Next Article in Journal
Physicochemical and Sensory Evaluation of Spreads Derived from Fruit Processing By-Products
Previous Article in Journal
Research Progress on Non-Destructive Testing Technology and Equipment for Poultry Eggshell Quality
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Mechanistic Insights into Vegetable Color Stability: Discoloration Pathways and Emerging Protective Strategies

Agricultural Product Processing and Storage Lab, School of Food and Biological Engineering, Jiangsu University, Zhenjiang 212013, China
*
Author to whom correspondence should be addressed.
Foods 2025, 14(13), 2222; https://doi.org/10.3390/foods14132222
Submission received: 29 May 2025 / Revised: 21 June 2025 / Accepted: 23 June 2025 / Published: 24 June 2025

Abstract

During processing and storage, vegetables often experience undesirable color changes, including fading, lightening, or yellowing and softening, due to browning (enzymatic and non-enzymatic) and chlorophyll degradation. These changes diminish commercial and nutritional value. Therefore, it is necessary to maintain vegetable color and improve the quality of vegetable-based dishes. To address these issues, it is a scientific and practical necessity to summarize and discuss existing strategies and innovative techniques. This review first highlights the mechanisms of vegetable browning. This review then provides a comprehensive overview of recent advances in methods for color preservation, focusing on underlying mechanisms and techniques for inhibiting color changes from physical, chemical, and biological perspectives. A review of innovative technologies suggests that effective color preservation in vegetables is achieved by inhibiting the conditions that lead to three unfavorable color change reactions: enzymatic browning, non-enzymatic browning, and chlorophyll degradation. Current research frequently employs combined approaches that integrate two or more techniques to mitigate these adverse color changes. Moreover, most of these methods could simultaneously inhibit the three reaction processes. Future research directions are proposed for in-depth investigations into the molecular mechanisms of color changes in vegetables and the impact of treatments on the nutritional value.

1. Introduction

Each year, over 30% of harvested vegetables are discarded globally due to color degradation, resulting in economic losses exceeding USD 50 billion and diminishing nutritional quality [1]. Color deterioration not only reduces consumer appeal but also serves as a visible marker of underlying biochemical spoilage that may compromise food safety [2,3]. As color is the first visual cue consumers use to assess freshness and quality, discoloration severely limits the commercial value and shelf-life of vegetable products.
Vegetables especially play a vital role in such diets due to their rich nutritional profiles and health-promoting properties [4]. According to the Food and Agriculture Organization of the United Nations (FAO), vegetables supply more than 90% of the daily required intake of vitamin C and over 65% of vitamin A for the human body [5]. Additionally, vegetables are rich in phenolic compounds, dietary fiber, organic acids, and bioactive sulfur-containing compounds such as dipropylene and methyl sulfide compounds, which contribute to enhance immune function and the prevention of chronic diseases [6]. Consequently, ensuring the quality and availability of vegetables is critical to addressing global nutritional deficiencies and advancing public health.
However, vegetables are highly perishable and especially vulnerable to spoilage during post-harvest handling, storage, and distribution [7]. Among all quality attributes, color is the most rapidly affected and strongly influences consumer purchasing decisions [8]. Discoloration—particularly browning and chlorophyll degradation—has been identified as a major cause of rejection in the global vegetable market, leading to substantial economic losses [9,10].
Although various reviews have examined enzymatic browning and its inhibition strategies, there is a lack of comprehensive analyses that integrate all forms of discoloration—including non-enzymatic browning and chlorophyll degradation—with emerging control methods. Furthermore, existing literature tends to focus either on degradation mechanisms or on individual preservation strategies, without bridging the two perspectives. In this review, we address this gap by systematically connecting the molecular mechanisms behind vegetable discoloration with the most recent advances in color stabilization. We synthesize studies published between 2020 and 2025, focusing on three major intervention approaches: physical, chemical, and natural anti-browning strategies. By presenting an integrated overview, this review offers an actionable framework for improving vegetable shelf-life, reducing waste, and promoting the commercialization of visually appealing, nutritious plant-based products. The content of the review structure is shown in Figure 1.
Relevant studies were retrieved from the Web of Science and PubMed databases (2020–2025) using keywords such as ‘vegetable color stability’, ‘pigment degradation’, and ‘protective interventions’. We applied inclusion criteria of English-language original research articles and comprehensive reviews that reported quantitative data or mechanistic insights into vegetable pigment stability. Exclusion criteria included conference abstracts, book chapters, patents, non-English studies, and irrelevant topics. A total of 512 records were initially identified, of which 189 met the eligibility criteria and were included in this review.

2. Mechanism of Vegetable Color Change

There are two main factors for color changes in vegetable-based products: browning and chlorophyll degradation. Browning is categorized into enzymatic and non-enzymatic types [9]. Among them, enzymatic browning significantly influences the color change in vegetable-based products.

2.1. Enzymatic Browning

2.1.1. Mechanism of Enzymatic Browning

Enzymatic browning is a complex biochemical process that is usually described by the theory of the regionalized distribution of phenolic compounds in plant tissues and the enzymes that induce their reactions [11]. According to this theory, the spatial separation of these components plays a critical role in preventing undesirable browning reactions. Phenolic compounds, the primary substrates of enzymatic browning, are typically sequestered in specific intracellular vesicles, while polyphenol oxidase (PPO), the enzyme responsible for their oxidation, is located in distinct cytoplasmic compartments [12]. This compartmentalization ensures that phenolics and PPO remain separated under normal physiological conditions, thereby inhibiting browning reactions despite the presence of oxygen, a key reactant in the process [13].
The cell is a highly organized structure with organelles surrounded by a selectively permeable film, which helps delineate the intracellular environment. In intact cells, this organization provides protection against the enzymatic browning reaction despite the presence of oxygen (a key component of enzymatic browning). However, compartmentalization is disrupted when the cell wall is physically damaged, mechanically damaged, or damaged by enzymatic breakdown. This allows phenolic compounds that were in the vesicles to come into direct contact with PPO and oxygen in the cytoplasm, triggering a series of biochemical reactions. Specifically, these reactions are the oxidation of phenolic substrates. The phenolic substrate is polymerized into larger, more complex structures during the reaction. The resulting melanoidins then lead, at a macroscopic level, to the dark brown discoloration observed in damaged or cut vegetable tissues [14].
The formation of melanoidins is often associated with the degradation of antioxidants and other essential nutrients, thereby not only affecting the visual quality of vegetables but also diminishing their nutritional value and shelf-life [15]. The rate and extent of enzymatic browning are influenced by several factors, including temperature, oxygen availability, and PPO activity, making its control and inhibition particularly challenging.

