef] [Green Version]
  235. Nakamura, Y.; Miyoshi, N. Electrophiles in Foods: The Current Status of Isothiocyanates and Their Chemical Biology. Biosci. Biotechnol. Biochem. 2010, 74, 242–255. [Google Scholar] [CrossRef] [Green Version]
  236. Abdel-Kader, M.S.; Alam, P.; Kamal, Y.T.; Alkharfy, K.M.; Foudah, A.I.; Alqasoumi, S.I. Optimization of the Extraction Condition for Benzyl Isothiocyanate Contents in Salvadora persica Roots “Siwak”. Saudi Pharm. J. 2019, 27, 753–755. [Google Scholar] [CrossRef]
  237. Abdel-Kader, M.S.; Al Shahrani, K.S.; Alqarni, M.H.; Salkini, M.A.; Khamis, E.H.; Ghabbour, H.A.; Alqasoumi, S.I. Effect of Hydroxylated Solvents on the Active Constituents of Salvadora persica Root “Siwak”. Saudi Pharm. J. 2019, 27, 220–224. [Google Scholar] [CrossRef]
  238. Hegazi, G.; El-Hanafy, N.; Abu-Elkheir, Z.; Hussein, I. Benzyl Isothiocyanate Production from Salvadora persica L. Callus Cultures. IOSR J. Biotechnol. Biochem. 2016, 2, 19–25. [Google Scholar]
  239. De Nicola, G.R.; Nyegue, M.; Montaut, S.; Iori, R.; Menut, C.; Tatibouët, A.; Rollin, P.; Ndoyé, C.; Zollo, P.-H.A. Profile and Quantification of Glucosinolates in Pentadiplandra Brazzeana Baillon. Phytochemistry 2012, 73, 51–56. [Google Scholar] [CrossRef] [PubMed]
  240. Conaway, C.C.; Krzeminski, J.; Amin, S.; Chung, F.-L. Decomposition Rates of Isothiocyanate Conjugates Determine Their Activity as Inhibitors of Cytochrome P450 Enzymes. Chem. Res. Toxicol. 2001, 14, 1170–1176. [Google Scholar] [CrossRef] [PubMed]
  241. De Nicola, G.R.; Montaut, S.; Rollin, P.; Nyegue, M.; Menut, C.; Iori, R.; Tatibouët, A. Stability of Benzylic-Type Isothiocyanates in Hydrodistillation-Mimicking Conditions. J. Agric. Food Chem. 2013, 61, 137–142. [Google Scholar] [CrossRef]
  242. Oerlemans, K.; Barrett, D.M.; Suades, C.B.; Verkerk, R.; Dekker, M. Thermal Degradation of Glucosinolates in Red Cabbage. Food Chem. 2006, 95, 19–29. [Google Scholar] [CrossRef]
  243. Hanschen, F.S.; Rohn, S.; Mewis, I.; Schreiner, M.; Kroh, L.W. Influence of the Chemical Structure on the Thermal Degradation of the Glucosinolates in Broccoli Sprouts. Food Chem. 2012, 130, 1–8. [Google Scholar] [CrossRef]
  244. Hasapis, X.; MacLeod, A.J. Benzylglucosinolate Degradation in Heat-Treated Lepidium sativum Seeds and Detection of a Thiocyanate-Forming Factor. Phytochemistry 1982, 21, 1009–1013. [Google Scholar] [CrossRef]
  245. Li, N.; Xu, L. Thermal Analysis of β-Cyclodextrin/Berberine Chloride Inclusion Compounds. Thermochim. Acta 2010, 499, 166–170. [Google Scholar] [CrossRef]
  246. Li, W.; Liu, X.; Yang, Q.; Zhang, N.; Du, Y.; Zhu, H. Preparation and Characterization of Inclusion Complex of Benzyl Isothiocyanate Extracted from Papaya Seed with β-Cyclodextrin. Food Chem. 2015, 184, 99–104. [Google Scholar] [CrossRef]
  247. Uppal, S.; Kaur, K.; Kumar, R.; Kahlon, N.K.; Singh, R.; Mehta, S.K. Encompassment of Benzyl Isothiocyanate in Cyclodextrin Using Ultrasonication Methodology to Enhance Its Stability for Biological Applications. Ultrason. Sonochem. 2017, 39, 25–33. [Google Scholar] [CrossRef]
  248. Dayalan Naidu, S.; Suzuki, T.; Yamamoto, M.; Fahey, J.W.; Dinkova-Kostova, A.T. Phenethyl Isothiocyanate, a Dual Activator of Transcription Factors NRF2 and HSF1. Mol. Nutr. Food Res. 2018, 62, 1700908. [Google Scholar] [CrossRef] [Green Version]
  249. Sarkar, R.P.; Mukherjee, S.; Biswas, J.; Roy, M. Phenethyl Isothiocyanate, by Virtue of Its Antioxidant Activity, Inhibits Invasiveness and Metastatic Potential of Breast Cancer Cells: HIF-1α as a Putative Target. Free Radic. Res. 2016, 50, 84–100. [Google Scholar] [CrossRef] [PubMed]
