Next Article in Journal
Effect of Near-Freezing Storage Combined with High-Voltage Electric Fields on the Freshness of Large Yellow Croaker
Previous Article in Journal
Effects of Cultivar Factors on Fermentation Characteristics and Volatile Organic Components of Strawberry Wine
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Chloropropanols and Their Esters in Food: An Updated Review

by
Gizem Ozluk
1,
Miguel Ángel González-Curbelo
2 and
Bulent Kabak
1,3,*
1
Department of Food Engineering, Faculty of Engineering, Hitit University, Corum 19030, Turkey
2
Departamento de Ciencias Básicas, Facultad de Ingeniería, Universidad EAN, Calle 79 no 11–45, Bogotá 110221, Colombia
3
Biotechnology Laboratory, Machinery and Manufacturing Technology Application and Research Center, Hitit University, Corum 19030, Turkey
*
Author to whom correspondence should be addressed.
Foods 2024, 13(18), 2876; https://doi.org/10.3390/foods13182876
Submission received: 15 August 2024 / Revised: 29 August 2024 / Accepted: 9 September 2024 / Published: 11 September 2024
(This article belongs to the Section Food Toxicology)

Abstract

:
Chloropropanols, their fatty acid esters, and glycidol and its fatty acid esters (GEs) are process contaminants in foods that pose potential health risks. These contaminants typically arise during the deodorization process of vegetable oils, particularly in high concentrations within oils like palm oil and products derived from them, such as margarine, baked goods, pastries, and infant formula. Chloropropanol esters and GE can hydrolyze under the influence of lipases, forming chloropropanols. Elevated temperatures during food production can lead to the release of free 3-chloro-1,2-propanediol (3-MCPD) or free 2-chloro-1,3-propanediol (2-MCPD) in products containing both fat and salt. The exposure to these contaminants, especially for infants and young children, raises concerns about potential health hazards. While extensive research has focused on 3-MCPD, 2-MCPD, and GE, knowledge regarding other chloropropanols such as 1,3-dichloro-2-propanol (1,3-DCP), 2,3-dichloro-1-propanol (2,3-DCP), and their fatty acid esters remains limited. This review aims to provide a comprehensive overview encompassing formation mechanisms, analysis methods, toxicological implications, occurrence patterns, exposure levels, mitigation strategies, and legislative considerations concerning these contaminants.

1. Introduction

Recently, there has been an increasing interest in the potential carcinogenic effects of contaminants that arise from heat treatment processes. Chloropropanols are a class of chemicals in which the hydroxyl group(s) of glycerol is replaced by chlorine. Based on the replacement position and the number of chlorines, it can be separated into 2-chloro-1,3-propanediol (2-MCPD), 3-chloro-1,2-propanediol (3-MCPD), 1,3-dichloro-2-propanol (1,3-DCP), and 2,3-dichloro-1-propanol (2,3-DCP) [1].
Since the inception of food processing, chloropropanol derivatives have been found in foods, but the free form of 3-MCPD was first identified in acid-hydrolyzed vegetable proteins (acid-HVPs) in 1978 [2]. The esterified derivative of 3-MCPD, conjugated with fatty acids, was identified in 1980 [3]. In 2006, the presence of 3-MCPD esters in refined vegetable oils was confirmed, and the fatty acid ester configurations of MCPDs were identified as process contaminants [4]. In December 2007, 3-MCPD esters were identified in high amounts in infant formulas [5]. Their presence was also observed in human breast milk [6]. Later, high quantities of 3-MCPD have been found in a wide range of heat-treated foods, including meat products, milk and dairy products, cereal-derived products, coffee, fried cheese, biscuits, and smoked foods [7]. While the formation of bound 2- and 3-MCPD, as well as glycidyl fatty acid esters (GEs), typically occurs during the deodorization process in the refining of oils and fats, free MCPD is usually produced during the process of roasting and smoking [8].
The presence of 1,3-DCP, 2,3 DCP, and 2-MCPD has been observed in a wide range of foodstuffs, although these compounds have received relatively less attention in research. Notably, 1,3-DCP has also been detected in drinking water [9], highlighting the importance of considering these substances. The comprehensive investigation of all three compounds was initiated in the early 2000s, with a primary focus on soy sauce [10,11,12].
Besides vegetable oils and foodstuffs, chloropropanols have also been recorded in several food contact paper products globally, causing significant concern. It has been reported that the primary origin of chloropropanols in food contact paper products is the residual substance from polyamide epichlorohydrin resin, which is a wet-strength agent frequently employed in pulp manufacturing. During this process, epichlorohydrin can undergo hydrolysis, resulting in the formation of 3-MCPD and 1,3-DCP [8]. Thus, 3-MCPD and/or 1,3-DCP may also be found in foodstuffs through migration from contact materials such as sausage casings, tea bags, and coffee filters [13].
GEs are considered food processing contaminants that are generated during the physical refining process of vegetable oils and fats at temperatures exceeding 200 °C, specifically during the deodorization step. Glycidol is produced from the hydrolysis of GE ingested through food consumption during human digestion [1].
Chloropropanols and their esters are of significant toxicological concern due to their potential carcinogenicity. Studies have shown that 3-MCPD, 2-MCPD, and DCPs (such as 1,3-DCP) can induce kidney and testicular toxicity in animal models [1], with 3-MCPD classified as a possible human carcinogen by the International Agency for Research on Cancer (IARC) [14]. Additionally, glycidol, a hydrolysis product of glycidyl esters, is recognized for its genotoxic and carcinogenic properties, posing a severe risk to human health [1]. The carcinogenic properties of glycidol were initially discovered through rodent studies conducted by the National Toxicology Program (NTP) in 1990 [15]. Chronic exposure to these compounds, especially through processed foods, highlights the importance of assessing and minimizing their presence in the food supply chain.
This review provides a comprehensive overview of carcinogenic chloropropanol derivatives, specifically MCPDs, DCPs, glycidol, and GE. This document provides an overview of the scientific research related to MCPD and GE formation mechanisms, occurrence, exposure, toxicological effects, analytical methods, mitigation strategies, and existing legislative frameworks.

2. Definitions, Properties, and Chemical Structures

Chloropropanols are chlorohydrins, which are propanols that include a chloride functional group. Chloropropanols are classified as monochloropropanol (MCPD) or dichloropropanols (DCP) based on the amount of chlorine atoms connected to the structure. Monochloropropanols are accessible in both monoesters and diesters, whereas dichloropropanols are only found in monoesters. They are designated after the location of the chlorine in the propanol molecule to which it is connected, such as 3-MCPD and 2-MCPD [1]. The chemical structures of chloropropanol isomers are shown in Figure 1.
These compounds are typically encountered in their esterified form with fatty acids. However, it is worth noting that they can also be found in small quantities in their free form. The chloropropanol esters are characterized by the presence of hydroxyl groups that have been esterified with fatty acids on the chloropropanol skeleton. MCPD-monoesters or MCPD-diesters are formed through the esterification of hydroxyl groups with fatty acids, either partially or fully. The compounds 3-MCPD, 2-MCPD, and glycidol are chloropropanols that are obtained from chlorinated derivatives of glycerol. The positioning of the chlorine atom differs between 3-MCPD and 2-MCPD. In 3-MCPD, the chlorine atom is located at position 3, whereas in 2-MCPD, it is situated at position 2 [9,16,17].
3-MCPD consists of three carbon (C) atoms, two functional alcohol (-OH) groups, and one chlorine (Cl) atom. The compound has a molecular formula of C3H7ClO2 and a molecular weight of 110.54 g/mol. The compound 3-MCPD is soluble in water, alcohol, diethyl ether, and acetone. It is colorless but may gradually turn straw yellow [18]. 3-MCPD has been classified as a rodenticide by the United States Environmental Protection Agency (US EPA) and is referred to as α-chlorohydrin [18]. 2-MCPD is an isomer of 3-MCPD; hence, the two have identical physical and chemical properties. 2-MCPD has a density of 1.3 g/cm3, and it, along with 3-MCPD and glycidol, can be esterified with fatty acids [1,19].
Contrary to free MCPDs, the fatty acid esters of MCPDs are insoluble in water and polar solvents. The fatty acid esters of MCPDs can exist in nearly 100 different isomers due to the ability of the ester bonds to form at one or two of the available hydroxyl positions with any of the fatty acids. The number of GEs is extremely low due to the single hydroxyl group and the absence of positional isomers [1].
The compound 1,3-DCP consists of two chloride ions and has the chemical formula C3H6Cl2O. 2,3-DCP is an isomer of 1,3-DCP and includes two Cl ions. 1,3-DCP and 2,3-DCP are liquids at room temperature, with molecular weights of 128.98 g/mol and 163 g/mol, respectively. 1,3-DCP, which functions as a protic solvent and crosslinking reagent, is utilized as a solvent for hard resins and nitrocellulose. 2,3-DCP is utilized in the manufacturing process of epichlorohydrin. The utilization of 1,3-DCP and 2,3-DCP in chemical processes has raised concerns due to the potential for toxic effects resulting from the discharge of their waste into soil and water [20,21].
Glycidol, also known as 2,3-epoxy-1-propanol, is a colorless liquid that is soluble in water and most polar solvents. It has a molecular formula of C3H6O2 (Figure 2) and a molecular weight of 74.08 g/mol. It is an odorless liquid cyclic ether. Glycidol has a single hydroxyl group and forms only monoesters [1]. GE is another food process contaminant derived from refined oil. It contains an epoxide group esterified with fatty acids. Even though GE and MCPD esters are evaluated concurrently by researchers on account of their structural similarities, it is hypothesized that these compounds may exhibit distinct toxic effects because of the varying functional groups they possess [22,23,24].

3. Toxicological Effects

The toxicological impact of chloropropanols and their derivatives encompasses a range of mechanisms, including carcinogenicity, genotoxicity, and reproductive toxicity, observed across multiple species. The carcinogenic effects of chloropropanol and dichloropropanol were observed in rats, resulting in the development of various neoplasms. While chloropropanol’s genotoxicity was confirmed in vitro, dichloropropanol showed clear genotoxic effects in both in vivo and in vitro investigations. The specific method of action for the carcinogenicity of chloropropanol and dichloropropanol is unknown; however, evidence suggests probable hormone-mediated, cytotoxic, and genotoxic mechanisms [25]. The IARC has classified 1,3-DCP and 3-MCPD as “possibly carcinogenic to humans” (Group 2B). Additionally, glycidol has been classified as Group 2A, indicating that it is a probable carcinogen in humans [14,26].
It has been reported that the most important toxic effects of 3-MCPD are on the reproductive system [27,28,29]. 3-MCPD was previously considered a non-genotoxic carcinogen based on the observation that kidney tumors can arise from chronic nephropathy, and Leydig cell tumors can develop because of hormonal irregularities (known as endocrine disruptive effects) in the testicles [30,31]. It should be noted that 3-MCPD has been found to have toxic effects on the reproductive system and a potential carcinogenic effect on the kidneys, specifically renal tubular carcinoma [32]. It is effective in rats, rams, wild boars, guinea pigs, hamsters, rhesus monkeys, and human sperm but not in mice or rabbits [33]. Studies have demonstrated that 3-MCPD induces DNA strand breaks and causes genotoxicity in Chinese hamster ovary cells in vitro but does not cause DNA chain breaks in the liver, bone marrow, and lymphocyte cells in vivo [34].
The immunotoxic effects of 3-MCPD have also been recorded in female rats at a dose of 100 mg/kg body weight (b.w.)/day [35]. The various isomers of 3-MCPD exhibit distinct biological activities. The (S) isomer exhibits fertility inhibitory activity, while the (R) isomer has been found to have a detrimental effect on the kidneys [36].
A tolerable daily intake (TDI) of 0.8 mg/kg b.w./day for 3-MCPD and its fatty acid esters has been established, based on a BMDL10 value of 0.077 mg/kg b.w./day for the induction of renal tubular hyperplasia in rats, whereas no health-based guidance value could be established for 2-MCPD and its fatty acid esters [1].
The available information regarding the toxic effects of 2-MCPD is currently limited. The lethal dose (LD50) value for rats in the acute toxicity test was determined to be between 50–60 mg/kg [1]. It has been shown that 2-MCPD does not cause diuresis in Sprague-Dawley rats; however, 3-MCPD causes kidney failure and even renal failure [37]. No genotoxic potential has been observed for 2-MCPD [1]. The oral toxicological effects of chloropropanols and glycidol are summarized in Table 1.
Glycidol, like 3-MCPD, has been shown to cause infertility at low doses. This impact is assumed to be caused by the conversion of glycidol to 3-MCPD in the stomach. The rat’s lowest observed adverse effect level (LOAEL) for glycidol was determined to be 25 mg/kg b.w. per day. The administration of glycidol resulted in a reduction in both sperm count and motility in mice, with observations of testicular atrophy across all dosage levels. The LOAEL for mice was determined to be 75 mg/kg b.w. per day [1]. Terminal axon damage was observed when glycidol was administered to pregnant mice. It was found that the offspring of pregnant mice exposed to 1600 mg/kg glycidol had fewer interneurons. This shows that glycidole exposure during pregnancy may result in worse cognitive performance in kids later in life [42]. Upon exposure to a dosage of 200 mg/kg of glycidol, rats exhibited a gradual deterioration in gait abnormalities, as well as histological and immunohistochemical differences, and lesions appeared in the central and peripheral nerve systems. Glycidol has been observed to stimulate the growth of new neurons in the hippocampus and cause damage to axons during the later stages of development [43]. The compound has been reported to cause gene mutations and unscheduled DNA synthesis [44].
Glycidol has been shown to cause numerous genotoxic effects in mammalian cells, including genetic mutations, chromosomal abnormalities, sister chromatid exchange, and unprogrammed DNA synthesis. In addition to the reproductive system, investigations have revealed that glycidol can cause gastrointestinal, liver, and skin cancers when ingested orally. Considering the genotoxic and carcinogenic properties of glycidol, a margin of exposure (MOE) method was used. An MOE equal to or exceeding 25,000 was deemed to indicate a low level of health concern [1].
The LD50 values of glycidol for mice and rats were determined to be 431 mg/kg and 420 mg/kg, respectively [45,46]. Mihalache and Dall’Asta [47] calculated the potential cancer risk associated with glycidol, estimating an annual impact of 5.37/100,000 persons disability-adjusted life year (DALY). This suggests an elevated chance of developing cancer over one’s lifetime.
GE has been shown to have a lower genotoxicity potential than glycidol, and the MOE for individuals with average and highest exposure levels is specified as 17,800 and 10,900, respectively [48].
Based on the exposure data for 1,3-DCP and 2,3-DCP, The Committee on Muta-genicity of Chemicals in Food [49] concluded that the estimated 1,3-DCP intakes are of low concern to human health. Later, the Committee on Carcinogenicity of Chemicals in Food [50] reported that 1,3-DCP should be listed as a carcinogen. It has been observed that 1,3-DCP not only exhibits infertility effects but also induces the formation of tumors in the liver, kidney, ureter, and bladder. The oral LD50 values of 1,3-DCP and 2,3-DCP were 110 mg/kg and 90 mg/kg in rat, respectively [40]. These values are much lower than the specified value for glycidol, indicating that 1,3-DCP and 2,3-DCP have more acute toxicity than glycidol.