2.1.2. Conditions of Enzymatic Browning

Enzymatic browning is a chemical reaction involving several interrelated processes. As shown in Figure 2, these chemical reactions proceed sequentially leading to characteristic browning of plant tissues. First, monophenols are oxidized to o-diphenols by PPO in the presence of oxygen. This initial reaction forms colorless o-diphenol, which serves as the main substrate for further oxidation. These o-diphenols are further oxidized to form o-quinone under the catalytic action of PPO. O-quinone is a highly reactive intermediate. Immediately thereafter, o-quinones through interactions with proteins, amino acids and other cellular components. Complex brown polymers are formed which ultimately lead to visible browning of plant tissues [14].
The enzymatic browning reaction is continuous. This means that the reaction continues as long as reducing substances, such as phenolic compounds and sufficient enzyme activity, are still present. Once these substrates are depleted or the enzyme activity is reduced due to various factors, the reaction slows down and eventually stops [16]. The enzymatic browning process fundamentally depends on the presence of three critical components: oxygen, PPO, and phenolic substrates [17]. All three must be simultaneously present for browning to proceed efficiently. Consequently, effective control of these factors is key to minimizing enzymatic browning, which is particularly crucial in the post-harvest preservation of vegetables [18].
In addition to these primary factors, enzymatic browning is also affected by a number of secondary environmental conditions such as temperature, light and pressure. High temperatures can provide more kinetic energy to increase the rate of chemical reactions. On the contrary, low temperatures can slow down the activity of PPO, which in turn reduces the browning process [19]. Light exposure has been reported to enhance reactive oxygen species production, which can increase PPO activity and thus promote browning [20]. In addition, external pressure affects enzyme conformation and activity, and high pressure may inhibit PPO activity or alter its substrate specificity [21].
Thus, while oxygen, enzymes, and phenolics are necessary for enzymatic browning, external conditions such as temperature, light, and pressure are secondary factors that influence the enzymatic reaction [22]. A comprehensive understanding of the interactions between these primary and secondary factors is essential for developing effective strategies to mitigate enzymatic browning in fresh produce.

2.2. Non-Enzymatic Browning

2.2.1. Mechanism of Non-Enzymatic Browning

Non-enzymatic browning is also a key factor in the color stability, quality and shelf-life of vegetable products. Unlike enzymatic browning, non-enzymatic browning occurs through a number of different chemical reactions. These chemical reactions do not require enzymes, but still result in undesired color changes and quality deterioration. Previous research has identified several reactions that fall under the category of non-enzymatic browning, including the Maillard reaction, caramelization reaction, degradation of vitamin C, and polyphenol oxidative condensation [23]. The Maillard reaction is the most widely studied and the most important contributor to non-enzymatic browning of foods, including vegetables.
The Maillard reaction involves a complex sequence of interactions between reducing sugars and amino compounds, typically amino acids or proteins. Initially, the carbonyl group of a reducing sugar (e.g., glucose or fructose) reacts with the amino group of an amino compound to form an unstable Schiff base. This intermediate rapidly undergoes an Amadori rearrangement to yield more stable Amadori products, which are highly reactive. Under various environmental conditions, these intermediates degrade into a range of compounds, including unsaturated aldehydes with strong browning potential. Ultimately, these reactions lead to the formation of high-molecular-weight melanoidins, which are responsible for the characteristic browning observed in processed and stored foods [24,25].
Current research has focused on the extensive study and characterization of the formation of the intermediate products of the Maillard reaction, and the mechanism of the production of the final high-molecular-weight polymer is still poorly understood. Therefore, recent studies have focused on the effect of various processing conditions (e.g., temperature, moisture content, and pH value) on the formation of the final product melanoidin. As with enzymatic browning, these high-molecular-weight products not only lead to browning, but also alter the flavor, texture, and nutritional composition of the food. In addition, the Maillard reaction interacts with other non-enzymatic browning reactions (e.g., caramelization and polyphenol oxidative condensation), further complicating the overall non-enzymatic browning process. Understanding the whole process of non-enzymatic browning reactions and their underlying mechanisms, especially the formation of high-molecular-weight melanoidins, is essential for developing strategies to control and mitigate browning of vegetable products in order to improve their color stability and extend their shelf-life.

2.2.2. Conditions of Non-Enzymatic Browning

The mechanism of non-enzymatic browning is complex and involves many chemical reactions that produce a large number of intermediates. These reaction processes are influenced by a variety of factors and are difficult to control completely. Therefore, effective control and inhibition of non-enzymatic browning can only be achieved by looking at the external conditions that affect the chemical reactions, such as temperature, time, pH and the presence of specific precursor compounds. Previous study has found that the complexity of non-enzymatic browning stems from its multi-step nature. These reactions are sensitive to environmental conditions external to the reaction, which in turn significantly affects the rate and extent of browning in different foods [26].
For example, Sacchetti et al. reported that high-temperature, short-time treatments led to less non-enzymatic browning in cocoa beans compared to prolonged heating, indicating that the balance between temperature and time is crucial in controlling browning intensity. Their study also highlighted that extended cooking or overheating can accelerate the Maillard reaction and other non-enzymatic pathways, resulting in excessive pigment formation and deterioration of flavor and nutritional quality [27]. In addition, Liu et al. found that pH plays a vital role in modulating browning; lower pH values were shown to inhibit the formation of reactive intermediates, thereby reducing the generation of high-molecular-weight browning compounds. This suggests that pH adjustment during food processing or storage may be an effective strategy to suppress non-enzymatic browning in vegetables and other products [28].
In addition to processing conditions, the chemical composition of a food also plays a key role in determining the rate of non-enzymatic Browning. It has been found that ascorbic acid, fructose, and arginine can act as pre-reactive agents of non-enzymatic Browning, accelerating or altering the process of non-enzymatic Browning [29]. Different amino acids have different interactions with reducing sugars and have different effects on non-enzymatic Browning. For example, basic amino acids, such as lysine and arginine, tend to form more reactive intermediates with sugars, which can lead to faster Browning. In contrast, non-polar or neutral amino acids may produce slower Browning rates due to their chemical properties [30]. The structure of sugars also plays a role—simple sugars such as glucose and fructose are more susceptible to browning than complex carbohydrates due to their higher reactivity in Maillard-type reactions [31].
At the same time, other conditions can also affect the non-enzymatic Browning process. Guo et al. found that the presence of certain metal ions, such as calcium and magnesium, may be involved in active intermediates or enzymatic interactions of Browning, preventing the formation of unwanted pigments and thereby inhibiting the Maillard reaction [32]. Grant-Preece et al. have shown that exposure to light may speed up or slow down the rate of non-enzymatic Browning, depending on the specific wavelength and intensity of the light [33]. Shishir et al. found that modified atmosphere packaging could inhibit browning by controlling oxygen levels or incorporating light-blocking materials [34].
In summary, non-enzymatic Browning is a highly complex process that depends on a range of factors, including temperature, time, pH, and the chemical composition of the food. While certain conditions can accelerate Browning, other conditions, such as acidic pH or the presence of metal ions, can help slow it down. Understanding these factors and their interactions is critical to developing strategies to control non-enzymatic Browning.