  250. Hoffman, J.D.; Ward, W.M.; Loo, G. Effect of Antioxidants on the Genotoxicity of Phenethyl Isothiocyanate. Mutagenesis 2015, 30, 421–430. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  251. Yazdanparast, R.; Bahramikia, S.; Ardestani, A. Nasturtium officinale Reduces Oxidative Stress and Enhances Antioxidant Capacity in Hypercholesterolaemic Rats. Chem. Biol. Interact. 2008, 172, 176–184. [Google Scholar] [CrossRef] [PubMed]
  252. Chen, P.-Y.; Lin, K.-C.; Lin, J.-P.; Tang, N.-Y.; Yang, J.-S.; Lu, K.-W.; Chung, J.-G. Phenethyl Isothiocyanate (PEITC) Inhibits the Growth of Human Oral Squamous Carcinoma HSC-3 Cells through G 0/G 1 Phase Arrest and Mitochondria-Mediated Apoptotic Cell Death. Evidence-Based Complement. Altern. Med. 2012, 2012, 718320. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  253. Koschorke, A.; Faraci, S.; Giani, D.; Chiodoni, C.; Iorio, E.; Canese, R.; Colombo, M.P.; Lamolinara, A.; Iezzi, M.; Ladomery, M.; et al. Phenethyl Isothiocyanate Hampers Growth and Progression of HER2-Positive Breast and Ovarian Carcinoma by Targeting Their Stem Cell Compartment. Cell. Oncol. 2019, 42, 815–828. [Google Scholar] [CrossRef]
  254. Lawson, A.P.; Long, M.J.C.; Coffey, R.T.; Qian, Y.; Weerapana, E.; El Oualid, F.; Hedstrom, L. Naturally Occurring Isothiocyanates Exert Anticancer Effects by Inhibiting Deubiquitinating Enzymes. Cancer Res. 2015, 75, 5130–5142. [Google Scholar] [CrossRef] [Green Version]
  255. Zhang, T.; Zhang, W.; Hao, M. Phenethyl Isothiocyanate Reduces Breast Cancer Stem Cell-like Properties by Epigenetic Reactivation of CDH1. Oncol Rep. 2021, 45, 337–348. [Google Scholar] [CrossRef]
  256. Bommareddy, A.; Hahm, E.-R.; Xiao, D.; Powolny, A.A.; Fisher, A.L.; Jiang, Y.; Singh, S.V. Atg5 Regulates Phenethyl Isothiocyanate–Induced Autophagic and Apoptotic Cell Death in Human Prostate Cancer Cells. Cancer Res. 2009, 69, 3704–3712. [Google Scholar] [CrossRef] [Green Version]
  257. Boggs, D.A.; Palmer, J.R.; Wise, L.A.; Spiegelman, D.; Stampfer, M.J.; Adams-Campbell, L.L.; Rosenberg, L. Fruit and Vegetable Intake in Relation to Risk of Breast Cancer in the Black Women’s Health Study. Am. J. Epidemiol. 2010, 172, 1268–1279. [Google Scholar] [CrossRef]
  258. Gupta, P.; Kim, B.; Kim, S.-H.; Srivastava, S.K. Molecular Targets of Isothiocyanates in Cancer: Recent Advances. Mol. Nutr. Food Res. 2014, 58, 1685–1707. [Google Scholar] [CrossRef] [Green Version]
  259. Chen, Y.; Li, Y.; Wang, X.; Meng, Y.; Zhang, Q.; Zhu, J.; Chen, J.; Cao, W.; Wang, X.; Xie, C.; et al. Phenethyl Isothiocyanate Inhibits Colorectal Cancer Stem Cells by Suppressing Wnt/β-Catenin Pathway. Phyther. Res. 2018, 32, 2447–2455. [Google Scholar] [CrossRef] [PubMed]
  260. Jaya Seema, D.M.; Saifullah, B.; Selvanayagam, M.; Gothai, S.; Hussein, M.Z.; Subbiah, S.K.; Mohd Esa, N.; Arulselvan, P. Designing of the Anticancer Nanocomposite with Sustained Release Properties by Using Graphene Oxide Nanocarrier with Phenethyl Isothiocyanate as Anticancer Agent. Pharmaceutics 2018, 10, 109. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  261. Yang, C.-X.; Wu, H.-T.; Li, X.-X.; Wu, H.-Y.; Niu, T.-X.; Wang, X.-N.; Lian, R.; Zhang, G.-L.; Hou, H.-M. Comparison of the Inhibitory Potential of Benzyl Isothiocyanate and Phenethyl Isothiocyanate on Shiga Toxin-Producing and Enterotoxigenic Escherichia coli. LWT 2020, 118, 108806. [Google Scholar] [CrossRef]
  262. Kim, M.G.; Lee, H.S. Growth-Inhibiting Activities of Phenethyl Isothiocyanate and Its Derivatives against Intestinal Bacteria. J. Food Sci. 2009, 74, M467–M471. [Google Scholar] [CrossRef] [PubMed]