4. Formation

Chloropropanols are known to be formed during food processing. The formation of 3-MCPD esters has been reported to occur through primary reactions during food processing in the presence of chlorine. This reaction occurs when foods that have a high fat and salt content are subjected to heat treatment. The synthesis of MCPD esters is primarily influenced by chlorine ions, acylglycerols, temperature, and time. On the other hand, GEs are formed under acidic conditions and high temperatures, primarily by monoglycerides and diglycerides [51]. The formation mechanism of GE under thermal conditions is illustrated in Figure 3 [52]. GE is mostly formed from diacetylglycerol during the deodorization process in the physical refining of oils, specifically at temperatures above 200 °C [1]. The hydrolysis of esters into their free forms, 3-MCPD and 2-MCPD, is facilitated by the lipase enzyme secreted by the pancreas in the intestinal system [53].
There are three hypothesized mechanisms for the formation of 3-MCPD in foods: acid hydrolysis, heat treatment, and lipase activity releasing it from its esters. 3-MCPD is generated through the chemical reaction between cooking oils and acid. The raw material contains HCl, triacylglycerols, phospholipids, and glycerol, which serve as precursors for the formation of chloropropanols [54]. The process of hydrohydrin production through the reaction of hydrochloric acid (HCl) and glycerol was initially documented by Velisek et al. [2]. In a later study, Hamlet et al. [36] put forward a mechanism that elucidates the heat-induced formation of chloropropanols from triacylglycerols in acidic environments. The crucial step in this mechanism entails the nucleophilic substitution of acyl groups with chlorine ions, which are activated by neighboring ether groups. The resulting intermediate is the chloropropanediol diester that yields chloropropanols under hydrolytic conditions. 3-MCPD can also be formed through a process of heat treatment in the absence of acid-HVP. The presence of this phenomenon has been observed in sodium chloride and lipid-containing foods during cooking methods such as grilling and baking. The release of free glycerol occurs through the high-temperature hydrolysis of triglycerides. This free glycerol has the potential to react with chlorine content [55,56]. Many researchers have demonstrated that 3-MCPD, 2-MCPD, and GE are formed during oil refining, and it has been stated that the deodorization process, particularly at high temperatures, plays an important role in their formation [56,57,58,59,60,61,62]. The conversion mechanism of 2-MCPD, 3-MCPD, and GE among each other is illustrated in Figure 4.
The generation of 3-MCPD increases with moisture content up to 30% but declines with increasing humidity. The production of 3-MCPD is influenced by pH in addition to temperature and humidity. It has been observed that 3-MCPD is not stable when the pH exceeds 6 [55].
Glycidol, an organic molecule with both epoxide and alcohol functional groups, can be synthesized through the process of dehalogenation from MCPDs [9]. Glycidol and GE are quickly absorbed following food consumption. After presystemic hydrolysis, GE quickly transforms to glycidol. Glycidol undergoes rapid metabolism through various enzymatic pathways, such as glutathione conjugation and mercapturate formation. The primary route of elimination for glycidol is urinary excretion [51].
3-MCPD esters are commonly produced alongside various other chloropropanol fatty acid esters, including 1,3-DCP and 2,3-DCP. 1,3-DCP, utilized as an intermediary in the manufacturing of epichlorohydrin, a raw material used in the chemical and paper industries, may be formed from hydrochloric acid and leftover lipids from the application material [9]. Another proposed mechanism for the formation of free 1,3-DCP is through the degradation of the substance by bacteria [63]. It has been proposed that 1,3-DCP can be formed through the direct conversion to glutathione conjugates or by metabolizing to β-chlorolactaldehyde, which undergoes oxidative dehalogenation and subsequent reduction to 3-MCPD [9,22].

5. Occurrence in Foods

Chloropropanols, their esters, and GE are commonly found in foods including hydrolyzed plant proteins, soy sausage, vegetable oils and fats, margarine, baked products, smoked foods, fried foods, infant formula, and other foods.
The Rapid Alert System for Food and Feed (RASFF) reported a total of 54 notifications on 3-MCPD between the years 2020 and 2023. Most of these notifications were related to vegetable oils. A total of 36 notifications were documented for GE, whereas nine notifications were recorded for glycidol, and two notifications were recorded for 1,3-DCP. It should be noted that there were no recorded notifications for any chloropropanol compounds before the year 2020. The presence of chloropropanol has gained attention subsequent to the publication of the EFSA report. However, an analysis of the distribution of notifications over the years reveals a lack of comprehensive global efforts for mitigation. The number of RASFF notifications on chloropropanols, GE, and glycidol by product category during the period 2020–2023 is presented in Figure 5. Palm oil and other vegetable oils are the primary sources of food containing chloropropanol, followed by bakery and pastry products. The presence of 1,3-DCP is limited to food contact materials, while there is no available information regarding 2-MCPD and 2,3-DCP, potentially due to the absence of regulatory thresholds [64].

5.1. Vegetable Oils

The esters of chloropropanols, particularly 3- and 2-MCPD and GE, have been widely reported in various refined vegetable oils and fats, especially refined palm oil [65,66,67,68,69,70,71]. The occurrence and levels of fatty acid esters of 3-MCPD and 2-MCPD in vegetable oils from various countries are summarized in Table 2 and Table 3, respectively.
Among the vegetable oils, the highest middle bound (MB) values of 3-MCPD esters have been reported in palm oil/fat (2912 μg/kg), with >99 incidence, followed by palm kernel oil (100%, 624 μg/kg) and coconut oil (100%, 608 μg/kg). However, the MB levels of 3-MCPD esters in sunflower seed oil (93% incidence), maize oil (97%), soya bean oil (96%), rapeseed oil (84%), peanut oil (100%), and olive oil (100%) were 521 μg/kg, 503 μg/kg, 394 μg/kg, 232 μg/kg, 229 μg/kg, and 48 μg/kg, respectively. For 2-MCPD, the highest mean MB value was recorded in palm oil/fat (1565 μg/kg, 96% incidence) among the vegetable oil and fats group. Other vegetable oils and fats had a relatively lower MB level of 2-MCPD esters (86–270 μg/kg), with an incidence rate of 44–100%. For both fatty acid esters of 3-MCPD and 2-MCPD, the lowest MB values were recorded in olive oil [1].
Palm oil, classified as a fruit oil, has a higher water content compared to other seed oils, rendering it more susceptible to hydrolysis processes. These reactions lead to the formation of fatty acid esters of 3-MCPD and GE. The high 3-MCPD ester formation in palm oil was explained by the palm tree absorbing chlorine ions from the soil and water, which facilitates the formation of 3-MCPD. Furthermore, the synthesis of 3-MCPD esters in palm oil is directly proportional to its exposure to high temperatures during refining [77,78,79]. The study conducted by Razak et al. [73] demonstrated that the levels of 3-MCPD esters were higher in samples obtained from the bleaching and final stages of palm oil processing. In a recent study, Zhang et al. [80] discovered that the highest levels of 3-MCPD esters in palm oil can be attributed to the low activation energy required for the formation of 1,2-bis-palmitoyl-3-chloropropanediol.
The formation of 3-MCPD esters is influenced by several critical parameters, including the procedure of oil extraction from fresh fruit, fruit quality, chloride concentration, heating temperature, and heating time during the deodorization step of oil refining [72,73,76,78,80,81]. Moreover, a correlation between the presence of 3-MCPD and 1,3-DCP has been found in edible vegetable oils with a coefficient of 0.819 [82].
The occurrence and levels of GE in vegetable oils are summarized in Table 4. The highest GE level (18,000 µg/kg) was observed in the palm oil sample. Similarly, with a 99% incidence among the vegetable oil and fats group, palm oil/fat had the highest mean MB value for GE (3955 µg/kg) in the EFSA report. While no detectable level of GE was found in olive oil, the incidence of GE was 94% in maize oil (650 µg/kg), 92% in coconut oil/fat (mean MB = 476 µg/kg), 87% in palm kernel oil (421 µg/kg), 80% in sunflower seed oil (269 µg/kg), 70% in soya bean oil (171 µg/kg), 51% in rapeseed oil (166 µg/kg), and 50% in peanut oil (148 µg/kg) [1].
The two-fold higher concentrations of fatty acid esters of 3-MCPD, 2-MCPD, and GE values in the palm olein fractions were observed compared to the palm stearin fractions [67,73]. The data indicate that fatty acid esters of 3-MCPD, 2-MCPD, and GE tend to partition into the liquid phase, with a higher concentration found in the unsaturated fraction compared to the solid fraction. The explanation for this phenomenon is based on the physical properties of oils, which enable them to retain higher levels of fatty acid esters of 2-MCPD, 3-MCPD, and GE within their triacylglyceride structures compared to the solid fraction [67,73,83].
The concentration of MCPD esters increases with higher frying temperatures, while the GE content of oils decreases during the frying process. The impact of frying temperature on GE content is relatively insignificant while frying time has a more pronounced effect. The occurrence of glycidol from GEs at higher temperatures explains this phenomenon, resulting in a decrease in GE content and an increase in glycidol content during frying [84,85].