2.3. Chlorophyll Degradation

2.3.1. Mechanism of Chlorophyll Degradation

In green vegetables, the degradation of chlorophyll is a key factor leading to color loss and significantly affects the overall quality of the product, including its visual appeal, nutritional content and consumer acceptability [35]. Chlorophyll refers to the green pigment responsible for capturing light energy during photosynthesis. It exists mainly in two forms: chlorophyll-a and chlorophyll-b. The two types of chlorophyll differ in their chemical structure. Chlorophyll-a has a methyl group (-CH3) on the II pyrrole ring, while chlorophyll-b has an aldehyde group (-CHO) in the same location on the pyrrole ring [36]. Both forms of chlorophyll play a crucial role in photosynthesis. Part of this chlorophyll-a is the main pigment in the light harvesting process, while chlorophyll-b assists in capturing light energy and transferring it to chlorophyll-a. In the process of green vegetable storage, due to temperature, light, pH value, mechanical damage and other environmental factors will cause chlorophyll degradation, which leads to adverse color reactions in vegetables.
The degradation of chlorophyll is catalyzed by multiple chemical reactions. First, the chlorophyll enzyme catalyzes the shedding of the phthalein group in chlorophyll-a, converting it into demethyl-chlorophyll-a and chlorophyll-alcohol. This makes the pigment less stable and more prone to further degradation. Subsequently, de-magnesium chelatase catalyzes the removal of magnesium ions in chlorophyll-a to produce de-magnesium chlorophyll-a, which is a key step in chlorophyll degradation. The stability of pheophyll-a is significantly lower than that of chlorophyll-A, making it more susceptible to oxidation and further transformation. Then, pheophyll-a is further enzymolized by pheophyllase oxygenase and red chlorophyll metabolite reductase, resulting in the formation of fluorescent chlorophyll metabolites. As a result, green pigments are gradually replaced by other decomposition products, producing brown or yellow color changes [37,38].

2.3.2. Conditions of Chlorophyll Degradation

Chlorophyll enzymes play a central role in the breakdown of chlorophyll molecules during post-harvest storage and food processing. However, studies have shown that the relationship between chlorophyll enzyme activity and chlorophyll degradation rate is not linear, suggesting that various factors may affect enzyme activity in complex ways [39]. Therefore, any factor that directly or indirectly affects the activity of chlorophyll enzymes will have an impact on the chlorophyll degradation rate.
For instance, acidic environments have been reported to inhibit chlorophyllase activity by altering its structure and reducing its catalytic efficiency, thereby slowing chlorophyll breakdown [40]. Therefore, protecting the color of green vegetables during processing and controlling pH is a key step. Especially in curing or pickling operations, acidifiers are often added to preserve color and extend shelf-life. In addition, light is another key factor affecting chlorophyll degradation. Zhang et al. showed that light can induce ROS production, promote oxidation reaction and accelerate the decomposition of chlorophyll in plant tissues [41]. Reducing light during storage or processing is a key strategy to maintain chlorophyll content and color stability in vegetables. Therefore, it is often recommended to use light-blocking packaging materials or store in dark conditions to extend the shelf-life of green vegetables.
Temperature also significantly impacts chlorophyll degradation. Elevated temperatures can enhance enzyme activity, including that of chlorophyll-degrading enzymes, thereby accelerating chlorophyll breakdown and resulting in rapid color loss [42]. However, temperature interacts with other environmental variables, such as oxygen availability and humidity, to collectively determine the degradation rate. Proper control of temperature and humidity during storage and transportation is therefore essential to preserving vegetable color.
Metallic elements can also interact with chlorophyll molecules, catalyzing or inhibiting various degradation pathways. Yang et al. found that certain metals, such as iron and copper, can act as catalysts during oxidation, promoting the breakdown of chlorophyll. In contrast, magnesium, as the central atom in the chlorophyll molecule, can play a stabilizing role by binding to the chlorophyll structure and maintaining its integrity [43]. It is worth noting that oxygen levels also play a crucial role in the chlorophyll degradation process. Studies have shown that low-oxygen environments can effectively inhibit chlorophyll degradation even if they do not directly affect chlorophyll enzyme activity [44]. Therefore, controlling the oxygen environment during storage and packaging may be an effective way to both maintain the green color of vegetables and minimize the loss of valuable nutrients.
In short, the preservation of chlorophyll in vegetable products is a complex process, which is affected by many factors. Strategies such as controlling pH, limiting light, optimizing temperature conditions and using air-conditioned packaging help to slow chlorophyll degradation and ensure that the vegetable’s color and nutrients are preserved throughout its shelf-life. These approaches are critical to both consumer satisfaction and the economic viability of the fresh produce industry [45].

3. Methods to Inhibit Browning

Adverse color changes in vegetables during processing and storage not only diminish market appeal but also lead to nutritional loss and may lead to the formation of harmful compounds [10]. To address the issue, several strategies have been employed, including physical, chemical, and natural anti-browning method. The inhibition methods typically target and mitigate the three factors of color deterioration on varying degrees, thereby exhibiting a multifaceted effect.

3.1. Physical Methods

3.1.1. Cold Treatment

Cold treatment remains a cornerstone technology for inhibiting enzymatic and non-enzymatic browning in fresh-cut vegetables, primarily through metabolic suppression and enzyme inactivation. However, conventional freezing methods often compromise cellular integrity due to the formation of macro-scale ice crystals, which mechanically disrupt cell walls and accelerate pigment degradation [46]. This structural damage creates a paradoxical effect where the initial preservation method exacerbates quality deterioration through two pathways: (1) facilitating contact between phenolic substrates and PPO, and (2) promoting ascorbic acid oxidation through cellular compartmentalization loss [10,47]. Recent technological advancements attempt to resolve this dilemma through ice crystal morphology control. Notably, researchers demonstrated that high-pressure carbon dioxide freezing (HPCF) generates smaller, more uniform ice crystals in lettuce, achieving 53.50% lower polyphenol oxidase activity and 74.63% lower total phenolic content [48]. While these results are promising, the industrial scalability of HPCF warrants critical evaluation due to its high equipment costs and energy demands [49].
Emerging evidence suggests that synergistic approaches combining physical and biochemical interventions may offer more practical solutions. Isochoric freezing, which maintains constant volume during phase change, has shown particular effectiveness in preserving chlorophyll and ascorbic acid in leafy greens compared to isobaric systems [50]. This method, characterized by its capacity to maintain uniform temperature distribution and minimize ice crystal damage, has demonstrated significant advantages across different vegetable types. The schematic of isochoric freezing is shown in Figure S1. For instance, in spinach, thawed samples subjected to isometric freezing retained cellular integrity and exhibited levels of cell lysis comparable to fresh counterparts, indicating effective structural preservation [49]. In the case of potato, this technique not only delayed enzymatic browning by approximately one week but also enhanced the total phenolic content and antioxidant capacity of the samples, suggesting that isometric freezing may induce beneficial stress responses while inhibiting undesirable quality degradation [51]. These findings underscore the dual benefits of isometric freezing in maintaining both the physicochemical stability and nutritional functionality of vegetables during storage.
Future research should prioritize three key areas: (1) developing energy-efficient phase change materials for sustainable cold chains, (2) establishing vegetable-specific freezing protocols considering cellular architecture differences, and (3) integrating real-time quality monitoring systems with adaptive preservation parameter adjustment. Recent advances in hyperspectral imaging for early browning detection [52] and AI-driven predictive models for shelf-life estimation [53] provide critical enabling technologies for these developments.