  263. Chen, H.; Wang, C.; Ye, J.; Zhou, H.; Chen, X. Antimicrobial Activities of Phenethyl Isothiocyanate Isolated from Horseradish. Nat. Prod. Res. 2012, 26, 1016–1021. [Google Scholar] [CrossRef] [PubMed]
  264. Nowicki, D.; Maciąg-Dorszyńska, M.; Bogucka, K.; Szalewska-Pałasz, A.; Herman-Antosiewicz, A. Various Modes of Action of Dietary Phytochemicals, Sulforaphane and Phenethyl Isothiocyanate, on Pathogenic Bacteria. Sci. Rep. 2019, 9, 13677. [Google Scholar] [CrossRef] [PubMed]
  265. Van Eylen, D.; Oey, I.; Hendrickx, M.; Van Loey, A. Effects of Pressure/Temperature Treatments on Stability and Activity of Endogenous Broccoli (Brassica oleracea L. Cv italica) Myrosinase and on Cell Permeability. J. Food Eng. 2008, 89, 178–186. [Google Scholar] [CrossRef]
  266. Fusari, C.M.; Ramirez, D.A.; Camargo, A.B. Simplified analytical methodology for glucosinolate hydrolysis products: A mniaturized extraction technique and multivariate optimization. Anal. Methods 2019, 11, 309–316. [Google Scholar] [CrossRef]
  267. Øverby, A.; Stokland, R.A.; Åsberg, S.E.; Sporsheim, B.; Bones, A.M. Allyl isothiocyanate depletes glutathione and upregulates expression of glutathione S-transferases in Arabidopsis thaliana. Front. Plant Sci. 2015, 6, 277. [Google Scholar] [CrossRef] [Green Version]
  268. Fahey, J.W. Method of Extraction of Isothiocyanates into Oil from Glucosinolsate-Containing Plants. US20060127996A1, 15 June 2006. WO 2006/065736 A3. [Google Scholar]
  269. Rodrigues, L.; Silva, I.; Poejo, J.; Serra, A.T.; Matias, A.A.; Simplício, A.L.; Bronze, M.R.; Duarte, C.M.M. Recovery of Antioxidant and Antiproliferative Compounds from Watercress Using Pressurized Fluid Extraction. RSC Adv. 2016, 6, 30905–30918. [Google Scholar] [CrossRef]
  270. Ji, Y.; Morris, M.E. Determination of Phenethyl Isothiocyanate in Human Plasma and Urine by Ammonia Derivatization and Liquid Chromatography–Tandem Mass Spectrometry. Anal. Biochem. 2003, 323, 39–47. [Google Scholar] [CrossRef] [PubMed]
  271. Chatzilazarou, A. Application of Cloud Point Extraction Using Surfactants in the Isolation of Physical Antioxidants (Phenols) from Olive Mill Wastewater. Fresenius Environ. Bull. 2006, 15, 1122–1125. [Google Scholar]
  272. Vieira, F.A.; Guilherme, R.J.R.; Neves, M.C.; Abreu, H.; Rodrigues, E.R.O.; Maraschin, M.; Coutinho, J.A.P.; Ventura, S.P.M. Single-Step Extraction of Carotenoids from Brown Macroalgae Using Non-Ionic Surfactants. Sep. Purif. Technol. 2017, 172, 268–276. [Google Scholar] [CrossRef]
  273. Tani, H.; Kamidate, T.; Watanabe, H. Aqueous Micellar Two-Phase Systems for Protein Separation. Anal. Sci. 1998, 14, 875–888. [Google Scholar] [CrossRef] [Green Version]
  274. Molina-Bolívar, J.A.; Aguiar, J.; Ruiz, C.C. Light Scattering and Fluorescence Probe Studies on Micellar Properties of Triton X-100 in KCl Solutions. Mol. Phys. 2001, 99, 1729–1741. [Google Scholar] [CrossRef]
  275. Raja, S.; Murty, V.R.; Thivaharan, V.; Rajasekar, V.; Ramesh, V. Aqueous Two-Phase Systems for the Recovery of Biomolecules—A Review. Sci. Technol. 2012, 1, 7–16. [Google Scholar] [CrossRef]
  276. Cordisco, E.; Haidar, C.N.; Goñi, R.; Nerli, B.B.; Malpiedi, L.P. Physicochemical Characterization of Aqueous Micellar Systems Formed by Environmentally Friendly Salts. Fluid Phase Equilib. 2015, 393, 111–116. [Google Scholar] [CrossRef]
  277. Xian, M.; Wawrzyniak, P.; Rückert, B.; Duan, S.; Meng, Y.; Sokolowska, M.; Globinska, A.; Zhang, L.; Akdis, M.; Akdis, C.A. Anionic surfactants and commercial detergents decrease tight junction barrier integrity in human keratinocytes. J. Allergy Clin. Immunol. 2016, 138, 890–893. [Google Scholar] [CrossRef] [Green Version]