5.2. Foods Other than Vegetable Oils

Many studies have been conducted in the past two decades to investigate the concentrations of 3-MCPD, 2-MCPD, glycidol, and GE in various foodstuffs. The results of these studies are summarized in Table 5. The free and bound 3-MCPD values in other foodstuffs are significantly lower when compared to vegetable oils. However, the issue is in the excessive consumption of certain foods, which results in high exposure to compounds mostly because of the large quantities consumed rather than the actual levels of contaminants present.
In a survey by Chung et al. [13], a total of 318 different food samples consumed in Hong Kong, China, were monitored for the presence of chlorapropanols. The study revealed that approximately one-third of commonly consumed products were found to contain MCPD esters. Among various food categories, snack foods were found to have the highest levels of MCPD esters. The likely cause for this occurrence can be attributed to the oil content and the frying process employed in the production of snacks. 1,3-DCP esters were also found in 4% of the analyzed samples at lower concentrations (ranging from 1 to 9 µg/kg).
Ilko et al. [90] showed that the primary origin of 3-MCPD in potato products is the frying oils that are absorbed. Hence, it is reported that the concentrations of 3-MCPD, 2-MCPD, and GE are associated with the palm oils used as ingredients. In another study, Shi et al. [70] revealed a significant correlation between the levels of 3-MCPD esters or GE and the percentage of absorbed fat content in domestically oil-cooked food. Stauff et al. [88] measured the amounts of 2-MCPD, 3-MCPD, and GE in fine bakery products and found that these substances are present only in products made with refined vegetable oils. They also found that the GE levels, ranging from the limit of detection (LOD) to 1280 μg/kg, have been significantly decreased when compared to previous studies conducted on bakery products.
The presence of 3-MCPD has been detected in bread crust at levels of up to 400 µg/kg [97]. The formation of 3-MCPD increases when the toasted bread is exposed to high temperatures, resulting in an increased intake [55]. The formation of 3-MCPD compounds in commercial bread dough is primarily attributed to glycerol, which accounts for 68% of these compounds. During the baking process, chlorine ions and glycerol in yeast dough undergo a reaction with the precursor molecule, resulting in the formation of 3-MCPD [86].
The occurrence of chloropropanols in foods consumed by vulnerable groups, such as infants and young children, is of great concern. In the Czech Republic, 14 infant and baby food products were analyzed for the presence of 1,3-DCP, 3-MCPD, and their fatty acid esters. While none of the products contained free chloropropanols at levels above the method’s LOD, all of them contained high amounts of 3-MCPD esters. The concentrations of 3-MCPD esters in the infant and baby foods varied from 62 to 588 µg/kg, which corresponds to <300–2060 μg/kg on the fat basis [98].
In a United States survey from 2013–2016, 98 infant formula samples were tested for 3-MCPD esters and glycidol. The concentrations of 3-MCPD esters and glycidol in formulas containing palm and palm olein varied from 0.021 to 0.92 mg/kg and from less than the limit of quantification (LOQ) to 0.40 mg/kg, respectively. 3-MCPD esters and glycidol concentrations in formulas not containing palm and palm olein ranged from 72 to 160 μg/kg and from 5 to 150 μg/kg, respectively [91].
The presence of 3-MCPD in coffee is typically at very low levels, while coffee beverages do not contain detectable amounts of 3-MCPD due to the dilution effect of water [99]. The color of the coffee bean at the end of the process is closely linked to the formation of 3-MCPD and the presence of a high amount of dark beans. The presence of lipids and chlorine in coffee beans contributes to the formation of 3-MCPD during the roasting process [87,99]. Higher concentrations of 3-MCPD were detected in instant coffee, whereas the greatest concentrations were found in decaffeinated coffee and products that undergo extended roasting [89,100].
It has also been shown that smoked meat products contain 3-MCPD and its esters [101,102]. The level of 3-MCPD has been found to vary depending on the type of wood used for smoking and the duration of smoke treatment [103].
Air frying is an emerging frying technique that utilizes a newly designed appliance for home cooking. It involves the use of minimal or even no added oil for the frying process. A study was conducted by Ostermeyer et al. [104] to address concerns regarding the potential increase of chloropropanols resulting from the use of an air fryer. The effect of different domestic cooking procedures (baking in an electric oven, hot-air fryer, roasting in a pan, and roasting in a mini kitchen fryer) on the MCPD and GE amounts of fish fingers was determined. Sunflower oil was employed for the purpose of stir-frying, while frying oil was not utilized in the hot air fryer. The longer cooking duration in the hot air fryer resulted in a higher reduction in product weight compared to the oven. The levels of 3-MCPD and GE remain unchanged when food is cooked using a hot air fryer. The utilization of a hot air fryer can potentially impact the weight loss of food items under specific circumstances, although it appears to have minimal influence on their composition. The concentration of chloropropanol present in ready-to-eat frozen fish is influenced by the initial level of contaminants in the product and/or oil before cooking. On the contrary, Goh et al. [105] demonstrated that the use of an air fryer can still result in the production of 2- and 3-MCPD and GE, attributed to the elevated temperature and duration of the cooking process. In addition, the formation of chloropropanols can be facilitated by the presence of chloride ions, moisture, and partial acylglycerol.
Together with 3-MCPD, other chloropropanols such as 2-MCPD, 1,3-DCP, 2,3-DCP, glycidol, and GE have been found to occur in a range of foodstuffs. In addition to the presence of chloropropanol derivatives in foodstuffs, 1,3-DCP may also be found in drinking water. This is owing to its existence as a contaminant of epichlorohydrin-linked cationic polymer resins in flocculants used for water purification in a restricted number of treatment facilities [12]. The occurrence and levels of glycidol and chloropropanols other than 3-MCPD in foodstuffs are summarized in Table 6.
The available data regarding the content of 2-MCPD in food was primarily limited to soy sauces and oils before the publication of the EFSA report in 2016. It was observed that the level of 2-MCPD was lower than that of 3-MCPD, suggesting a potential correlation between these two contaminants. The levels of 2-MCPD were determined to be significantly low in dairy products, cereal products, meat, and fish. The presence of the substance was not detected in fruits and vegetables, confectionaries, bakery wares, salt, spices, soups and sauces (excluding soy sauce), coffee, and tea [12]. It has been reported that a correlation between 2-MCPD and 3-MCPD presence in various foodstuffs was found specifically in a 1:10 proportion [1]. Although free 2-MCPD is typically found in lower quantities than 3-MCPD in food, it exhibits greater resistance to the method used to reduce 3-MCPD [108].
It has been reported that all samples that tested positive for 1,3-DCP were also found to contain 3-MCPD. Additionally, the concentration of 1,3-DCP was noticeably lower compared to that of 3-MCPD. Meat, malt, and fish products were found to have very low quantities of 1,3-DCP, but cereals, spices, and sauces (excluding soy sauce) did not contain it [12]. However, Chung et al. [13] found higher amounts of 1,3-DCP in roasted pork and sausage than 3-MCPD. In another study, 1,3-DCP levels were found to be greater than 3-MCPD in cereals, snacks, toasted bread, and biscuits [109]. Zelinkova et al. [98] reported that the presence of 1,3-DCP was not detected in infant formula, follow-up formula, or growing-up milk. The presence of 1,3-DCP in food contact material, particularly in paper, has recently gained attention. This is due to the use of bleaching agents that contain chlorine [110].

6. Analytical Method

The analysis of chloropropanol derivatives is challenging due to their wide variety and structural complexity. There are two analytical approaches, namely direct and indirect, that can be used to determine 3-MCPD and related substances in foodstuffs [111,112].
The direct analytical approach is based on the quantitative determination of 3-MCPD and GE using liquid chromatography–mass spectrophotometry (LC-MS/MS) by analyzing each fatty acid that makes up the structure of mono and diglycerides separately. The analysis of GE involves purification using solid phase extraction (SPE) and gel permeation chromatography (GPC). The American Oil Chemists’ Society (AOCS) standardized the direct analytical method for determining GE in oils. However, the direct method for measuring 3-MCPD in food has not been fully validated. The fact that GE may be identified separately from MCPD and that the direct technique is simpler to use has increased the adoption of LC-MS/MS in recent years [113].
The indirect analytical approach is based on the quantitative examination by gas chromatography–mass spectroscopy (GC-MS/MS) of 3-MCPD and glycidyl components, which are esterified with fatty acid under acidic or alkaline conditions. Although several methods have been developed for indirect analysis, the purification, derivatization, and GC-MS/MS procedures are the same across all approaches. These approaches calculate the total MCPD or the sum of MCPD and GE in the food sample [108,114]. Although colorimetric and GC with flame ionization detection have been employed for detection in the air matrix, there is no currently appropriate method available for detecting unstable free glycidols in food [1].
The detection of 1,3-DCP and 2,3-DCP can be achieved using indirect methods, such as the elution of the compounds with a mixture of hexane and ether or ethyl acetate [55]. In another study, Cao et al. [115] developed a method for detecting 1,3-DCP and 2,3-DCP, as well as 3-MCPD, using 1-trimethylsilylimidazole as a silylating agent and GC-MS/MS.
In the analysis of chloropropanols and their esters, as well as glycidyl fatty acid esters, it is critical to report the LOD and LOQ to ensure the reliability and sensitivity of the analytical methods used. In chromatographic methods, the LOD for chlorapropanols and their derivatives can range from as low as 1 to 100 µg/kg [75,94,112,116,117,118,119], depending on the food matrix, sample preparation techniques, and specific analytical approach used (direct vs. indirect analysis). The LOQ for these contaminants typically falls within a slightly higher range, often reported between 5 and 200 µg/kg [1,75,76,94,112,117,119,120].
Apart from chromatographic techniques, a few approaches for the detection of chlorapropanols have been proposed, such as electrochemical and optical detections. Electrochemical methods, such as those using gold nanoparticle-enhanced sensors [121] and magnetic molecularly imprinted polymers [122], offer high sensitivity with detection limits down to 0.25 mg/L. Additionally, optical methods, including fluorescence-based assays on paper substrates [123], provide rapid analysis with LODs as low as 0.6 ng/mL. These methods complement traditional instrumental techniques by offering alternative approaches for detecting chloropropanols in food matrices.

7. Mitigation

Mitigating the formation of chloropropanols in food products is a complex but crucial endeavor involving a range of targeted interventions aimed at reducing their presence or preventing their formation during processing. Several strategies exist to prevent and/or reduce the formation of chloropropanols in foods. These strategies include the following:
  • Removing precursor compounds from the raw material;
  • Removing chloropropanols from the final product;
  • Preventing chloropropanol synthesis by optimizing process parameters.
One precaution that can be performed before oil refining is the elimination of chlorine [115]. The chlorinated components present in crude oils can be effectively reduced by washing the oils with different solvents. The best results were obtained by washing the palm fruit pulp before oil extraction [52,124]. Additional precautions can be taken to mitigate the acidity that leads to the formation of contaminants. The acidity of the oil can be kept low by following certain precautions. These include carefully selecting the fruits that will be processed into oil, promptly collecting and sterilizing the fruits to deactivate the lipase enzyme, washing the fruits from palm trees that grow in acidic soil before processing them into raw oil, and storing the oil seeds at temperatures below 25 °C and with moisture content ideally below 7% [125,126].
Various adsorbents can be employed to eliminate chloropropanol derivatives from food. Zeolite and synthetic magnesium silicate resulted in a 40% reduction in MCPD esters [127]. An alternative approach involves the enzymatic conversion of 3-MCPD and its esters into glycerol using bacterial enzymes [128]. Additionally, the formation of contaminants can be prevented by introducing antioxidants to the oil. However, the inclusion of adsorbents, enzymes, and antioxidants appears to be impractical due to the need for supplementary procedural steps. However, the addition of antioxidants, such as α-tocopherol, stigmasterol, and squalene, to vegetable oil at higher concentrations led to the stimulation and increased formation of MCPD esters [129].
In recent years, researchers have concentrated on the anti-inflammatory, antioxidant, and chemopreventive characteristics of bioactive compounds, particularly those that protect against chloropropanol-induced toxic damage [130,131,132,133,134,135]. Flavonoids have been found to offer protection against the harmful effects of chloropropanols. Anthocyanins, notably cyanidin-3-glucoside, present in fruits and vegetables, have antioxidant characteristics and control gut microbiota, protecting against 3-MCPD-induced damage to the intestines and testis. Apigenin, found in fruits, vegetables, and plant-derived beverages, protects against 3-MCPD-induced kidney damage by altering mitochondrial activity. Furthermore, microRNAs, specifically miR-223-3p, play an important role in preventing acute kidney injury caused by 3-MCPD esters by inhibiting RIPK3 expression [131,132,133,134].
The CAC [126] advises implementing measures to reduce the levels of 3-MCPD esters and GE in refined oils and food products made with refined oils. These measures include following good agricultural practices (GAP), good manufacturing practices (GMP), and carefully selecting and using refined oils. It is recommended to reduce 3-MCPD formation in vegetable oil beginning with the early phases of vegetable plantation on the farm. Promoting the use of chlorine-free fertilizers, specifically supporting alternatives such as magnesium-rich synthetic gypsum (MRSG), may be beneficial for the low chlorine content of vegetables and oils [59]. Despite some positive developments in reducing contaminants, the desired outcome has not yet been achieved at the industrial level. Further investigation is required to reduce the chloropropanol formation in food products.