3.1.2. Heat Treatment

Heat treatment has emerged as a dual-functional intervention for fresh-cut vegetable preservation, achieving enzymatic inactivation through both thermal denaturation of PPO and strategic modulation of cellular compartmentalization [54,55]. While conventional hot water blanching at 80–100 °C can reduce PPO activity by 60–80%, it simultaneously disrupts pectin-cellulose networks, causing texture softening and creating aqueous microenvironments that paradoxically enhance microbial proliferation [56]. This fundamental limitation has spurred the development of precision thermal technologies that decouple enzyme inactivation from structural degradation.
High-humidity hot air impingement blanching (HHAIB) has emerged as an effective thermal pretreatment for preserving the post-harvest texture and visual quality of vegetables. Specifically, HHAIB treatment at 110 °C with 40% relative humidity for 38 s has been shown to maintain the crispness of chili peppers while achieving minimal POD residual activity (0.52%) and a low browning index difference (7.09) [57]. These effects are primarily attributed to the rapid vapor condensation on the vegetable surface, which generates a protective thermal boundary layer that prevents excessive moisture loss and structural collapse. By limiting enzymatic activity and preserving cellular integrity, HHAIB offers a promising approach for extending the shelf-life and sensory attributes of fresh produce with minimal thermal damage [58]. The heat transfer mechanism description of HHAIB was shown in Figure S2.
Similarly, microwave-assisted blanching (MAB) at 300 W offers a time-efficient and nutrient-conserving alternative to conventional water blanching. By leveraging selective dielectric heating, MAB preferentially targets aqueous regions rich in PPO, thus accelerating enzyme inactivation. As a result, this method enhances ascorbic acid retention (13.59 mg/100 g), outperforming traditional water blanching (10.24 mg/100 g), and simultaneously reduces thermal exposure time [59]. However, these technologies face critical scalability challenges. HHHAS requires complex humidity control systems increasing capital costs by 30%, while microwave heterogeneity causes uneven treatment in dense vegetable tissues [60].
Nevertheless, critical knowledge gaps persist: (1) Nutrient–thermal stability trade-offs: While vitamin C degradation follows first-order kinetics, recent metabolomics reveals unexpected flavonoid enhancement through stress-induced biosynthesis [61]. (2) Matrix-specific thermal tolerance: Cruciferous vegetables exhibit higher heat resistance than leafy greens due to glucosinolate-mediated membrane stabilization, necessitating species-specific protocols [62]. Future research should prioritize intelligent thermal systems incorporating real-time spectroscopy feedback and adaptive control algorithms. The integration of pulsed electric fields with mild heating presents promise, with preliminary data showing that at reduced heat loads, PEFs inhibit browning by suppressing respiration rates during storage, and minimize nutrient loss from food by maintaining structural integrity [63,64]. Such hybrid approaches may ultimately reconcile the conflicting demands of microbial safety, nutritional preservation, and environmental sustainability in fresh-cut vegetable processing.

3.1.3. Gas-Conditioning Treatment

Gas-conditioning technologies, particularly modified atmosphere packaging (MAP), have evolved beyond simple gas composition adjustments to become sophisticated systems that dynamically interact with the physiological and biochemical pathways of fresh-cut vegetables [11]. The functional mechanism is illustrated in Figure S3. While traditional MAP can reduce browning, its fixed gas permeability often fails to accommodate the fluctuating respiration rates of high-metabolic produce, leading to anaerobic fermentation and the accumulation of off-flavors [65,66]. Although recent advancements in gas-responsive packaging have addressed some of these preservation challenges, a fundamental tension remains between technical performance and practical scalability.
Emerging systems employing pH-activated nanocomposite membranes exhibit enhanced gas modulation, with certain moisture-sensitive designs enabling CO2 enrichment while simultaneously inhibiting spoilage [67]. These adaptive materials offer promising gas selectivity and have demonstrated efficacy in reducing microbial populations and enzymatic browning. However, practical implementation remains difficult, especially with high-respiration produce. In leafy vegetables, excessive moisture exchange can paradoxically induce hypercapnic conditions, triggering secondary metabolic shifts such as alcohol accumulation and quality degradation. This challenge underscores the need for next-generation materials that can adapt dynamically to environmental changes, balancing gas composition and humidity management [68].
Concurrently, noble gas blends redefine inert gas roles by quenching singlet oxygen and stabilizing membranes, reducing lipid peroxidation by 50% and an effective reduction in microbial load [69]. Emerging 3D-MAP system integrates spatiotemporal control with 3D-printed gradient films and β-CD inclusions, but sustainability trade-offs remain: reduced carbon footprint but lack of mechanical robustness in bio-based films [70]. Future advancements must bridge nanoscale interventions with systems biology and digital twins, transitioning from empirical gas recipes to precision-controlled “gas-microbiome-vegetable” ecosystems that harmonize preservation efficacy, nutritional integrity, and environmental sustainability.

3.1.4. High-Pressure Treatment

High-pressure processing (HPP) has emerged as a dual-functional technology for fresh-cut vegetable preservation, leveraging hydrostatic pressure (typically 100–600 MPa) to simultaneously inhibit enzymatic browning and microbial proliferation while modulating cellular water dynamics [71]. By creating a transient pressure gradient across vegetable tissues, HPP restricts intracellular water migration, reducing dehydration-induced mass loss and preserving ascorbic acid through physical suppression of oxidative reactions, as demonstrated in pressurized argon-treated potatoes [72].
The technology’s ability to downregulate ethylene biosynthesis and respiratory rates extends shelf-life by delaying senescence, while its non-thermal nature minimizes thermal degradation of heat-sensitive nutrients [73]. However, the purported “sterilization-preservation synergy” remains contentious: although HPP achieves reductions in microorganisms, its sublethal stress (200–400 MPa) may induce microbial stress responses, enhancing biofilm formation and antibiotic resistance in surviving populations—a critical food safety paradox requiring mechanistic resolution [74]. Furthermore, the narrow therapeutic window between effective preservation and cellular damage poses significant challenges, with excessive pressure disrupting chloroplast ultrastructure and plasmalemma integrity, leading to irreversible texture softening and chlorophyll leakage. Current limitations in scalability and species-specific optimization—rooted in variable cellular architecture (e.g., parenchyma-rich vs. lignified tissues)—hinder broad adoption, as evidenced by inconsistent outcomes between leafy greens and root vegetables under identical protocols.
Future advancements must bridge fundamental research and industrial pragmatism through: (1) multiscale modeling of pressure-vegetable-microbiome interactions to define species-specific pressure thresholds, (2) hybrid systems integrating HPP with edible coatings to mitigate sublethal microbial recovery, and (3) energy-efficient pressure cycling technologies that minimize structural damage while enhancing process uniformity. By addressing these interdisciplinary challenges, HPP could transition from a niche intervention to a cornerstone of sustainable post-harvest strategies, balancing microbial safety with organoleptic and nutritional integrity.