  278. Alibrahim, M. Cloud Point Extraction of Polycyclic Aromatic Hydrocarbons in Aqueous Solution with Nonionic Surfactants. Tenside Surfactants Deterg. 2014, 51, 333–338. [Google Scholar] [CrossRef]
  279. Yamaguchi, T. Mutagenicity of Isothiocyanates, Isocyanates and Thioureas on Salmonella typhimurium. Agric. Biol. Chem. 2014, 44, 3017–3018. [Google Scholar] [CrossRef]
  280. Kaschula, C.H.; Hunter, R. Synthesis and Structure–Activity Relations in Allylsulfide and Isothiocyanate Compounds From Garlic and Broccoli Against In Vitro Cancer Cell Growth. In Studies in Natural Products Chemistry; Rahman, A., Ed.; Elsevier: Amsterdam, The Netherlands, 2016; Volume 50, pp. 1–43. [Google Scholar] [CrossRef]
  281. Schlesinger, D.H. PROTEINS | Traditional Methods of Sequence Determination. In Encyclopedia of Analytical Sciences; Worsfold, P., Poole, C., Townshend, A., Miró, M., Eds.; Academic Press: Oxford, UK, 2005; pp. 420–424. [Google Scholar] [CrossRef]
  282. Oe, T.; Maekawa, M.; Satoh, R.; Lee, S.H.; Goto, T. Combining [C-13]-Phenylisothiocyanate and the Edman Degradation Reaction: A Possible Breakthrough for Absolute Quantitative Proteomics Together with Protein Identification. Rapid Commun. Mass Spectrom. 2010, 24, 173–179. [Google Scholar] [CrossRef] [PubMed]
  283. Hanna, P.E.; Anders, M.W. The Mercapturic Acid Pathway. Crit. Rev. Toxicol. 2019, 49, 819–929. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  284. Kuhn, C.; Kupke, F.; Baldermann, S.; Klopsch, R.; Lamy, E.; Hornemann, S.; Pfeiffer, F.H.A.; Schreiner, M.; Hanschen, S.F.; Rohn, S. Diverse Excretion Pathways of Benzyl Glucosinolate in Humans after Consumption of Nasturtium (Tropaeolum majus L.) – A Pilot Study. Mol. Nutr. Food. Res 2018, 62, 1–37. [Google Scholar] [CrossRef] [PubMed]
  285. Kim, Y.-J.; Lee, D.-H.; Ahn, J.; Chung, W.-J.; Jang, Y.J.; Seong, K.-S.; Moon, J.-H.; Ha, T.Y.; Jung, C.H. Pharmacokinetics, Tissue Distribution, and Anti-Lipogenic/Adipogenic Effects of Allyl-Isothiocyanate Metabolites. PLoS ONE 2015, 10, e0132151. [Google Scholar] [CrossRef] [PubMed]
  286. Charron, C.S.; Vinyard, B.T.; Ross, S.A.; Seifried, H.E.; Jeffery, E.H.; Novotny, J.A. Absorption and Metabolism of Isothiocyanates Formed from Broccoli Glucosinolates: Effects of BMI and Daily Consumption in a Randomised Clinical Trial. Br. J. Nutr. 2018, 120, 1370–1379. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  287. Dyba, M.; Wang, A.; Noone, A.-M.; Goerlitz, D.; Shields, P.; Zheng, Y.-L.; Rivlin, R.; Chung, F.-L. Metabolism of Isothiocyanates in Individuals with Positive and Null GSTT1 and M1 Genotypes after Drinking Watercress Juice. Clin. Nutr. 2010, 29, 813–818. [Google Scholar] [CrossRef] [Green Version]
  288. Fusari, C.M.; Locatelli, D.A.; Altamirano, J.C.; Camargo, A.B. UAE-HPLC-UV: New Contribution for Fast Determination of Total Isothiocyanates in Brassicaceae Vegetables. J. Chem. 2015, 2015, 294601. [Google Scholar] [CrossRef] [Green Version]
  289. Zhang, Y.; Cho, C.G.; Posner, G.H.; Talalay, P. Spectroscopic Quantitation of Organic Isothiocyanates by Cyclocondensation with Vicinal Dithiols. Anal. Biochem. 1992, 205, 100–107. [Google Scholar] [CrossRef]
  290. Zhang, Y. The 1,2-Benzenedithiole-Based Cyclocondensation Assay: A Valuable Tool for the Measurement of Chemopreventive Isothiocyanates. Crit. Rev. Food Sci. Nutr. 2012, 52, 525–532. [Google Scholar] [CrossRef] [Green Version]
  291. Mukerjee, A.K.; Ashare, R. Isothiocyanates in the Chemistry of Heterocycles. Chem. Rev. 1991, 91, 1–24. [Google Scholar] [CrossRef]