8. Legal Regulations

Several countries have established regulations for 3-MCPD and GE in hydrolyzed vegetable proteins, soy sauce, vegetable oils and fats, and infant formulas. The maximum levels (ML) of 3-MCPD in soy sauce in various countries are listed in Table 7. As can be seen in Table 7, the MLs of 3-MCPD in soy sauce vary from 20 µg/kg (in Europe, Turkey, and Malaysia) to 1000 µg/kg (in Japan, Canada, Taiwan, and the United States).
Worldwide, the most extensive regulations for 3-MCPD and GE have been established in the European Union (EU). For 3-MCPD, an ML of 20 µg/kg in hydrolyzed vegetable proteins and soy sauce has been established by Commission Regulation (EC) No 1881/2006 [145]. The original MLs for the sum of 3-MCPD and 3-MCPD fatty acid esters (expressed as 3-MCPD) and GE (expressed as glycidol) in certain foods, including specific food products for infants and young children were established by Regulation (EU) 2020/1322 [146]. The MLs in vegetable oils and fats, fish oils, and oils from other marine organisms are in the range of 1250 to 2500 µg/kg for 3-MCPD and 1000 µg/kg for glycidol. The MLs of 750 μg/kg 3-MCPD and 500 μg/kg glycidol are applied to vegetable oils and fats, fish oils, and oils from marine organisms that are specifically used to produce baby food and processed cereal-based food for infants and young children. The MLs of 3-MCPD and glycidol for baby formulae, follow-on formulae, food for specific medical purposes, and young-child formulae vary, with limitations of 125 and 50 μg/kg for powdered items and 15 and 6 μg/kg for liquids. However, the EU has reduced the MLs for the sum of 3-MCPD and 3-MCPD fatty acid esters in certain food products for infants and young children from 125 μg/kg to 80 μg/kg (for powder form) and from 15 μg/kg to 12 μg/kg (for liquid form). These new MLs will apply from 1 January 2025 [147].
No EU regulations were available for other chloropropanol derivates such as 2-MCPD, 1,3-DCP, and 2,3-DCP. The recommended MLs for 1,3-DCP in foodstuffs varied from 5 µg/kg in New Zealand and Australia [148] to 50 µg/kg in the United States and Switzerland [63].

9. Conclusions

Food products that contain carcinogenic chloropropanol derivatives can be consumed directly or used in the manufacture of other foods. The consumption of oils and foods containing these pollutants has been reported to be detrimental to health. Hence, it is crucial to implement measures to mitigate the occurrence of chloropropanol derivatives in food products and to eliminate any pre-existing contaminants. The presence of 3-MCPD in soy sauce has been well-known for decades, but a variety of foods have been shown to contain chloropropanol derivates. The EFSA report has highlighted increasing concerns regarding potential health risks associated with glycidol. The manufacturers focus their efforts on mitigating the presence of 3-MCPD and GE, as these substances have stringent restrictions. However, measures to reduce glycidol have not yet been put into action. The levels of exposure to 1,3-DCP and 2,3-DCP are quite low; however, the levels of GE have caused more significant concerns due to their larger concentrations. The utilization of 1,3-DCP and 2,3-DCP in chemical processes gives rise to concerns regarding potential toxic effects caused by the discharge of their waste into soil and water. It is important to give greater consideration not just to 3-MCPD but also to other chloropropanol derivatives.
In the future, research on chloropropanols and their esters is expected to focus on easier-to-use sample preparation techniques and detection methods, enhancing both sensitivity and specificity, especially for compounds like 1,3-DCP and 2,3-DCP. Efforts will also be geared towards identifying safer processing technologies to reduce the formation of these contaminants, possibly through novel enzyme treatments or alternative food processing methods that minimize high temperatures and the use of salts. Regulatory frameworks are expected to evolve in response to emerging toxicological data, resulting in stricter permissible limits and possibly new guidelines for previously overlooked contaminants, such as lesser-studied chloropropanols.

Author Contributions

Conceptualization, G.O., M.Á.G.-C., and B.K.; investigation, G.O. and B.K.; writing—original draft preparation, G.O., M.Á.G.-C., and B.K.; writing—review and editing, M.Á.G.-C. and B.K. All authors have read and agreed to the published version of the manuscript.

Funding

This research received no external funding. The APC was funded by Universidad EAN.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

No new data were created or analyzed in this study. Data sharing is not applicable to this article.

Conflicts of Interest

The authors declare no conflicts of interest.