3.1.5. Irradiation Treatment

Irradiation technologies, particularly ultraviolet (UV-C at 254 nm) and blue light (450 nm), have demonstrated dual functionality in fresh-cut vegetable preservation by simultaneously targeting microbial inactivation and modulating enzymatic browning pathways [75]. The bactericidal action of UV irradiation primarily arises from the formation of thymine dimers in microbial DNA, which disrupt replication and lead to cell death. Meanwhile, blue light exerts selective inhibitory effects on PPO isoforms via flavin-mediated photochemical reactions, as demonstrated by Zhao et al., who reported a significant reduction in browning indices of treated yam samples [76]. However, the practical effectiveness of UV-based interventions is hampered by their limited penetration depth. Surface moisture can scatter up to 70% of UV photons, creating protective niches for microbes within epidermal crevices—a limitation further exacerbated by the irregular and heterogeneous surface morphologies of many vegetables [77]. While combined UV-blue light strategies offer enhanced efficacy, they may also induce oxidative stress in plant tissues, accelerating ascorbic acid degradation and promoting membrane lipid peroxidation [78]. Emerging solutions like pulsed light [79] and plasma-activated UV systems [80] demonstrate improved uniformity, yet their scalability remains constrained by energy inefficiency and photoreactivation risks.
The broader landscape of physical preservation methods—from irradiation to high-pressure processing—faces a universal paradox: interventions suppressing browning often inadvertently compromise nutritional or textural integrity. For instance, gamma irradiation (3 kGy) reduces Bamboo shoot PPO activity but degrades nutrient composition through radiolytic cleavage [81]. Future innovations must prioritize intelligent synergy, such as UV-adaptive packaging with light-diffusing nanostructures to enhance irradiation uniformity, or hybrid systems coupling e-beam irradiation with antioxidant-releasing edible films to offset oxidative collateral damage. Table 1 provides an overview of recent innovations in the application of physical methods to inhibit vegetable Browning, showing the various advances and their respective results.

3.2. Chemical Methods

3.2.1. Acidifying Agents

Acidification approaches in fresh-cut produce preservation operate through pH-mediated enzyme suppression, capitalizing on the inverse correlation between acidic environments and oxidative enzyme functionality. Browning reactions driven by PPO and POD exhibit maximum activity in neutral pH conditions where specific metal ion coordination within the enzyme’s catalytic site becomes optimal. By strategically lowering pH to mildly acidic ranges, common organic acid treatments disrupt critical amino acid residues essential for enzymatic catalysis, particularly through protonation mechanisms affecting metal-binding domains. This pH-dependent interference leads to substantial reduction in oxidative enzyme performance, effectively decelerating pigment-generating biochemical pathways [121]. This pH manipulation concurrently disrupts non-enzymatic browning pathways: acidic environments inhibit Maillard reactions by protonating amino groups (-NH2 → -NH3+), effectively blocking Schiff base formation between reducing sugars and amino acids [28]. However, the technology’s purported dual antimicrobial–antioxidant benefits mask critical trade-offs—low pH induces proton stress in plant cells, accelerating chlorophyll degradation via Mg2+ displacement from porphyrin rings and triggering up to anthocyanin loss in pigmented vegetables (carrots, red cabbage), paradoxically diminishing visual appeal despite reduced browning [122].
Current limitations stem from oversimplified application paradigms. While citric and ascorbic acids remain industry standards, their buffering capacity varies dramatically. Repeated applications must therefore be made or there is a risk of sensory fatigue (acidity greater than the acceptability threshold of 2.5%) [123]. Future innovation must address the spatial heterogeneity of acid diffusion in vegetable tissues. Hybrid approaches integrating pH-triggered nanoemulsions (e.g., alginate-coated malic acid microparticles) with predictive AI-driven dosing systems could resolve these gradients, enabling precision acidification tailored to cellular architecture and real-time metabolic feedback. Such advancements would transform acidification from a blunt empirical tool to a dynamic, tissue-specific preservation modality.

3.2.2. Antioxidants

Antioxidants, while pivotal in mitigating enzymatic browning through redox modulation of quinones and stabilization of phenolic intermediates, face context-dependent efficacy challenges that demand critical reappraisal. Ascorbic acid, the industry benchmark, suppresses browning via a dual mechanism: chemically reducing o-quinones and competitively binding to PPO active sites [124]. However, its non-enzymatic regeneration cycle is pH- and light-sensitive—under fluorescent storage lighting, ascorbate degradation accelerates by 40% at pH > 5.5, paradoxically increasing semiquinone radical accumulation and subsequent browning rebound [125]. This temporal instability highlights a critical gap in single-antioxidant strategies, exacerbated by interspecies variability: cruciferous vegetables are more ascorbate stability than leafy greens due to glucosinolate-mediated redox buffering [126]. Emerging alternatives like melatonin present a paradigm shift, leveraging pleiotropic mechanisms beyond mere radical scavenging. At 100 μM, melatonin not only quenches peroxyl radicals (k = 2.1 × 106 M−1s−1) but also upregulates glutathione biosynthesis by 50% through NPR1-mediated salicylic acid pathways, as evidenced in Agaricus bisporus [127].
However, melatonin’s hydrophobic nature limits its solubility and diffusion in aqueous tissues, necessitating advanced delivery systems such as nanoemulsion encapsulation to improve bioavailability and tissue penetration [128]. The field now confronts a trilemma: balancing antioxidant efficacy, economic viability, and sensory compatibility. Future innovation must prioritize smart delivery systems—pH-responsive hydrogels or light-activated nanozymes—that spatially and temporally control antioxidant release, aligning redox activity with vegetable metabolic rhythms to achieve precision preservation without compromising affordability or sustainability.

3.2.3. Chelating Agents

Chelators, while effective in disrupting metal-catalyzed browning pathways, present a complex interplay between targeted inhibition and unintended metabolic consequences that demand nuanced evaluation. By sequestering Cu2+ from PPO active sites, conventional chelators like EDTA reduce enzymatic browning in model systems [129]. However, their performance in real food matrices is highly variable and species dependent. For example, cruciferous vegetables exhibit significantly higher resistance to PPO-metal chelation than leafy greens, largely due to glucosinolate-mediated shielding of metal ions [130,131]. Non-enzymatic preservation strategies employing metal chelation demonstrate dualistic metabolic impacts in plant tissues. While iron-binding approaches effectively inhibit lipid oxidation processes, they concurrently disrupt critical iron-dependent complexes in mitochondrial energy production systems. This collateral interference with electron transport mechanisms impairs cellular energy generation, paradoxically accelerating deterioration in physiologically active tissues despite oxidative stabilization [132]. The resulting metabolic trade-off—achieving pigment preservation while compromising energy homeostasis—reveals fundamental limitations in conventional chelation methodologies. Such antagonistic effects underscore the necessity for intervention strategies that achieve molecular specificity, particularly in decoupling antioxidant effects from essential cofactor depletion through temporally or spatially controlled chelation approaches [133].
Emerging plant-derived alternatives demonstrate selective chelation via ortho-dihydroxyphenyl moieties, binding Cu2+ while sparing essential Mg2+ and Ca2+ ions critical for texture preservation [134]. Lactoferrin, though promising for its dual antimicrobial–chelation capacity, faces scalability challenges due to its pH-dependent iron release, which risks pro-oxidant effects in low-acid vegetables [135]. The field now confronts a trilemma: optimizing chelator specificity to avoid nutrient stripping, balancing redox homeostasis, and maintaining cost-effectiveness. Future innovation must prioritize metabolically intelligent designs: CRISPR-edited vegetables with endogenous metal-binding peptides, or light-responsive chelators activated only during storage stress. Such systems could transcend the current trade-offs, enabling dynamic metal regulation aligned with vegetable physiology rather than static inhibition.
Recent innovations in the application of chemical methods to inhibit vegetable browning are presented in Table 2.