  292. Cho, C.-G.; Posner, G.H. Alkyl and Aryl Isothiocyanates as Masked Primary Amines. Tetrahedron Lett. 1992, 33, 3599–3602. [Google Scholar] [CrossRef]
  293. Zhang, Y.; Talalay, P. Mechanism of Differential Potencies of Isothiocyanates as Inducers of Anticarcinogenic Phase 2 Enzymes. Cancer Res. 1998, 58, 4632–4639. [Google Scholar] [PubMed]
  294. Masutomi, N.; Toyoda, K.; Shibutani, M.; Niho, N.; Uneyama, C.; Takahashi, N.; Masao, H. Toxic Effects of Benzyl and Allyl Isothiocyanates and Benzyl-Isoform Specific Metabolites in the Urinary Bladder After a Single Intravesical Application to Rats. Toxicol. Pathol. 2001, 29, 617–622. [Google Scholar] [CrossRef]
  295. Shapiro, T.A.; Fahey, J.W.; Wade, K.L.; Stephenson, K.K.; Talalay, P. Chemoprotective glucosinolates and isothiocyanates of broccoli sprouts: Metabolism and excretion in humans. Cancer Epidemiol Biomark. Prev. 2001, 10, 501–508. [Google Scholar]
  296. Kassahun, K.; Davis, M.; Hu, P.; Martin, B.; Baillie, T. Biotransformation of the Naturally Occurring Isothiocyanate Sulforaphane in the Rat:  Identification of Phase I Metabolites and Glutathione Conjugates. Chem. Res. Toxicol. 1997, 10, 1228–1233. [Google Scholar] [CrossRef] [PubMed]
  297. Zhang, Y.; Wade, K.L.; Prestera, T.; Talalay, P. Quantitative Determination of Isothiocyanates, Dithiocarbamates, Carbon Disulfide, and Related Thiocarbonyl Compounds by Cyclocondensation with 1,2-Benzenedithiol. Anal. Biochem. 1996, 239, 160–167. [Google Scholar] [CrossRef] [PubMed]
  298. Ye, L.; Dinkova-Kostova, A.T.; Wade, K.L.; Zhang, Y.; Shapiro, T.A.; Talalay, P. Quantitative Determination of Dithiocarbamates in Human Plasma, Serum, Erythrocytes and Urine: Pharmacokinetics of Broccoli Sprout Isothiocyanates in Humans. Clin. Chim. Acta. 2002, 316, 43–53. [Google Scholar] [CrossRef]
  299. Moy, K.A.; Yuan, J.-M.; Chung, F.-L.; Wang, X.-L.; Van Den Berg, D.; Wang, R.; Gao, Y.-T.; Yu, M.C. Isothiocyanates, Glutathione S-Transferase M1 and T1 Polymorphisms and Gastric Cancer Risk: A Prospective Study of Men in Shanghai, China. Int. J. Cancer 2009, 125, 2652–2659. [Google Scholar] [CrossRef] [Green Version]
  300. London, S.J.; Yuan, J.M.; Chung, F.L.; Gao, Y.T.; Coetzee, G.A.; Ross, R.K.; Yu, M.C. Isothiocyanates, Glutathione S-Transferase M1 and T1 Polymorphisms, and Lung-Cancer Risk: A Prospective Study of Men in Shanghai, China. Lancet 2000, 356, 724–729. [Google Scholar] [CrossRef]
  301. Ina, K.; Nobukuni, M.; Sano, A.; Kishima, I. Stability of Allyl Isothiocyanate. Nippon Shokuhin Kogyo Gakkaishi 1981, 28, 627–631. [Google Scholar] [CrossRef]
  302. Luo, B.; Wang, J.; Li, X.; Lu, W.; Yang, J.; Hu, Y.; Huang, P.; Wen, S. ; New Mild and Simple Approach to Isothiocyanates: A Class of Potent Anticancer Agents. Molecules 2017, 22, 773. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  303. Agerbirk, N.; De Nicola, G.R.; Olsen, C.E.; Müller, C.; Iori, R. Derivatization of Isothiocyanates and Their Reactive Adducts for Chromatographic Analysis. Phytochemistry 2015, 118, 109–115. [Google Scholar] [CrossRef] [PubMed]
  304. Song, L.; Morrison, J.J.; Botting, N.P.; Thornalley, P.J. Analysis of gucosinolates, isothiocyanates, and amine degradation products in vegetable extracts and blood plasma by LC-MS/MS. Anal. Biochem. 2005, 347, 234–243. [Google Scholar] [CrossRef] [PubMed]
  305. Wilson, E.A.; Ennahar, S.; Zhao, M.; Bergaentzle, M.; Marchioni, E.; Bindler, F. Simultaneous Determination of Various Isothiocyanates by RP-LC Following Precolumn Derivatization with Mercaptoethanol. Chromatographia 2011, 73, 137–142. [Google Scholar] [CrossRef] [Green Version]