References

  1. EFSA (European Food Safety Authority). Risks for human health related to the presence of 3- and 2-monochloropropanediol (MCPD), and their fatty acid esters, and glycidyl fatty acid esters in food. EFSA J. 2016, 14, 4426. [Google Scholar]
  2. Velisek, J.; Davidek, J.; Hajšlová, J.; Kubelka, V.; Janíček, G.; Mánková, B. Chlorohydrins in protein hydrolysates chlorhydrinein. Z. Lebensm. Unters. Forsch 1978, 167, 241–244. [Google Scholar] [CrossRef] [PubMed]
  3. Davidek, J.; Velíšek, J.; Kubelka, V.; Janíček, G.; Šimicová, Z. Glycerol chlorohydrins and their esters as products of the hydrolysis of tripalmitin, tristearin and triolein with hydrochloric acid. Z. Lebensm. Unters. Forsch 1980, 171, 14–17. [Google Scholar] [CrossRef]
  4. Zelinkova, Z.; Svejkovska, B.; Velisek, J.; Dolezal, M. Fatty acid esters of 3-chloropropane-1,2-diol in edible oils. Food Addit. Contam. 2006, 23, 1290–1298. [Google Scholar] [CrossRef]
  5. CVUA (Chemisches- und Veterinäruntersuchungsamt Stuttgart Sitz Fellbach). 3-MCPD ester in Raffinierten Speisefetten und ölen—Ein neu Erkanntes Weltweits Problem. 2007. Available online: http://www.cvuas.de/pub/beitrag.asp?ID=717&subid=1 (accessed on 11 March 2024).
  6. Zelinková, Z.; Novotný, O.; Schůrek, J.; Velíšek, J.; Hajšlová, J.; Doležal, M. Occurrence of 3-MCPD fatty acid esters in human breast milk. Food Addit. Contam. 2008, 25, 669–676. [Google Scholar] [CrossRef]
  7. Casado, N.; Berenguer, C.V.; Câmara, J.S.; Pereira, J.A.M. What are we eating? surveying the presence of toxic molecules in the food supply chain using chromatographic approaches. Molecules 2024, 29, 579. [Google Scholar] [CrossRef]
  8. SGS (Société Générale de Surveillance). Available online: https://www.sgs.com/en-hk/news/2022/03/20220310-sgs-risk-of-chloropropanols-in-fcm-paper-products (accessed on 10 May 2024).
  9. Andres, S.; Appel, K.E.; Lampen, A. Toxicology, occurrence and risk characterisation of the chloropropanols in food: 2-monochloro-1,3-propanediol, 1,3-dichloro-2-propanol and 2,3-dichloro-1-propanol. Food Chem. Toxicol. 2013, 58, 467–478. [Google Scholar] [CrossRef]
  10. Joint FAO/WHO Expert Committee on Food Additives. Summary of the Fifty-Seventh Meeting of the Joint FAO/WHO Expert Committee on Food Additives (JECFA), Rome, Italy, 5–14 June 2004; pp. 20–24. Available online: http://www.leffingwell.com/Summary%20and%20Conclusions%20of%20the%20Fifty-seventh%20meeting.pdf (accessed on 11 May 2024).
  11. Crews, C.; Hasnip, S.; Chapman, S.; Hough, P.; Potter, N.; Todd, J.; Brereton, P.; Matthews, W. Survey of chloropropanols in soy sauces and related products purchased in the UK in 2000 and 2002. Food Addit. Contam. 2003, 20, 916–922. [Google Scholar] [CrossRef]
  12. EC (European Commission). Collection and Collation of Data on Levels of 3-Monochloropropanediol (3-MCPD) and Related Substances in Foodstuffs, Reports on Tasks for Scientific Cooperation. 2004. Available online: https://food.ec.europa.eu/system/files/2016-10/cs_contaminants_catalogue_mcpd_scoop_3-2-9_final_report_chloropropanols_en.pdf (accessed on 11 March 2024).
  13. Chung, S.W.C.; Kwong, K.P.; Yau, J.C.W.; Wong, A.M.C.; Xiao, Y. Chloropropanols levels in foodstuffs marketed in Hong Kong. J. Food Compos. Anal. 2008, 21, 569–573. [Google Scholar] [CrossRef]
  14. IARC (International Agency for Research on Cancer). Some chemicals present in industrial and consumer products, food and drinking-water. IARC Monogr. Eval. Carcinog. Risks Hum. 2013, 101, 349–374. [Google Scholar]
  15. NTP (National Toxicology Program). NTP Toxicology and Carcinogenesis Studies of Glycidol (CAS No. 556-52-5) in F344/N Rats and B6C3F1 Mice (Gavage Studies); Technical Report Series No. 374; National Institutes of Health Publication No. 90-2829; NIH Publication: Chapel Hill, NC, USA, 1990.
  16. Seefelder, W.; Scholz, G.; Schilter, B. Structural diversity of dietary fatty esters of chloropropanols and related substances. Eur. J. Lipid Sci. Technol. 2011, 113, 319–322. [Google Scholar] [CrossRef]
  17. Tanguler, H.; Kabak, B. Chemical hazards in foods. In Health and Safety Aspects of Food Processing Technologies; Malik, A., Erginkaya, Z., Erten, H., Eds.; Springer International Publishing: Cham, Swizterland, 2019; pp. 349–402. [Google Scholar]
  18. Lee, B.Q.; Khor, S.M. 3-chloropropane-1,2-diol (3-MCPD) in soy sauce: A review on the formation, reduction, and detection of this potential carcinogen. Compr. Rev. Food Sci. Food Saf. 2015, 14, 48–66. [Google Scholar] [CrossRef] [PubMed]
  19. Velisek, J.; Doležal, M.; Crews, C.; Dvořák, T. Optical isomers of chloropropanediols: Mechanisms of their formation and decomposition in protein hydrolysates. Czech J. Food Sci. 2002, 20, 161–170. [Google Scholar] [CrossRef]
  20. NCBI (National Center for Biotechnology Information). PubChem Compound Summary for CID 7289, 1,3-Dichloro-2-propanol. Available online: https://pubchem.ncbi.nlm.nih.gov/compound/1_3-Dichloro-2-propanol (accessed on 7 March 2024).
  21. NCBI (National Center for Biotechnology Information). PubChem Compound Summary for CID 12018, 2,3-Dichloro-1-propanol. Available online: https://pubchem.ncbi.nlm.nih.gov/compound/2_3-Dichloro-1-propanol (accessed on 7 March 2024).
  22. Schilter, B.; Scholz, G.; Seefelder, W. Fatty acid esters of chloropropanols and related compounds in food: Toxicological aspects. Eur. J. Lipid Sci. Technol. 2011, 113, 309–313. [Google Scholar] [CrossRef]
  23. Habermeyer, M.; Guth, S.; Eisenbrand, G. Identification of gaps in knowledge concerning toxicology of 3-MCPD and glycidol esters. Eur. J. Lipid Sci. Technol. 2011, 113, 314–318. [Google Scholar] [CrossRef]
  24. NCBI (National Center for Biotechnology Information). PubChem Compound Summary for CID 11164, Glycidol. Available online: https://pubchem.ncbi.nlm.nih.gov/compound/Glycidol (accessed on 7 March 2024).
  25. Kobets, T.; Smith, B.P.C.; Williams, G.M. Food-borne chemical carcinogens and the evidence for human cancer risk. Foods 2022, 11, 2828. [Google Scholar] [CrossRef]
  26. IARC (International Agency for Research on Cancer). Some industrial chemicals. IARC Monogr. Eval. Carcinog. Risk Chem. Hum. 2000, 77, 469–487. [Google Scholar]
  27. Barocelli, E.; Corradi, A.; Mutti, A.; Petronini, P.G. Comparison between 3-MCPD and its palmitic esters in a 90-day toxicological study. EFSA Support. Publ. 2011, 8, 1–131. [Google Scholar] [CrossRef]
  28. Li, J.; Wang, S.; Wang, M.; Shi, W.; Du, X.; Sun, C. The toxicity of 3-chloropropane-1,2-dipalmitate in Wistar rats and a metabonomics analysis of rat urine by ultra-performance liquid chromatography-mass spectrometry. Chem. Biol. Interact. 2013, 206, 337–345. [Google Scholar] [CrossRef]
  29. Onami, S.; Cho, Y.; Toyoda, T.; Mizuta, Y.; Yoshida, M.; Nishikawa, A.; Ogawa, K. A 13-week repeated dose study of three 3-monochloropropane-1,2-diol fatty acid esters in F344 rats. Arch. Toxicol. 2014, 88, 871–880. [Google Scholar] [CrossRef]
  30. Lynch, B.S.; Bryant, D.W.; Hook, G.J.; Earle, N.R.; Ian, C.M. Carcinogenicity of monochloro-1,2-propanediol (α-chlorohydrin, 3-MCPD). Int. J. Toxicol. 1998, 17, 47–76. [Google Scholar] [CrossRef]
  31. WHO (World Health Organization). Safety Evaluation of Certain Food Additives and Contaminants—3-Chloro-1,2-propanediol; WHO Food Additives Series; World Health Organization: Geneva, Switzerland, 2007; Volume 58. Available online: https://www.google.com/url?sa=t&source=web&rct=j&opi=89978449&url=https://apps.who.int/food-additives-contaminants-jecfa-database/Document/Index/9008&ved=2ahUKEwj39LKf_LWIAxUkSvEDHcvjAicQFnoECBQQAQ&usg=AOvVaw2YlbF3xXrl76Pnx0-qxzN5 (accessed on 15 March 2024).
  32. Cho, W.S.; Han, B.S.; Lee, H.; Kim, C.; Nam, K.T.; Park, K.; Choi, M.; Kim, S.J.; Kim, S.H.; Jeong, J.; et al. Subchronic toxicity study of 3-monochloropropane-1,2-diol administered by drinking water to B6C3F1 mice. Food Chem. Toxicol. 2008, 46, 1666–1673. [Google Scholar] [CrossRef]
  33. Jones, A.R. Antifertility actions of alpha-chlorohydrin in the male. Int. J. Biol. Sci. 1983, 36, 333–350. [Google Scholar]
  34. El Ramy, R.; Ould Elhkim, M.; Lezmi, S.; Poul, J.M. Evaluation of the genotoxic potential of 3-monochloropropane-1,2-diol (3-MCPD) and its metabolites, glycidol and beta-chlorolactic acid, using the single cell gel/comet assay. Food Chem. Toxicol. 2007, 45, 41–48. [Google Scholar] [CrossRef]
  35. Lee, J.K.; Byun, J.A.; Park, S.H.; Kim, H.S.; Park, J.H.; Eom, J.H.; Oh, H.Y. Evaluation of the potential immunotoxicity of 3-monochloro-1, 2-propanediol in Balb/c mice: I. Effect on antibody forming cell, mitogen-stimulated lymphocyte proliferation, splenic subset, and natural killer cell activity. Toxicology 2004, 204, 1–11. [Google Scholar] [CrossRef]
  36. Hamlet, C.G.; Sadd, P.A.; Crews, C.; Velíšek, J.; Baxter, D.E. Occurrence of 3-chloro-propane-1,2-diol (3-MCPD) and related compounds in foods: A review. Food Addit. Contam. 2002, 19, 619–631. [Google Scholar] [CrossRef]
  37. Jones, A.R.; Fakhouri, G. Epoxides as obligatory intermediates in the metabolism of a-halohydrins. Xenobiotica 1979, 9, 595–599. [Google Scholar] [CrossRef]
  38. Qian, G.; Zhang, H.; Zhang, G.; Yin, L. Study on acute toxicity of R, S and (R,S)-3-monchloropropane-1, 2-diol. J. Hyg. Res. 2007, 36, 137–140. [Google Scholar]
  39. Kluwe, W.M.; Gupta, B.N.; Lamb, J.C. The comparative effects of 1,2-dibromo-3-chloropropane (DBCP) and metabolites, 3-chloro 1,2-propaneoxide (epichlorohydrin), 3-chloro-1,2-propanediol (alphachlorohydrin), and oxalic acid, on the urogenital system of male rats. Toxicol. Appl. Pharmacol. 1983, 70, 67–86. [Google Scholar] [CrossRef]
  40. Lewis, R.J. Sax’s Dangerous Properties of Industrial Materials, 9th ed.; Van Nostrand Reinhold: New York, NY, USA, 1996; Volume 1–3, p. 1124. [Google Scholar]
  41. Lewis, R.J., Sr. (Ed.) Sax’s Dangerous Properties of Industrial Materials, 11th ed.; Wiley-Interscience, Wiley & Sons, Inc.: Hoboken, NJ, USA, 2004; p. 1869. [Google Scholar]
  42. Kawashima, M.; Watanabe, Y.; Nakajima, K.; Murayama, H.; Nagahara, R.; Jin, M.; Yoshida, T.; Shibutani, M. Late effect of developmental exposure to glycidol on hippocampal neurogenesis in mice: Loss of parvalbumin-expressing interneurons. Exp. Toxicol. Pathol. 2017, 69, 517–526. [Google Scholar] [CrossRef]
  43. Akane, H.; Shiraki, A.; Imatanaka, N.; Akahori, Y.; Itahashi, M.; Abe, H.; Shibutani, M. Glycidol induces axonopathy and aberrations of hippocampal neurogenesis affecting late-stage differentiation by exposure to rats in a framework of 28-day toxicity study. Toxicol. Lett. 2014, 224, 424–432. [Google Scholar] [CrossRef]
  44. Bakhiya, N.; Abraham, K.; Gürtler, R.; Appel, K.E.; Lampen, A. Toxicological assessment of 3-chloropropane-1,2-diol and glycidol fatty acid esters in food. Mol. Nutr. Food Res. 2011, 55, 509–521. [Google Scholar] [CrossRef]
  45. Rae, J.M.C.; Craig, L.; Slone, T.W.; Frame, S.R.; Buxton, L.W.; Kennedy, G.L. Evaluation of chronic toxicity and carcinogenicity of ammonium 2,3,3,3-tetrafluoro-2-(heptafluoropropoxy)-propanoate in Sprague-Dawley rats. Toxicol. Rep. 2015, 2, 939–949. [Google Scholar]
  46. Aasa, J.; Granath, F.; Törnqvist, M. Cancer risk estimation of glycidol based on rodent carcinogenicity studies, a multiplicative risk model and in vivo dosimetry. Food Chem. Toxicol. 2019, 128, 54–60. [Google Scholar] [CrossRef]