3.3. Natural Anti-Browning Method

The shift toward natural anti-browning agents reflects both consumer demand for clean-label products and scientific recognition of their multifunctional mechanisms. While plant-derived inhibitors like seabuckthorn extract and CGA demonstrate competitive binding to PPO active sites via hydrophobic interactions, their industrial adoption faces unresolved challenges [155]. Molecular docking studies showed that quercetin forms a stable π-π stack with amino acid residues of PPO. The number of interacting forces increases slightly at acidic pH, reversibly binding to PPO molecules and decreasing PPO activity [156]. However, batch-to-batch variability in phenolic content due to extraction methods undermines reproducibility—a critical barrier for standardization. Similarly, Similarly, the anti-browning efficacy of the inhibitor in foods decreases at pH values greater than 6.0 due to the ionization of its moiety, which weakens the interaction with the PPO amino acid residues [157]. These pH-dependent limitations highlight the need for microenvironment-tailored formulations [158]. Bioactive peptides exhibited broad-spectrum enzyme inhibition by disrupting FAD cofactor binding in peroxidases [159]. Yet, However, they are susceptible to hydrolysis during storage and are susceptible to environmental influences, so they need to be encapsulated for sustained release [160].
Synergistic combinations, such as 4-hexylresorcinol (4-HR) with ascorbic acid, demonstrate additive PPO inhibition (85% vs. 55% alone) by simultaneously targeting the enzyme’s oxy (4-HR) and deoxy (CGA) states [161,162]. However, natural extracts face an organoleptic trade-off: tannins are potent inhibitors of PPO, but excessive additions produce astringent flavors (tannin content > 2.5 mg/g) that exceed the threshold of consumer acceptance [163]. Future research must prioritize precision synergy—AI-driven optimization of inhibitor ratios to balance efficacy and palatability—while addressing extraction scalability. The path forward demands systems-level approaches integrating multi-omics (e.g., metabolomics-guided inhibitor screening), advanced delivery systems (light-responsive nanocapsules), and circular economy models (upcycled agricultural byproducts). Standardization frameworks must reconcile the inherent variability of natural extracts through robust quantification of active biomarkers (e.g., HPLC-MS for swertiamarin), ensuring batch consistency without compromising cost-effectiveness. Only by addressing these multidimensional challenges can natural anti-browning agents transition from laboratory curiosities to scalable, sensorially neutral solutions for the fresh-cut industry. Although the browning inhibition rate of certain natural extracts is generally lower than that of chemical preservatives, their Generally Recognized as Safe (GRAS) status and absence of synthetic residues align more closely with the clean-label movement and consumer demand for natural ingredients [164]. In contrast, while physical treatments such as heat or irradiation avoid chemical residues, they often exhibit limited efficacy in browning control and may compromise textural integrity-particularly in fresh-cut produce. Therefore, natural techniques offer a balanced alternative, prioritizing safety and sustainability, with performance potentially enhanced through formulation optimization or synergistic applications.
Table 3 summarized recent research exploring the use of natural extracts and bioactive compounds as anti-browning agents, highlighting the effectiveness and versatility of natural methods in preserving color and quality in fresh-cut vegetables. Future research can focus on optimizing the extraction and application of these natural inhibitors to improve their stability and effectiveness in commercial food processing.

4. Conclusions and Prospects

The preservation of color in vegetables remains a significant scientific and technological challenge requiring ongoing innovation. Color critically influences consumer perceptions of quality, encompassing not only appearance but also freshness, safety, and nutritional value. This review has summarized current research on color protection technologies, focusing on mechanisms of color degradation such as enzymatic browning, non-enzymatic browning, and chlorophyll degradation. By evaluating physical, chemical, and natural preservation approaches, we have highlighted their respective strengths and limitations, underscoring the need for further research to resolve key gaps.

4.1. Current Research and Gaps in Understanding Mechanisms

While advances in physical methods (e.g., modified atmosphere packaging), chemical agents (e.g., anti-browning compounds), and natural extracts (e.g., organic acids, plant antioxidants) show promise, the detailed cellular and molecular interactions of these agents with vegetable tissues remain poorly understood. Most studies focus on efficacy without exploring how these treatments affect enzyme activity, cellular integrity, or membrane permeability across diverse vegetable types. Furthermore, the impact of preservation methods on the nutritional quality—especially antioxidants, vitamins, and phenolic compounds—is insufficiently studied [186]. Deeper mechanistic insights are critical to optimize treatments that balance color retention with nutrient preservation.

4.2. Outlook on Novel Anti-Browning Agents and Natural Preservatives

Natural anti-browning agents derived from plant extracts such as rosemary, sage, and green tea offer safer, consumer-friendly alternatives to synthetic chemicals [187]. However, optimizing their dosage, effectiveness, and interactions with other food components requires further research. Exploring underutilized plants and agricultural by-products may reveal sustainable and economically viable new sources of effective color-preserving compounds.

4.3. The Promise of Biotechnology: Gene Editing and Plant Breeding

Biotechnological approaches like CRISPR-Cas9 gene editing and selective breeding hold great potential to develop vegetable varieties with inherent resistance to browning and chlorophyll degradation. Targeting genes responsible for PPO activity or enhancing natural antioxidant levels could reduce reliance on post-harvest treatments and extend shelf-life intrinsically [188]. However, regulatory, ethical, and consumer acceptance challenges must be addressed through transparent labeling and education to facilitate adoption.

4.4. Towards Integrated, Holistic Preservation Systems

Given the complexity of color preservation, integrated strategies combining physical, chemical, natural, and biotechnological methods are likely to be most effective. Advances in intelligent packaging—embedding sensors that detect spoilage and trigger-controlled release of anti-browning agents—offer promising, resource-efficient solutions that minimize preservative use and prolong shelf-life sustainably.