  306. Revelou, P.K.; Kokotou, M.G.; Pappas, C.S.; Constantinou-Kokotou, V. Direct Determination of Total Isothiocyanate Content in Broccoli Using Attenuated Total Reflectance Infrared Fourier Transform Spectroscopy. J. Food Compos. Anal. 2017, 61, 47–51. [Google Scholar] [CrossRef]
  307. Socrates, G. Infrared and Raman Characteristic Group Frequencies. J. Raman Spectrosc. 2004, 35, 905. [Google Scholar] [CrossRef]
  308. Pocasap, P.; Weerapreeyakul, N. ; Sulforaphene and Sulforaphane in commonly consumed cruciferous plants contributed to antiproliferation in HCT116 colon cancer cells. Asian Pac. J. Trop. Biomed. 2016, 6, 119–124. [Google Scholar] [CrossRef] [Green Version]
  309. Jin, Y.; Wang, M.; Rosen, R.T.; Ho, C.T. Thermal Degradation of Sulforaphane in Aqueous Solution. J. Agric. Food Chem. 1999, 47, 3121–3123. [Google Scholar] [CrossRef]
  310. Choi, M.M.F.; Shuang, S.; Lai, H.Y.; Cheng, S.C.; Cheng, R.C.W.; Cheung, B.K.B.; Lee, A.W.M. Gas Chromatography-Mass Spectrometric Determination of Total Isothiocyanates in Chinese Medicinal Herbs. Anal. Chim. Acta. 2004, 516, 155–163. [Google Scholar] [CrossRef]
  311. Hong, E.; Kim, G.H. GC-MS Analysis of the Extracts from Korean Cabbage (Brassica campestris L. ssp. pekinensis) and Its Seed. Prev. Nutr. Food Sci. 2013, 18, 218–221. [Google Scholar] [CrossRef] [Green Version]
  312. Abdel-Kader, M.S.; Khamis, E.H.; Foudah, A.I.; Alqarni, M.H. GC Quantitative Analysis of Benzyl Isothiocyanate in Salvadora persica Roots Extract and Dental Care Herbal Products. Saudi Pharm. J. 2018, 26, 462–466. [Google Scholar] [CrossRef] [PubMed]
  313. Sofrata, A.; Santangelo, E.M.; Azeem, M.; Borg-Karlson, A.-K.; Gustafsson, A.; Pütsep, K. Benzyl Isothiocyanate, a Major Component from the Roots of Salvadora Persica Is Highly Active against Gram-Negative Bacteria. PLoS ONE 2011, 6, e23045. [Google Scholar] [CrossRef] [PubMed]
  314. Karanikolopoulou, S.; Revelou, P.-K.; Xagoraris, M.; Kokotou, M.G.; Constantinou-Kokotou, V. Current Methods for the Extraction and Analysis of Isothiocyanates and Indoles in Cruciferous Vegetables. Analytica 2021, 2, 93–120. [Google Scholar] [CrossRef]
  315. Zheng, L.; Zheng, F. Development and Validation of an LC-APCI-MS/MS Method for the Determination of Phenethyl Isothiocyanate in Human Plasma. Biomed. Chromatogr. 2015, 29, 619–625. [Google Scholar] [CrossRef] [PubMed]
  316. Yu, X.; Ma, F.; Zhang, L.; Li, P. Extraction and Quantification of Sulforaphane and Indole-3-Carbinol from Rapeseed Tissues Using QuEChERS Coupled with UHPLC-MS/MS. Molecules 2020, 25, 2149. [Google Scholar] [CrossRef]
  317. Alvarez-Jubete, L.; Smyth, T.J.; Valverde, J.; Rai, D.K.; Barry-Ryan, C. Simultaneous Determination of Sulphoraphane and Sulphoraphane Nitrile in Brassica Vegetables Using Ultra-Performance Liquid Chromatography with Tandem Mass Spectrometry. Phytochem. Anal. 2014, 25, 141–146. [Google Scholar] [CrossRef]
  318. Liang, H.; Li, C.; Yuan, Q.; Vriesekoop, F. Application of High-Speed Countercurrent Chromatography for the Isolation of Sulforaphane from Broccoli Seed Meal. J. Agric. Food Chem. 2008, 56, 7746–7749. [Google Scholar] [CrossRef]
Figure 1. Schematic representation of the segregation of ITCs based on the type of the R group (sulfoxides, alkenyls, and aromatics).
Figure 1. Schematic representation of the segregation of ITCs based on the type of the R group (sulfoxides, alkenyls, and aromatics).
Antioxidants 11 00642 g001
Figure 2. Hydrolysis of GSLs via enzymatic catalysis by myrosinase is activated upon tissue plant disruption. The formed aglucone can be rearranged into oxazolidin-2-thiones, nitriles, epithionitres, thiocyanates, and ITCs.