  47. Mihalache, O.A.; Dall’Asta, C. Food processing contaminants: Dietary exposure to 3-MCPD and glycidol and associated burden of disease for Italian consumers. Environ. Res. 2023, 234, 116559. [Google Scholar] [CrossRef]
  48. FSCJ (Food Safety Commission of Japan). Considerations on glycidol and its fatty acid esters in foods. Exec. Summ. 2015, 3, 67–69. Available online: https://www.jstage.jst.go.jp/article/foodsafetyfscj/3/2/3_2015010e/_pdf/-char/en (accessed on 20 March 2024).
  49. Committee on Mutagenicity of Chemicals in Food, Consumer Products and the Environment. Annual Report. 2004. Available online: https://cot.food.gov.uk/sites/default/files/cot/comsection.pdf (accessed on 5 March 2024).
  50. Committee on the Carcinogenicity of Chemicals in Food, Consumer Products and the Environment. Annual Report. 2004. Available online: https://cot.food.gov.uk/sites/default/files/cot/cocsection.pdf (accessed on 5 March 2024).
  51. Hamlet, C.G.; Asuncion, L.; Velíšek, J.; Doležal, M.; Zelinková, Z.; Crews, C. Formation and occurrence of esters of 3-chloropropane-1, 2-diol (3-CPD) in foods: What we know and what we assume. Eur. J. Lipid Sci. Technol. 2011, 113, 279–303. [Google Scholar] [CrossRef]
  52. Craft, B.D.; Nagy, K.; Sandoz, L.; Destaillats, F. Factors impacting the formation of monochloropropanediol (MCPD) fatty acid diesters during palm (Elaeis guineensis) oil production. Food Addit. Contam. Part A 2012, 29, 354–361. [Google Scholar] [CrossRef]
  53. Seefelder, W.; Varga, N.; Studer, A.; Williamson, G.; Scanlan, F.P.; Stadler, R.H. Esters of 3-chloro-1,2-propanediol (3-MCPD) in vegetable oils: Significance in the formation of 3-MCPD. Food Addit. Contam. 2008, 25, 391–400. [Google Scholar] [CrossRef]
  54. Jędrkiewicz, R.; Kupska, M.; Głowacz, A.; Gromadzka, J.; Namieśnik, J. 3-MCPD: A worldwide problem of food chemistry. Crit. Rev. Food Sci. Nutr. 2016, 56, 2268–2277. [Google Scholar] [CrossRef]
  55. Baer, I.; Calle, B.; Taylor, P. 3-MCPD in food other than soy sauce or hydrolyzed vegetable protein (HVP). Anal. Bioanal. Chem. 2010, 396, 443–456. [Google Scholar] [CrossRef]
  56. Hori, K.; Hori-Koriyama, N.; Tsumura, K.; Fukusaki, E.; Bamba, T. Insights into the formation mechanism of chloropropanol fatty acid esters under laboratory-scale deodorization conditions. J. Biosci. Bioeng. 2016, 122, 246–251. [Google Scholar] [CrossRef]
  57. Matthäus, B.; Pudel, F.; Fehling, P.; Vosmann, K.; Freudenstein, A. Strategies for the reduction of 3-MCPD esters and related compounds in vegetable oils. Eur. J. Lipid Sci. Technol. 2011, 113, 380–386. [Google Scholar] [CrossRef]
  58. Destaillats, F.; Craft, B.D.; Sandoz, L.; Nagy, K. Formation mechanisms of monochloropropanediol (MCPD) fatty acid diesters in refined palm (Elaeis guineensis) oil and related fractions. Food Addit. Contam. Part A 2012, 29, 29–37. [Google Scholar] [CrossRef]
  59. Arris, F.A.; Thai, V.T.S.; Manan, W.N.; Sajab, M.S. A revisit to the formation and mitigation of 3-chloropropane-1,2-diol in palm oil production. Foods 2020, 9, 1769. [Google Scholar] [CrossRef]
  60. Tivanello, R.; Capristo, M.; Vicente, E.; Ferrari, R.; Sampaio, K.; Arisseto, A. Effects of deodorization temperature and time on the formation of 3-MCPD, 2-MCPD, and glycidyl esters and physicochemical changes of palm oil. J. Food Sci. 2020, 85, 2255–2260. [Google Scholar] [CrossRef]
  61. Ji, Y.; Lan, D.; Wang, W.; Goh, K.M.; Tan, C.P.; Wang, Y. The formation of 3-monochloropropanediol esters and glycidyl esters during heat-induced processing using an olive-based edible oil. Foods 2022, 11, 4073. [Google Scholar] [CrossRef]
  62. Sun, C.; Wu, N.; Kou, S.; Wu, H.; Liu, Y.; Pei, A. Occurrence, formation mechanism, detection methods, and removal approaches for chloropropanols and their esters in food: An updated systematic review. Food Chem. X 2023, 17, 100529. [Google Scholar] [CrossRef]
  63. IARC (International Agency for Research on Cancer). IARC Monographs on the Evaluation of the Carcinogenic Risks to Humans. 1,3-Dichloro-2-propanol. 2012. Available online: https://www.google.com/url?sa=t&source=web&rct=j&opi=89978449&url=https://publications.iarc.fr/_publications/media/download/5695/5a8db1ef7524cce85ad21b3836bea40ab9082535.pdf&ved=2ahUKEwjitZCv8LeIAxWPR_EDHc2sJ4YQFnoECBUQAQ&usg=AOvVaw0cu25qbzNOoJSfM5u3j6O4 (accessed on 3 March 2024).
  64. RASFF (Rapid Alarm System for Food and Feed Portal). Available online: https://ec.europa.eu/food/safety/rasff_en (accessed on 4 March 2024).
  65. Pudel, F.; Benecke, P.; Fehling, P.; Freudenstein, A.; Matthaus, B.; Schwaf, A. On the necessity of edible oil refining and possible sources of 3-MCPD and glycidyl esters. Eur. J. Lipid Sci. Technol. 2011, 113, 368–373. [Google Scholar] [CrossRef]
  66. MacMahon, S.; Begley, T.H.; Diachenko, G.W. Occurrence of 3-MCPD and glycidyl esters in edible oils in the United States. Food Addit. Contam. A. 2013, 30, 2081–2092. [Google Scholar] [CrossRef] [PubMed]
  67. Razak, R.A.A.; Tarmizi, A.H.A.; Hammid, A.N.A.; Kuntom, A.; Ismail, I.S.; Sanny, M. Verification and evaluation of monochloropropanediol (MCPD) esters and glycidyl esters in palm oil products of different regions in Malaysia. Food Addit. Contam. Part A 2019, 36, 1626–1636. [Google Scholar] [CrossRef] [PubMed]
  68. Albuquerque, T.G.; Costa, H.S.; Silva, M.A.; Oliveira, M.B.P.P. Are chloropropanols and glycidyl fatty acid esters a matter of concern in palm oil? Trends Food Sci. Technol. 2020, 105, 494–514. [Google Scholar] [CrossRef]
  69. Eisenreich, A.; Monien, B.H.; Götz, M.E.; Buhrke, T.; Oberemm, A.; Schultrich, K.; Abraham, K.; Braeuning, A.; Schäfer, B. 3-MCPD as contaminant in processed foods: State of knowledge and remaining challenges. Food Chem. 2023, 403, 134332. [Google Scholar] [CrossRef]
  70. Shi, R.R.S.; Shen, P.; Yu, W.Z.; Cai, M.; Tay, A.J.; Lim, I.; Chin, Y.S.; Ang, W.M.; Er, J.C.; Lim, G.S. Occurrence and dietary exposure of 3-MCPD esters and glycidyl esters in domestically and commercially prepared food in Singapore. Foods 2023, 12, 4331. [Google Scholar] [CrossRef]
  71. Liao, Z.; Gao, Z.; Yang, Q.; Cao, D. Occurrence and exposure evaluation of 2- and 3-monochloropropanediol (MCPD) esters and glycidyl esters in refined vegetable oils marketed in Tianjin of China. J. Food Compos. Anal. 2024, 130, 106150. [Google Scholar] [CrossRef]
  72. Li, C.; Nie, S.P.; Zhou, Y.Q.; Xie, M.Y. Exposure assessment of 3-monochloropropane-1,2-diol esters from edible oils and fats in China. Food Chem. Toxicol. 2015, 75, 8–13. [Google Scholar] [CrossRef] [PubMed]
  73. Razak, R.A.A.; Kuntom, A.; Siew, W.L.; Ibrahim, N.A.; Ramli, M.R.; Hussein, R.; Nesaretnam, K. Detection and monitoring of 3-monochloropropane-1,2-diol (3-MCPD) esters in cooking oils. Food Control 2012, 25, 355–360. [Google Scholar] [CrossRef]
  74. Kamikata, K.; Vicente, E.; Arisseto-Bragotto, A.P.; de Oliveira Miguel, A.M.R.; Milani, R.F.; Tfouni, S.A.V. Occurrence of 3-MCPD, 2-MCPD and glycidyl esters in extra virgin olive oils, olive oils and oil blends and correlation with identity and quality parameters. Food Control 2019, 95, 135–141. [Google Scholar] [CrossRef]
  75. Jedrkiewicz, R.; Głowacz, A.; Gromadzka, J.; Namie, J. Determination of 3-MCPD and 2-MCPD esters in edible oils, fish oils and lipid fractions of margarines available on Polish market. Food Control 2016, 59, 487–492. [Google Scholar] [CrossRef]
  76. Kuhlmann, J. Determination of bound 2, 3-epoxy-1-propanol (glycidol) and bound monochloropropanediol (MCPD) in refined oils. Eur. J. Lipid Sci. Technol. 2011, 113, 335–344. [Google Scholar] [CrossRef]
  77. Shahidi, F.; Zhong, Y. Lipid oxidation: Measurement methods. In Bailey’s Industrial Oil and Fat Products; Shahidi, F., Ed.; John Wiley & Sons Inc.: Hoboken, NJ, USA, 2005; pp. 357–385. [Google Scholar]
  78. Weisshaar, R. 3-MCPD-esters in edible fats and oils–a new and worldwide problem. Eur. J. Lipid Sci. Technol. 2008, 110, 671–672. [Google Scholar] [CrossRef]
  79. CAC (Codex Alimentarius Commission). Proposed Draft Code of Practice for the Reduction of 3-Monochloropropane-1,2-diol Esters (3-MCPDE) and Glycidyl Esters (GE) in Refined Oils and Products Made with Refined Oils, Especially Infant Formula. Joint FAO/WHO Food Standards Programme. 2018. Available online: http://www.fao.org/fao-who-codexalimentarius/resources/circular-letters/en/?y=2018 (accessed on 1 February 2024).
  80. Zhang, J.; Zhang, W.; Zhang, Y.; Huang, M.; Sun, B. Effects of food types, frying frequency, and frying temperature on 3-monochloropropane-1,2-diol esters and glycidyl esters content in palm oil during frying. Foods 2021, 10, 2266. [Google Scholar] [CrossRef] [PubMed]
  81. Weisshaar, R.; Perz, R. Fatty acid esters of glycidol in refined fats and oils. Eur. J. Lipid Sci. Technol. 2010, 112, 158–165. [Google Scholar] [CrossRef]
  82. Taghizadeh, S.F.; Naseri, M.; Ahmadpourmir, H.; Azizi, M.; Rezaee, R.; Karimi, G. Determination of 3 monochloropropane-1,2-diol (3-MCPD) and 1,3-dichloropropan-2-ol (1,3-DCP) levels in edible vegetable oils: A health risk assessment for Iranian consumers. Microchem. J. 2023, 192, 108946. [Google Scholar] [CrossRef]
  83. Yamazaki, K.; Ogiso, M.; Isagawa, S.; Urushiyama, T.; Ukena, T.; Kibune, N. A new, direct analytical method using LC-MS/MS for fatty acid esters of 3-chloro-1, 2-propanediol (3-MCPD esters) in edible oils. Food Addit. Contam. Part A 2013, 30, 52–68. [Google Scholar] [CrossRef] [PubMed]
  84. Aniołowska, M.; Kita, A. The effect of frying on glycidyl esters content in palm oil. Food Chem. 2016, 203, 95–103. [Google Scholar] [CrossRef]
  85. Hammouda, I.B.; Zribi, A.; Ben Mansour, A.; Matthäus, B.; Bouaziz, M. Effect of deep-frying on 3-MCPD esters and glycidyl esters contents and quality control of refined olive pomace oil blended with refined palm oil. Eur. Food Res. Technol. 2017, 243, 1219–1227. [Google Scholar] [CrossRef]
  86. Hamlet, A.; Sadd, P.A.; Gray, D.A. Generation of monochloropropanediols (MCPDs) in model dough systems. 1. Leavened doughs. J. Agric. Food Chem. 2004, 52, 2059–2066. [Google Scholar] [CrossRef]
  87. Dolezal, M.; Chaloupská, M.; Divinová, V.; Svejkovská, B.; Velišek, J. Occurrence of 3-chloropropane-1,2-diol and its esters in coffee. Eur. Food Res. Tech. 2005, 221, 221–225. [Google Scholar] [CrossRef]
  88. Stauff, A.; Schneider, E.; Heckel, F. 2-MCPD, 3-MCPD and fatty acid esters of 2-MCPD, 3-MCPD and glycidol in fine bakery wares. Eur. Food Res. Technol. 2020, 246, 1945–1953. [Google Scholar] [CrossRef]
  89. Mikulíková, R.; Svoboda, Z.; Benešová, K.; Běláková, S. 3-MCPD process contaminant in malt. Kvasny Prum. 2018, 64, 6–9. [Google Scholar] [CrossRef]
  90. Ilko, V.; Zelinkova, Z.; Doležal, M.; Velíšek, J. 3-chloropropane-1,2-diol fatty acid esters in potato products. Czech J. Food Sci. 2011, 29, 411–419. [Google Scholar] [CrossRef]
  91. Leigh, J.; MacMahon, S. Occurrence of 3-MCPD and glycidyl esters in commercial infant formulas in the United States. Food Addit. Contam. Part A 2017, 34, 356–370. [Google Scholar] [CrossRef] [PubMed]
  92. Azmi, N.N.A.; Tan, T.C.; Ang, M.Y.; Leong, Y.H. Occurrence and risk assessment of 3-monochloropropanediols esters (3-MCPDE), 2-monochloropropanediol esters (2-MCPDE), and glycidyl esters (GE) in commercial infant formula samples from Malaysia. Food Addit. Contam. Part A 2023, 40, 212–221. [Google Scholar] [CrossRef] [PubMed]