4.5. The Importance of Cross-Disciplinary Collaboration and Future Research Directions

Achieving breakthroughs in color preservation demands collaboration across food chemistry, plant science, biotechnology, and materials science. Integrating diverse expertise can lead to novel, practical solutions such as breeding vegetables optimized for post-harvest stability and developing advanced packaging tailored to perishable produce. Future research should also prioritize consumer preferences, safety, sustainability, and transparent communication to ensure public trust and acceptance of new technologies, especially those involving genetic modification or novel additives.
In summary, the future of color protection for vegetables is bright, with significant potential for enhancing both the shelf-life and nutritional value of fresh produce. By pursuing a balanced approach that combines scientific rigor with practical application, the food industry can meet consumer demands for high-quality, nutritious, and visually appealing products while addressing the global need for sustainable food systems. As research progresses, the development of integrated color-preservation systems that harness the strengths of multiple preservation technologies will be crucial, paving the way for a new era of innovation in vegetable quality preservation. From an application standpoint, several of the strategies exhibit promising translational potential. However, challenges remain in terms of production cost, shelf stability, and large-scale implementation. Future research should focus not only on mechanistic refinement but also on optimizing these technologies for commercial scalability.

Supplementary Materials

The following supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/foods14132222/s1, Figure S1. Schematic of isochoric chamber and generalized freezing process; Figure S2. Heat transfer mechanism depiction in HHAIB; Figure S3. Functional mechanism of gas transfer phenom.

Author Contributions

J.Z. (Jianing Zhang), writing—original draft, conceptualization, and funding acquisition. J.Z. (Junjun Zhang), editing and revisions and funding acquisition. L.Z., Y.X. and K.Z., editing and revisions. All authors have read and agreed to the published version of the manuscript.

Funding

The authors gratefully acknowledge the Project supported by the National Natural Science Foundation of China (Grant No. 32402185); the Natural Science Foundation of Jiangsu Province (BK20240841); the Postgraduate Research & Practice Innovation Program of Jiangsu Province (KYCX24_4015).

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

All authors declare that they have no conflicts of interest.

Abbreviations

PPOpolyphenol oxidase
MAPmodified atmosphere packaging
HHAIBhigh-humidity hot air impingement blanching
MABmicrowave-assisted blanching
PEFpulsed electric fields
HPPHigh-pressure processing
4-HR4-hexylresorcinol