Figure 2. Hydrolysis of GSLs via enzymatic catalysis by myrosinase is activated upon tissue plant disruption. The formed aglucone can be rearranged into oxazolidin-2-thiones, nitriles, epithionitres, thiocyanates, and ITCs.
Antioxidants 11 00642 g002
Figure 3. The involvement of ITCs in various cell signalling cascades associated with cell cycle arrest, epigenetic modulation, generation of oxidative stress, angiogenesis, and apoptosis. Arrowheads pointing upwards or downwards indicate an upregulation or downregulation of the expression of various indicated protein markers and ROS, respectively.
Figure 3. The involvement of ITCs in various cell signalling cascades associated with cell cycle arrest, epigenetic modulation, generation of oxidative stress, angiogenesis, and apoptosis. Arrowheads pointing upwards or downwards indicate an upregulation or downregulation of the expression of various indicated protein markers and ROS, respectively.
Antioxidants 11 00642 g003
Figure 4. The 4 pillars: (1) sample processing and storage, (2) hydrolysis of GSLs, (3) extraction of ITCs, and (4) quantification for the determination of ITCs from naturally—derived sources.
Figure 4. The 4 pillars: (1) sample processing and storage, (2) hydrolysis of GSLs, (3) extraction of ITCs, and (4) quantification for the determination of ITCs from naturally—derived sources.
Antioxidants 11 00642 g004
Figure 5. Graphical presentation of the proposed methodological pipelines for the determination of ITCs from various natural sources, taking into consideration the efficacy of each approach based on recovery yield of isolated ITCs.
Figure 5. Graphical presentation of the proposed methodological pipelines for the determination of ITCs from various natural sources, taking into consideration the efficacy of each approach based on recovery yield of isolated ITCs.
Antioxidants 11 00642 g005
Figure 6. The cyclo-condensation assay is used for the quantification of ITCs. Briefly, it involves the nucleophilic attack of both vicinal thiols of 1,2-benzenedithiol into the electrophilic carbon atom of ITC, leading to the formation of a stable 1,3-benzodithiole-2-thione product which absorbs at 365 nm.
Figure 6. The cyclo-condensation assay is used for the quantification of ITCs. Briefly, it involves the nucleophilic attack of both vicinal thiols of 1,2-benzenedithiol into the electrophilic carbon atom of ITC, leading to the formation of a stable 1,3-benzodithiole-2-thione product which absorbs at 365 nm.
Antioxidants 11 00642 g006
Table 1. Natural sources of ITCs including genus species.
Table 1. Natural sources of ITCs including genus species.
ITCs SourceGenus Species (Sub Species)Isothiocyanates (ITCs)Ref
BroccoliBrassica oleracea var. italicasulforaphane (SFN)[28,29]
Curly KaleBrassica oleracea var. laciniata (L.)SFN[30]
CauliflowerBrassica oleracea var. caulifloraSFNphenethyl isothiocyanate (PEITC)[30,31]
CabbageBrassica rapa var. pekinensisSFN[32]
Brussel sproutBrassica oleracea var. gemmiferaSFN[33]
HorseradishArmoracia lapathifolia (L.) Gilibiberin (IBN) allyl isothiocyanate (AITC)[34,35]
RadishRaphanus sativus (L.)IBN[34]
WatercressNasturtium officinalePEITC[36]
Wasabi rootsEutrema japonicum (L.) Koidz.AITC[37]
Mustard seedsSinaptis albabenzyl isothiocyanate (BITC)[38]
Papaya seedsCarica papaya (L.)BITC[39]
Table 2. Optimum pH conditions for complete conversion of various GSLs into their respective ITCs in various plants.
Table 2. Optimum pH conditions for complete conversion of various GSLs into their respective ITCs in various plants.
Plant SpeciespHGlucosinolate
(GSL)
Isothiocyanate
(ITC)
Extraction Yield (%)Refs
Eruca sativa var. sativa
Erysimum allionii
10.0glucoerucinerucin98.7
96.6
[61,85,86,87,88,89,90,91,92,93,94,95,96,97,98,99,100,101,102]
Brassica hirta (L.) Moench7.0, 10.04-hydroxybenzyl 4-hydroxybenzyl29.6[103]
Lesquerella fendleri7.0, 10.0glucoiberinIBN90.0[85]
Brassica vegetables
Mattiola longipetala
7.0, 10.0glucoraphinSFN53.0
15.0
[61,91,92,93,94,95,96,97,98,99,100,101,102,103,104]
Lobularia maritima (L.) Desv
Capsella bursa-pastoris (L.) Medik
7.0, 10.03-butenyl3-butenyl~100.0[85]
Erysimum cheiri (L.) Crantz7.0, 10.0glucocheirolincheirolin98.3[85]
Lobularia maritima (L.) Desv7.0, 10.0lesquerellinlesquerellin96.3[85]
Nasturtium officinale7.0gluconasturtiinPEITC89.095.0[87,105,106]
Carica papaya (L.)