  93. Wong, S.F.; Lee, B.Q.; Low, K.H.; Jenatabadi, H.S.; Bt Wan Mohamed Radzi, C.W.J.; Khor, S.M. Estimation of the dietary intake and risk assessment of food carcinogens (3-MCPD and 1,3-DCP) in soy sauces by Monte Carlo simulation. Food Chem. 2020, 311, 126033. [Google Scholar] [CrossRef]
  94. Arisseto, A.P.; Silva, W.C.; Scaranelo, G.R.; Vicente, E. 3-MCPD and glycidyl esters in infant formulas from the Brazilian market: Occurrence and risk assessment. Food Control 2017, 77, 76–81. [Google Scholar] [CrossRef]
  95. Beekman, J.K.; Grassi, K.; MacMahon, S. Updated occurrence of 3-monochloropropane-1,2-diol esters (3-MCPD) and glycidyl esters in infant formulas purchased in the United States between 2017 and 2019. Food Addit. Contam. Part A 2020, 37, 374–390. [Google Scholar] [CrossRef]
  96. Fu, W.S.; Zhao, Y.; Zhang, G.; Zhang, L.; Li, J.G.; Tang, C.D.; Miao, H.; Ma, J.B.; Zhang, Q.; Wu, Y.N. Occurrence of chloropropanols in soy sauce and other foods in China between 2002 and 2004. Food Addit. Contam. 2007, 24, 812–819. [Google Scholar] [CrossRef]
  97. Breitling-Utzmann, C.M.; Köbler, H.; Harbolzheimer, D.; Maier, A. 3-MCPD: Occurrence in bread crust and various food groups as well as formation in toast. Dtsch. Lebensm. Rundsch 2003, 99, 280–285. [Google Scholar]
  98. Zelinkova, Z.; Doležal, M.; Velíšek, J. Occurrence of 3-chloropropane-1,2-diol fatty acid esters in infant and baby foods. Eur. Food Res. Technol. 2009, 228, 571–578. [Google Scholar] [CrossRef]
  99. Divinova, V.; Doležal, M.; Velíšek, J. Free and bound 3-chloropropane-1,2-diol in coffee surrogates and malts. Czech J. Food Sci. 2007, 25, 39–47. [Google Scholar] [CrossRef]
  100. Da Costa, D.S.; Albuquerque, T.G.; Costa, H.S.; Bragotto, A.P.A. Thermal contaminants in coffee induced by roasting: A review. Int. J. Environ. Res. Public Health 2023, 20, 5586. [Google Scholar] [CrossRef] [PubMed]
  101. Jira, W. 3-monochloropropane-1,2-diol (3-MCPD) in smoked meat products: Investigation of contents and estimation of the uptake by the consumption of smoked meat products. Fleischwirtschaft 2013, 90, 115–118. [Google Scholar]
  102. Zastrow, L.; Albert, C.; Speer, K.; Schwind, K.-H.; Jira, W. Formation of pyrolysis-affected PAHs, oxygenated PAHs, and MCPDs in home smoked meat. LWT 2023, 184, 114971. [Google Scholar] [CrossRef]
  103. Kuntzer, J.; Weisshaar, R. The smoking process: A potent source of 3-chloropropane-1,2-diol (3-MCPD) in meat products. Dtsch. Lebensm. Rundsch 2006, 102, 397–400. [Google Scholar]
  104. Ostermeyer, U.; Merkle, S.; Karl, H. Free and bound MCPD and glycidyl esters in smoked and thermally treated fishery products of the German market. Eur. Food Res. Technol. 2021, 247, 1757–1769. [Google Scholar] [CrossRef]
  105. Goh, K.M.; Wong, Y.H.; Tan, C.P.; Nyam, K.L. A summary of 2-, 3-MCPD esters and glycidyl ester occurrence during frying and baking processes. Curr. Res. Food Sci. 2021, 4, 460–469. [Google Scholar] [CrossRef]
  106. Schuhmacher, R.; Nurmi-Legat, J.; Oberhauser, A.; Kainz, M.; Krska, R. A rapid and sensitive GC-MS method for determination of 1,3-dichloro-2-propanol in water. Anal. Bioanal. Chem. 2005, 382, 366–371. [Google Scholar] [CrossRef]
  107. Gonzalez, P.; Racamonde, I.; Carro, A.M.; Lorenzo, R.A. Combined solid-phase extraction and gas chromatography–mass spectrometry used for determination of chloropropanols in water. J. Sep. Sci. 2011, 34, 2697–2704. [Google Scholar] [CrossRef]
  108. Crews, C.; Chiodini, A.; Granvogl, M.; Hamlet, C.; Hrnčiřík, K.; Kuhlmann, J.; Lampen, A.; Scholz, G.; Weisshaar, R.; Wenzl, T.; et al. Analytical approaches for MCPD esters and glycidyl esters in food and biological samples: A review and future perspectives. Food Addit. Contam. Part A 2013, 30, 11–45. [Google Scholar] [CrossRef]
  109. Racamonde, I.; Gonzalez, P.; Lorenzo, R.A.; Carro, A.M. Determination of chloropropanols in foods by one-step extraction and derivatization using pressurized liquid extraction and gas chromatography-mass spectrometry. J. Chromatogr. A 2011, 1218, 6878–6883. [Google Scholar] [CrossRef]
  110. Zhong, H.-N.; Zeng, Y.; Zhu, L.; Pan, J.-J.; Wu, S.-L.; Li, D.; Dong, B.; Li, H.-K.; Wang, X.-H.; Zhang, H.; et al. The Occurrence of mono/di-chloropropanol contaminants in food contact papers and their potential health risk. Food Packag. Shelf Life 2022, 34, 101002. [Google Scholar] [CrossRef]
  111. Ramli, N.A.S.; Roslan, N.A.; Abdullah, F.; Ghazali, R.; Abd Razak, R.A.; Tarmizi, A.H.A.; Bilal, B. Analytical method for the determination of 2-and 3-monochloropropanediol esters and glycidyl ester in palm-based fatty acids by GC-MS. Food Control 2023, 151, 109824. [Google Scholar] [CrossRef]
  112. Li, Y.; Li, Y.; Zhang, N.; Wen, S.; Li, Q.; Gao, Y.; Yu, X. Methods, principles, challenges, and perspectives of determining chloropropanols and their esters. Crit. Rev. Food Sci. Nutr. 2024, 64, 1632–1652. [Google Scholar] [CrossRef]
  113. AOCS (American Oil Chemists’ Society). AOCS Official Method Cd 29a–13: 2- and 3-MCPD Fatty Acid Esters and Glycidol Fatty Acid Esters in Edible Oils and Fats by Acid Transesterification. 2013. Available online: https://myaccount.aocs.org/PersonifyEbusiness/Store/Product-Details/productId/118272 (accessed on 20 March 2024).
  114. Ermacora, A.; Hrncirik, K. A novel method for simultaneous monitoring of 2-MCPD, 3-MCPD and glycidyl esters in oils and fats. J. Am. Oil Chem. Soc. 2013, 90, 1–8. [Google Scholar] [CrossRef]
  115. Cao, X.; Song, G.; Gao, Y.; Zhao, J.; Zhang, M.; Wu, W.; Hu, Y. A Novel derivatization method coupled with GC–MS for the simultaneous determination of chloropropanols. Chromatographia 2009, 70, 661–664. [Google Scholar] [CrossRef]
  116. Hamlet, C.G.; Sutton, P.G. Determination of the chloropropanols, 3-chloro-1, 2-propandiol and 2-chloro-1, 3-propandiol, in hydrolysed vegetable proteins and seasonings by gas chromatography/ion trap tandem mass spectrometry. Rapid Commun. Mass Spectr. 1997, 11, 1417–1424. [Google Scholar] [CrossRef]
  117. Wang, L.; Ying, Y.; Hu, Z.; Wang, T.; Shen, X.; Wu, P. Simultaneous determination of 2-and 3- MCPD esters in infant formula milk powder by solid-phase extraction and GC-MS analysis. J. AOAC Int. 2016, 99, 786–791. [Google Scholar] [CrossRef]
  118. Oey, S.B.; van der Fels-Klerx, H.J.; Fogliano, V.; van Leeuwen, S.P. Chemical refining methods effectively mitigate 2-MCPD esters, 3-MCPD esters, and glycidyl esters formation in refined vegetable oils. Food Res. Int. 2022, 156, 111137. [Google Scholar] [CrossRef]
  119. Tran, S.C.; Nguyen, N.H.; Vu, T.N.; Bui, T.C.; Phung, L.C.; Tran, T.T.; Thai, T.N.H. Risk assessment of 3-MCPD esters and glycidyl esters from the formulas for infants and young children up to 36 months of age. Food Addit. Contam. Part A 2023, 40, 723–732. [Google Scholar] [CrossRef] [PubMed]
  120. Chai, Q.; Zhang, X.; Karangwa, E.; Dai, Q.; Xia, S.; Yu, J.; Gao, Y. Direct determination of 3-chloro-1, 2-propanediol esters in beef flavoring products by ultra-performance liquid chromatography tandem quadrupole mass spectrometry. RSC Adv. 2016, 6, 113576–113582. [Google Scholar] [CrossRef]
  121. Sun, X.; Zhang, L.; Zhang, H.; Qian, H.; Zhang, Y.; Tang, L.; Li, Z. Development and application of 3-chloro-1, 2-propandiol electrochemical sensor based on a polyaminothiophenol modified molecularly imprinted film. J. Agric. Food Chem. 2014, 62, 4552–4557. [Google Scholar] [CrossRef]
  122. Yuan, Y.; Wang, J.; Ni, X.; Cao, Y. A biosensor based on hemoglobin immobilized with magnetic molecularly imprinted nanoparticles and modified on a magnetic electrode for direct electrochemical determination of 3-chloro-1, 2-propandiol. J. Electroanal. Chem. 2019, 834, 233–240. [Google Scholar] [CrossRef]
  123. Fang, M.; Zhou, L.; Zhang, H.; Liu, L.; Gong, Z.Y. A molecularly imprinted polymers/carbon dots-grafted paper sensor for 3-monochloropropane-1,2-diol determination. Food Chem. 2019, 274, 156–161. [Google Scholar] [CrossRef]
  124. Silva, W.C.; Santiago, J.K.; Capristo, M.F.; Ferrari, R.A.; Vicente, E.; Sampaio, K.A.; Arisseto, A.P. Washing bleached palm oil to reduce monochloropropanediols and glycidyl esters. Food Addit. Contam. Part A 2019, 36, 244–253. [Google Scholar] [CrossRef]
  125. Ramli, M.R.; Siew, W.L.; Ibrahim, N.A.; Kuntom, A.; Razak, R.A.A. Other factors to consider in the formation of chloropropandiol fatty esters in oil processes. Food Addit. Contam. Part A 2015, 32, 817–824. [Google Scholar] [CrossRef]
  126. CXC 79-2019; Code of Practice for the Reduction of 3-Monochloropropane-1,2-diol Esters (3-MCPDEs) and Glycidyl Esters (GEs) in Refined Oils and Food Products Made with Refined Oils. Codex Alimentarius International Food Standards: Rome, Italy, 2019.
  127. Strijowski, U.; Heinz, V.; Franke, K. Removal of 3-MCPD esters and related substances after refining by adsorbent material. Eur. J. Lipid Sci. Technol. 2011, 113, 387–392. [Google Scholar] [CrossRef]
  128. Bornscheuer, U.T.; Hesseler, M. Enzymatic removal of 3-monochloro-1,2-propanediol (3-MCPD) and its esters from oils. Eur. J. Lipid Sci. Technol. 2010, 112, 552–556. [Google Scholar] [CrossRef]
  129. Liu, S.; Shen, M.; Xie, J.; Liu, B.; Li, C. Effects of endogenous antioxidants in camellia oil on the formation of 2-monochloropropane-1, 3-diol esters and 3-monochloropropane-1,2-diol esters during thermal processing. Foods 2024, 13, 261. [Google Scholar] [CrossRef] [PubMed]
  130. Tobón, J.F.O.; Meireles, M.A.A. Recent applications of pressurized fluid extraction: Curcuminoids extraction with pressurized liquids. Food Public Health 2013, 3, 289–303. [Google Scholar] [CrossRef]
  131. Chen, G.W.; Wang, G.; Zhu, C.J.; Jiang, X.W.; Sun, J.X.; Tian, L.M.; Bai, W. Effects of cyanidin-3-O-glucoside on 3-chloro-1,2-propanediol induced intestinal microbiota dysbiosis in rats. Food Chem. Toxicol. 2019, 133, 110767. [Google Scholar] [CrossRef]
  132. Tian, L.; Tan, Y.; Chen, G.; Wang, G.; Sun, J.; Ou, S.; Chen, W.; Bai, B. Metabolism of anthocyanins and consequent effects on the gut microbiota. Crit. Rev. Food Sci. Nutr. 2019, 59, 982–991. [Google Scholar] [CrossRef]
  133. Xie, J.; Luo, C.; Yang, X.; Ren, Y.; Zhang, X.; Chen, H.; Zhao, Y.; Wu, F. Study on wild medicinal plant resources and their applied ethnology in multiethnic areas of the Gansu-Ningxia-Inner Mongolia intersection zone. J. Ethnobiol. Ethnomed. 2023, 19, 18. [Google Scholar] [CrossRef]
  134. Mou, Y.; Sun, L.; Geng, Y.; Xie, Y.; Chen, F.; Xiao, J.; Hu, X.; Ji, J.; Ma, L. Chloropropanols and their esters in foods: Exposure, formation and mitigation strategies. Food Chem. Adv. 2023, 3, 100446. [Google Scholar] [CrossRef]
  135. Rondanelli, M.; Nichetti, M.; Martin, V.; Barrile, G.C.; Riva, A.; Petrangolini, G.; Gasparri, C.; Perna, S.; Giacosa, A. Phytoextracts for human health from raw and roasted hazelnuts and from hazelnut skin and oil: A narrative review. Nutrients 2023, 15, 2421. [Google Scholar] [CrossRef]
  136. ANZFA (Australia New Zealand Food Authority). Draft Assessment Proposal P243: Maximum Limit for Chloropropanols in Soy and Oyster Sauces. 2001. Available online: https://www.foodstandards.gov.au/sites/default/files/food-standards-code/proposals/Documents/P243%20DraftFAR.pdf (accessed on 3 May 2024).