Tropaeolum majus (L.) tropaeolaceae
Lepidium sativum (L.)
6.5–7.0glucotropaeolinBITCn.d.[107]
Brassica rapa (L.) oleifera8.03-butenyl glucosinolate3-butenyl ITC40.0[56]
Arabidopsis thaliana (L.) Heynh6.52-propenyl glucosinolate2-propenyl ITC32.0[56]
Brassica oleracra var. italica5.54-(methylsulphinyl)
butyl glucosinolate
4-(methylsulphinyl)
butyl isothiocyanate
82.0[56]
Armoracia rusticana7.0sinigrinAITC61.0[108]
Table 3. Optimum temperature(s) for complete degradation of GSLs into their respective ITCs in various plants.
Table 3. Optimum temperature(s) for complete degradation of GSLs into their respective ITCs in various plants.
Plant SpeciesTemperatureGlucosinolate
(GSL)
Isothiocyanate(ITC)Extraction Yield (%)Refs
Brassica vegetables37 °CglucoraphaneSFN53.0[91]
Nasturtium officinale25 °CgluconasturtiinPEITC89.0[87]
[106]
Brassica rapa var. oleifera25 °C3-butenyl
glucosinolate
3-butenyl ITC40.0[55]
Arabidopsis thaliana (L.) Heynh37 °C2-propenyl
glucosinolate
2-propenyl ITC32.0[55]
3-butenyl
glucosinolate
2- buteny
l glucosinolate
3-Butenyl ITC
2-Butenyl ITC
n.d.
n.d.
[62]
Brassica oleracra var. italica37 °C4-(methylsulphinyl)butyl glucosinolate4-(methylsulphinyl)butyl isothiocyanate82.0[56]
Carica papaya (L.)16–18 °CglucotropaeolinBEITC85.6[119]
Salix alba (L.) maire<50 °CsinigrinAITCn.d.[112]
Brassica nigra (L.) nigra<60 °CsinigrinAITCn.d.[112]
Brassica juncea (L.) Czern.<50 °CsinigrinAITCn.d.[112]
Table 4. Optimum conditions for the extraction of ITCs from naturally-derived sources.
Table 4. Optimum conditions for the extraction of ITCs from naturally-derived sources.
Plant SourceIsothiocyanates
(ITCs)
Extraction MethodologyExtraction
Yield (%)
Refs
BroccoliSFNSolid phase extraction94[139]
MatthiolalongipetalaLiquid-liquid extraction
(with n-hexane)
31.9[85]
Lesquerella fendleri (L.)IBNLiquid-liquid extraction
(with dichloromethane)
48.6[85]
Physaria fendleri56.7[210]
Wasabia Japonica (L.) matsumAITCSuprecritical fluid (CO2) extraction79.1[35]
Armoracia rusticanaHydrodistillation61[35]
Liquid-liquid extraction
(with diethyl ether)
96.5[35]
Green papaya BITCHydrodistillation80[122]
Salvadora persica (L.)Liquid-liquid extraction
(with dichloromethane)
75[236]
Liquid-liquid extraction (with chloroform)40[237]
Nasturdium officinalePEITCExtraction with non-anionic surfactants94[98]
Liquid-liquid extraction
(with n-hexane)
98.7[106]
Suprecritical fluid (CO2) extraction87.4, ND[206,271]
Lobularia maritima6-methylthiohexyl isothiocyanateLiquid-liquid extraction
(with n-hexane)
96.3[85]
MatthiolalongipetalaerysolinLiquid-liquid extraction
(with n-hexane)
38.7[85]
cherolin18.7
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Kyriakou, S.; Trafalis, D.T.; Deligiorgi, M.V.; Franco, R.; Pappa, A.; Panayiotidis, M.I. Assessment of Methodological Pipelines for the Determination of Isothiocyanates Derived from Natural Sources. Antioxidants 2022, 11, 642. https://doi.org/10.3390/antiox11040642

AMA Style

Kyriakou S, Trafalis DT, Deligiorgi MV, Franco R, Pappa A, Panayiotidis MI. Assessment of Methodological Pipelines for the Determination of Isothiocyanates Derived from Natural Sources. Antioxidants. 2022; 11(4):642. https://doi.org/10.3390/antiox11040642

Chicago/Turabian Style

Kyriakou, Sotiris, Dimitrios T. Trafalis, Maria V. Deligiorgi, Rodrigo Franco, Aglaia Pappa, and Mihalis I. Panayiotidis. 2022. "Assessment of Methodological Pipelines for the Determination of Isothiocyanates Derived from Natural Sources" Antioxidants 11, no. 4: 642. https://doi.org/10.3390/antiox11040642

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

Article Metrics

Back to TopTop