  137. Health Canada. Health Canada’s Maximum Levels for Chemical Contaminants in Foods. 2018. Available online: https://www.canada.ca/en/health-canada/services/food-nutrition/food-safety/chemical-contaminants/maximum-levels-chemical-contaminants-foods.html/ (accessed on 3 May 2024).
  138. GB 2762-2017; National Food Safety Standard, Maximum Levels of Contaminants in Foods. CFDA (China Food and Drug Administration): Beijing, China, 2017.
  139. European Commission. Commission Regulation (EU) 2023/915 on maximum levels for certain contaminants in food and repealing Regulation (EC) No 1881/2006. Off. J. EU 2023, L 119, 103. [Google Scholar]
  140. Malaysia Food Regulations. Maximum Permitted Proportion of 3-Monochloropropane-1,2-diol (3-MCPD) in Specific Food. 1985. Available online: https://faolex.fao.org/docs/pdf/mal27305.pdf (accessed on 11 May 2024).
  141. MFDS (Ministry of Food and Drug Safety). Risk Assessment of 3-MCPD and 1,3-DCP in Foods (11-1471057-01); Technical Report; National Institute of Food and Drug Safety Evaluation: Cheongju-si, Republic of Korea, 2019; pp. 47–61.
  142. TFC (Turkish Food Codex). Bulaşanlar Yönetmeliği. Resmî Gazete Tarihi: 5 November 2023, Resmî Gazete Sayısı: 32360. Available online: https://www.resmigazete.gov.tr/eskiler/2023/11/20231105-1.htm (accessed on 25 March 2024).
  143. FDA (Food and Drug Administration). Guidance Levels for 3-MCPD (3-chloro-1,2-propanediol) in Acid-Hydrolyzed Protein and Asian-Style Sauces. 2008. Available online: https://www.fda.gov/media/71760/download (accessed on 27 March 2024).
  144. TFDA (Taiwan Food and Drug Administration). Regulations for 3-MCPD and GEs in Infant Formulas. 2021. Available online: https://www.fda.gov.tw/ENG/site.aspx?sid=10172#arrow-up-a (accessed on 25 March 2024).
  145. European Commission. Commission Regulation (EC) No 1881/2006 of 19 December 2006 setting maximum levels for certain contaminants in foodstuffs. Off. J. EU 2006, L364, 5–24. [Google Scholar]
  146. European Commission. Commission Regulation (EU) 2020/1322 of 23 September 2020 amending Regulation (EC) No 1881/2006 as regards maximum levels of 3-monochloropropanediol (3-MCPD), 3-MCPD fatty acid esters and glycidyl fatty acid esters in certain foods. Off. J. EU 2020, L310, 2–5. [Google Scholar]
  147. AGRINFO. Maximum Levels for 3-MCPD in Infant Formulae. The Latest on EU Agri-Food Policies Impacting Low-Income & Middle-Income Countries. 2024. Available online: https://www.google.com/url?sa=i&url=https%3A%2F%2Fagrinfo.eu%2Fbook-of-reports%2Fmaximum-levels-for-3-mcpd-in-infant-formulae%2F&psig=AOvVaw3HZ2mSDqEd6H5EqOCFWH7e&ust=1715940653836000&source=images&cd=vfe&opi=89978449&ved=0CAgQr5oMahcKEwjo_raa95GGAxUAAAAAHQAAAAAQBA (accessed on 12 May 2024).
  148. FSANZ (Food Standards Australia New Zealand). Chloropropanols in Food. Technical Report Series No 15. 2003. Available online: https://www.foodstandards.gov.au/sites/default/files/publications/Documents/Chloropropanol%20Report%20(no%20appendices)%20-%2011%20Sep%2003b-2.pdf (accessed on 28 March 2024).
Figure 1. Chemical structures of chloropropanols and their esters.
Figure 1. Chemical structures of chloropropanols and their esters.
Foods 13 02876 g001
Figure 2. Glycidol and fatty acid ester of glycidol (R = alk(en)yl).
Figure 2. Glycidol and fatty acid ester of glycidol (R = alk(en)yl).
Foods 13 02876 g002
Figure 3. Possible formation mechanism of GE under thermal conditions.
Figure 3. Possible formation mechanism of GE under thermal conditions.
Foods 13 02876 g003
Figure 4. Conversion mechanism among 2-MCPD, 3-MCPD, and GE.
Figure 4. Conversion mechanism among 2-MCPD, 3-MCPD, and GE.
Foods 13 02876 g004
Figure 5. The numbers of RASFF notifications on chloropropanols, GE, and glycidol during the years 2020–2023 by product category.
Figure 5. The numbers of RASFF notifications on chloropropanols, GE, and glycidol during the years 2020–2023 by product category.
Foods 13 02876 g005
Table 1. Oral toxic effects and the LD50 values of chloropropanols and glycidol.
Table 1. Oral toxic effects and the LD50 values of chloropropanols and glycidol.
SubstanceOral Toxic EffectsLD50, mg/kg b.w. (Species)References
3-MCPDMutagenic effects on sperm
Male infertility
191 (mouse)
152 (rat)
[38]
[39]
2-MCPDToxic by ingestion50–60 (rat)[1]
1,3-DCPFetotoxic effects on reproductivity
Tumorogenic effects on liver, kidney, ureter, and bladder
100 (mouse)
110 (rat)
[40]
[40]
2,3-DCPNA90 (rat)[40]
GlycidolDNA damage in ovary and sperm
Effects on embryo/fetus
Male infertility
Tumorogenic effects on gastrointestine, liver, skin, endocrine, and brain
431 (mouse)
420 (rat)
[41]
[41]
Table 2. Occurrence and levels of 3-MCPD (from esters) in vegetable oils.
Table 2. Occurrence and levels of 3-MCPD (from esters) in vegetable oils.
ProductCountryNo. of Samples (n)Incidence (%)Range (µg/kg)MethodReferences
Camellia oilChina5NA988–2586DGF method C-VI 18[72]
Camellia oil (crude)China5NA250–555DGF method C-VI 18[72]
Canola oilMalaysiaNANA600BfR Method 008[73]
Maize germ oilChina12NA219–1826DGF method C–VI 18[72]
Maize oilMalaysiaNANA250–300BfR Method 008[73]
Olive oilChina1110034–1970ISO-18363-4[71]
Olive oilBrazil1392250–3900AOCS Cd 29a-13[74]
Olive oil (bland)Brazil1735250–620AOCS Cd 29a-13[74]
Olive oil (pomace)MalaysiaNANA1650BfR Method 008[73]
Olive oil (virgin)MalaysiaNANA350BfR Method 008[73]
Olive oil (virgin)Brazil4617250–1240AOCS Cd 29a-13[74]
Palm oil (bleached)Malaysia5650250–1800BfR Method 008[73]
Palm oil (crude)Malaysia10520250–900BfR Method 008[73]
Palm oil (refined)MalaysiaNA99250–5800BfR Method 008[73]
Palm oilChina18100270–8390ISO-18363-4[71]
Peanut oil (refined) China15NA450–1187DGF method C-VI 18[72]
Peanut/sesame oilMalaysiaNANA2450BfR Method 008[73]
Peanut oilChina14100190–3720ISO-18363-4[71]
Rapeseed oil (crude)China9NA250–438DGF method C-VI 18[72]
Rapeseed oil (refined) China18NA226–1069DGF method C-VI 18[72]
Rapeseed oilChina1210082–6400ISO-18363-4[71]
Refined vegetable oilSingapore36NA78–9592AOCS Cd 29a-13[70]
Rice bran oilMalaysiaNANA250–300BfR Method 008[73]
Sesame oil (crude)China6NA250–356DGF method C-VI 18[72]
Sesame oil (refined) China4NA651–1344DGF method C-VI 18[72]
Sesame oilChina1810069–5860ISO-18363-4[71]
Soybean oilMalaysiaNANA<250BfR Method 008[73]
Soybean oil (crude)China7NA<250DGF method C-VI 18[72]
Soybean oil (refined) China18NA224–1090DGF method C-VI 18[72]
Sunflower oilMalaysiaNANA600BfR Method 008[73]
Sunflower oil (crude)China8NA<250DGF method C-VI 18[72]
Sunflower oil (refined)China6NA504–1044DGF method C-VI 18[72]
Sunflower oilChina13100140–2580ISO-18363-4[71]
Vegetable oil (unrefined)Singapore24NA<30–1172AOCS Cd 29a-13[70]
NA: Not available.
Table 3. Occurrence and levels of 2-MCPD (from esters) in vegetable oils and fats.
Table 3. Occurrence and levels of 2-MCPD (from esters) in vegetable oils and fats.
ProductCountryNo. of Samples (n)Incidence (%)Range (µg/kg)MethodReferences
Edible oilsPoland2711180–230AOCS Cd 29b-13[75]
MargarinesPoland5100630–1700AOCS Cd 29b-13[75]
Palm oilGermany20NA200–5900AOCS Cd 29b-13[76]
Palm oilChina18100150–6950ISO-18363-4[71]
Rapeseed oilGermany5NA<LOQ–300AOCS Cd 29b-13[76]
Rapeseed oilChina1210071–4440ISO-18363-4[71]
Soybean oilGermany5NA<LOQ–300AOCS Cd 29b-13[76]
Soybean oilChina21100336–1200ISO-18363-4[71]
Sunflower oilGermany5NA<LOQ–300AOCS Cd 29b-13[76]
Sunflower oilChina1310028–1700ISO-18363-4[71]
Olive oilBrazil1385200–2010AOCS Cd 29a-13[74]
Olive oil (bland)Brazil1712200–280AOCS Cd 29a-13[74]
Olive oil (virgin)Brazil4613200–2160AOCS Cd 29a-13[74]
Olive oilChina1182800–1010ISO-18363-4[71]
NA: Not available. LOQ: Limit of quantification.
Table 4. Occurrence and levels of GE in vegetable oils.
Table 4. Occurrence and levels of GE in vegetable oils.
ProductCountryNo. of Samples (n)Incidence (%)Range (µg/kg)MethodReferences
Palm oilGermany20NA300–18,000AOCS Cd 29b-13[76]
Rapeseed oilGermany5NA<100–300AOCS Cd 29b-13[76]
Olive oilBrazil13100200–1910AOCS Cd 29a-13[74]
Olive oil (bland)Brazil17100200–1910AOCS Cd 29a-13[74]
Olive oil (virgin)Brazil4635200–2160AOCS Cd 29a-13[74]
Soybean oilGermany5NA<100–600AOCS Cd 29b-13[76]
Sunflower oilGermany5NA<100–400AOCS Cd 29b-13[76]
Vegetable oil (refined)Singapore36NA155–6204AOCS Cd 29a-13[70]
Vegetable oil (refined)Singapore24NA<30–1325AOCS Cd 29a-13[70]
NA: Not available.
Table 5. Occurrence and levels of free and bound 3-MCPD in various foodstuffs.
Table 5. Occurrence and levels of free and bound 3-MCPD in various foodstuffs.
SubstanceProductCountryNo. of Samples (n)Incidence (%)Range (µg/kg)References
Free 3-MCPDCereal productsCzech Rep.NANA1–477[86]
CoffeeCzech Rep.157310–19[87]
Fine bakeryGermany94275–265[88]
MaltCzech Rep.223210–95[89]
Bound 3-MCPDCereal productsUKNANA3–2916[86]
CoffeeCzech Rep.15806–390[87]
Fine bakeryGermany94NA18–1365[88]
Fried French friesCzech Rep.16NA100–258[90]
Infant formulaUnited States98NA21–920[91]
Infant formulaMalaysia161002–244[92]
Potato crispsCzech Rep.56NA229–1008[90]
Total 3-MCPDCereal and cereal productsHong Kong57544–23[13]
Chicken seasoning cubesMalaysia610090[93]
Dairy productsHong Kong120<LOD[13]
Egg and productsHong Kong180<LOD[13]
Fish and shellfish productsHong Kong66223–33[13]
FruitsHong Kong210<LOD[13]
Infant formulaBrazil4037<LOD–600[94]
Infant formulaUnited States222NA13–950[95]
Meat and poultry productsHong Kong87494–32[13]
SnacksHong Kong24386–66[13]
Soy-based saucesMalaysia4354<20–122[93]
Soy saucesChina629895–189,000[96]
Vegetable productsHong Kong390<LOD[13]
NA: Not available. LOD: Limit of detection.
Table 6. Other chloropropanol derivates than 3-MCPD in food.
Table 6. Other chloropropanol derivates than 3-MCPD in food.
SubstanceProductNo. of Samples (n)Incidence (%)Range (µg/kg)References
Bound 2-MCPDCerealsNANA1–853[86]
Fine bakery products94NA<3–624[88]
Infant formula16812–22[86]
Soy sauces3454810–20300[96]
1,3-DCPFish products66143–6[13]
Meat and poultry products8775–10[13]
Soy sauces28220100–1400[12]
Soy sauces345194–8260[96]
Soy-based sauces4354<LOD–25[93]
Water>300 <100[106]
Water24NA6–122[107]
2,3-DCPSoy sauce711013–28[12]
Soy sauces34543–500[96]
GEInfant formula 42<LOD–750[94]
Infant formula98NA5–400[91]
Infant formula222NA19–370[95]
Infant formula16382–69[92]
GlycidolFine bakery products94NA<3–128[88]
NA: Not available. LOD: Limit of detection.
Table 7. The ML of 3-MCPD in soy sauce in various countries.
Table 7. The ML of 3-MCPD in soy sauce in various countries.
CountryML (µg/kg)References
Australia200[136]
Canada1000[137]
China400[138]
European 20[139]
Malaysia20[140]
New Zealand200[136]
South Korea300[141]
Turkey20[142]
United States1000[143]
Japan1000[48]
Taiwan1000[144]
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Ozluk, G.; González-Curbelo, M.Á.; Kabak, B. Chloropropanols and Their Esters in Food: An Updated Review. Foods 2024, 13, 2876. https://doi.org/10.3390/foods13182876

AMA Style

Ozluk G, González-Curbelo MÁ, Kabak B. Chloropropanols and Their Esters in Food: An Updated Review. Foods. 2024; 13(18):2876. https://doi.org/10.3390/foods13182876

Chicago/Turabian Style

Ozluk, Gizem, Miguel Ángel González-Curbelo, and Bulent Kabak. 2024. "Chloropropanols and Their Esters in Food: An Updated Review" Foods 13, no. 18: 2876. https://doi.org/10.3390/foods13182876

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

Article Metrics

Article metric data becomes available approximately 24 hours after publication online.
Back to TopTop