References

  1. Zhang, Y.; Zareef, M.; Rong, Y.; Lin, H.; Chen, Q.; Ouyang, Q. Application of colorimetric sensor array coupled with chemometric methods for monitoring the freshness of snakehead fillets. Food Chem. 2024, 439, 138172. [Google Scholar] [CrossRef] [PubMed]
  2. Azam, S.M.R.; Ma, H.; Xu, B.; Devi, S.; Siddique, M.A.B.; Stanley, S.L.; Bhandari, B.; Zhu, J. Efficacy of ultrasound treatment in the removal of pesticide residues from fresh vegetables: A review. Trends Food Sci. Technol. 2020, 97, 417–432. [Google Scholar] [CrossRef]
  3. Gao, Q.; Li, Y.; Li, Y.; Zhang, Z.; Liang, Y. Antioxidant and prooxidant activities of phenolic acids commonly existed in vegetables and their relationship with structures. Food Sci. Technol. 2022, 42, e07622. [Google Scholar] [CrossRef]
  4. Chen, J.; Zhang, M.; Xu, B.; Sun, J.; Mujumdar, A.S. Artificial intelligence assisted technologies for controlling the drying of fruits and vegetables using physical fields: A review. Trends Food Sci. Technol. 2020, 105, 251–260. [Google Scholar] [CrossRef]
  5. Ali, M.; Batool, S.; Khalid, N.; Ali, S.; Raza, M.A.; Li, X.; Li, F.; Xinhua, Z. Recent trends in hydrogen-associated treatments for maintaining the postharvest quality of fresh and fresh-cut fruits and vegetables: A review. Food Control 2024, 156, 110114. [Google Scholar] [CrossRef]
  6. Xu, J.-G.; Huang, X.-N.; Meng, J.; Chen, J.-Y.; Han, B.-Z. Characterization and comparison of the bacterial community on environmental surfaces through a fresh-cut vegetables processing line in China. Food Res. Int. 2022, 155, 111075. [Google Scholar] [CrossRef]
  7. Ru, X.; You, W.; Zhang, J.; Xu, F.; Wu, Z.; Jin, P.; Zheng, Y.; Cao, S. γ-aminobutyric acid treatment inhibits browning and promotes storage quality by regulating reactive oxygen species and membrane lipid metabolism in fresh-cut stem lettuce. Food Chem. 2024, 459, 140420. [Google Scholar] [CrossRef]
  8. Hong, C.; Zhao, Y.M.; Zhou, C.; Guo, Y.; Ma, H. Ultrasonic washing as an abiotic elicitor to increase the phenolic content in fruits and vegetables: A review. Compr. Rev. Food Sci. Food Saf. 2023, 22, 785–808. [Google Scholar] [CrossRef]
  9. Moon, K.M.; Kwon, E.-B.; Lee, B.; Kim, C.Y. Recent Trends in Controlling the Enzymatic Browning of Fruit and Vegetable Products. Molecules 2020, 25, 2754. [Google Scholar] [CrossRef]
  10. Navina, B.; Keshav Huthaash, K.; Velmurugan, N.K.; Korumilli, T. Insights into recent innovations in anti browning strategies for fruit and vegetable preservation. Trends Food Sci. Technol. 2023, 139, 104128. [Google Scholar] [CrossRef]
  11. Obadi, M.; Guo, Q.; Sun, J.; Xu, B. Recent developments in the application of physical processing techniques for controlling browning in fresh wet noodles: A review. J. Cereal Sci. 2024, 118, 103951. [Google Scholar] [CrossRef]
  12. Arnold, M.; Gramza-Michałowska, A. Enzymatic browning in apple products and its inhibition treatments: A comprehensive review. Compr. Rev. Food Sci. Food Saf. 2022, 21, 5038–5076. [Google Scholar] [CrossRef]
  13. Hongyang, T.; Daming, H.; Xingyi, H.; Aheto, J.H.; Yi, R.; Yu, W.; Ji, L.; Shuai, N.; Mengqi, X. Detection of browning of fresh-cut potato chips based on machine vision and electronic nose. J. Food Process Eng. 2021, 44, e13631. [Google Scholar] [CrossRef]
  14. Winters, A.L.; Minchin, F.R.; Michaelson-Yeates, T.P.T.; Lee, M.R.F.; Morris, P. Latent and Active Polyphenol Oxidase (PPO) in Red Clover (Trifolium pratense) and Use of a Low PPO Mutant To Study the Role of PPO in Proteolysis Reduction. J. Agric. Food Chem. 2008, 56, 2817–2824. [Google Scholar] [CrossRef]
  15. Sui, X.; Meng, Z.; Dong, T.; Fan, X.; Wang, Q. Enzymatic browning and polyphenol oxidase control strategies. Curr. Opin. Biotechnol. 2023, 81, 102921. [Google Scholar] [CrossRef]
  16. Zhang, J.; Zhang, X.; Ye, M.; Li, X.W.; Lin, S.B.; Sun, X.-L. The Jasmonic Acid Pathway Positively Regulates the Polyphenol Oxidase-Based Defense against Tea Geometrid Caterpillars in the Tea Plant (Camellia sinensis). J. Chem. Ecol. 2020, 46, 308–316. [Google Scholar] [CrossRef]
  17. Terefe, N.S.; Delon, A.; Buckow, R.; Versteeg, C. Blueberry polyphenol oxidase: Characterization and the kinetics of thermal and high pressure activation and inactivation. Food Chem. 2015, 188, 193–200. [Google Scholar] [CrossRef]
  18. Zhan, L.; Hu, J.; Pang, L.; Li, Y.; Shao, J. Light exposure reduced browning enzyme activity and accumulated total phenols in cauliflower heads during cool storage. Postharvest Biol. Technol. 2014, 88, 17–20. [Google Scholar] [CrossRef]
  19. Zhou, C.; Hu, J.; Yu, X.; Yagoub, A.E.A.; Zhang, Y.; Ma, H.; Gao, X.; Otu, P.N.Y. Heat and/or ultrasound pretreatments motivated enzymolysis of corn gluten meal: Hydrolysis kinetics and protein structure. LWT 2017, 77, 488–496. [Google Scholar] [CrossRef]
  20. Wang, B.; Zhang, Y.; Venkitasamy, C.; Wu, B.; Pan, Z.; Ma, H. Effect of pulsed light on activity and structural changes of horseradish peroxidase. Food Chem. 2017, 234, 20–25. [Google Scholar] [CrossRef]
  21. Liu, Z.; Guo, Z.; Wu, D.; Fei, X.; Ei-Seedi, H.R.; Wang, C. High-pressure homogenization influences the functional properties of protein from oyster (Crassostrea gigas). LWT 2021, 151, 112107. [Google Scholar] [CrossRef]
  22. Xu, B.; Sylvain Tiliwa, E.; Yan, W.; Roknul Azam, S.M.; Wei, B.; Zhou, C.; Ma, H.; Bhandari, B. Recent development in high quality drying of fruits and vegetables assisted by ultrasound: A review. Food Res. Int. 2022, 152, 110744. [Google Scholar] [CrossRef] [PubMed]
  23. Bharate, S.S.; Bharate, S.B. Non-enzymatic browning in citrus juice: Chemical markers, their detection and ways to improve product quality. J. Food Sci. Technol. 2014, 51, 2271–2288. [Google Scholar] [CrossRef]
  24. Xu, Z.; Yang, Z.; Ji, J.; Mou, Y.; Chen, F.; Xiao, Z.; Liao, X.; Hu, X.; Ma, L. Polyphenol mediated non-enzymatic browning and its inhibition in apple juice. Food Chem. 2023, 404, 134504. [Google Scholar] [CrossRef]
  25. Pham, H.T.T.; Bista, A.; Kebede, B.; Buvé, C.; Hendrickx, M.; Van Loey, A. Insight into non-enzymatic browning of shelf-stable orange juice during storage: A fractionation and kinetic approach. J. Sci. Food Agric. 2020, 100, 3765–3775. [Google Scholar] [CrossRef]
  26. Pandiselvam, R.; Süfer, Ö.; Özaslan, Z.T.; Gowda, N.A.N.; Pulivarthi, M.K.; Charles, A.P.R.; Ramesh, B.; Ramniwas, S.; Rustagi, S.; Jafari, Z.; et al. Acrylamide in food products: Formation, technological strategies for mitigation, and future outlook. Food Front. 2024, 5, 1063–1095. [Google Scholar] [CrossRef]
  27. Sacchetti, G.; Ioannone, F.; De Gregorio, M.; Di Mattia, C.; Serafini, M.; Mastrocola, D. Non enzymatic browning during cocoa roasting as affected by processing time and temperature. J. Food Eng. 2016, 169, 44–52. [Google Scholar] [CrossRef]
  28. Liu, Q.-L.; Yi, Y.; Wang, S.-Q.; Wang, H.-X.; Xu, W.; Min, T.; Wang, L.-M. Non-enzymatic browning of lotus root during boiling. LWT 2023, 173, 114191. [Google Scholar] [CrossRef]
  29. Pham, H.T.T.; Kityo, P.; Buvé, C.; Hendrickx, M.E.; Van Loey, A.M. Influence of pH and Composition on Nonenzymatic Browning of Shelf-Stable Orange Juice during Storage. J. Agric. Food Chem. 2020, 68, 5402–5411. [Google Scholar] [CrossRef]
  30. Peng, H.; Gao, Y.; Zeng, C.; Hua, R.; Guo, Y.; Wang, Y.; Wang, Z. Effects of Maillard reaction and its product AGEs on aging and age-related diseases. Food Sci. Hum. Wellness 2024, 13, 1118–1134. [Google Scholar] [CrossRef]
  31. Cao, J.; Yang, C.; Zhang, J.; Zhang, L.; Tsao, R. Amadori compounds: Analysis, composition in food and potential health beneficial functions. Crit. Rev. Food Sci. Nutr. 2023, 65, 406–428. [Google Scholar] [CrossRef] [PubMed]
  32. Guo, A.; Kontoudakis, N.; Scollary, G.R.; Clark, A.C. Production and Isomeric Distribution of Xanthylium Cation Pigments and Their Precursors in Wine-like Conditions: Impact of Cu(II), Fe(II), Fe(III), Mn(II), Zn(II), and Al(III). J. Agric. Food Chem. 2017, 65, 2414–2425. [Google Scholar] [CrossRef]
  33. Grant-Preece, P.; Barril, C.; Schmidtke, L.M.; Scollary, G.R.; Clark, A.C. Light-induced changes in bottled white wine and underlying photochemical mechanisms. Crit. Rev. Food Sci. Nutr. 2017, 57, 743–754. [Google Scholar] [CrossRef]
  34. Shishir, M.R.I.; Taip, F.S.; Saifullah, M.; Aziz, N.A.; Talib, R.A. Effect of packaging materials and storage temperature on the retention of physicochemical properties of vacuum packed pink guava powder. Food Packag. Shelf Life 2017, 12, 83–90. [Google Scholar] [CrossRef]
  35. Barry, C.S. The stay-green revolution: Recent progress in deciphering the mechanisms of chlorophyll degradation in higher plants. Plant Sci. 2009, 176, 325–333. [Google Scholar] [CrossRef]
  36. Queiroz Zepka, L.; Jacob-Lopes, E.; Roca, M. Catabolism and bioactive properties of chlorophylls. Curr. Opin. Food Sci. 2019, 26, 94–100. [Google Scholar] [CrossRef]
  37. Lin, Y.-P.; Charng, Y.-Y. Chlorophyll dephytylation in chlorophyll metabolism: A simple reaction catalyzed by various enzymes. Plant Sci. 2021, 302, 110682. [Google Scholar] [CrossRef]
  38. Zhong, S.; Bird, A.; Kopec, R.E. The Metabolism and Potential Bioactivity of Chlorophyll and Metallo-chlorophyll Derivatives in the Gastrointestinal Tract. Mol. Nutr. Food Res. 2021, 65, 2000761. [Google Scholar] [CrossRef]
  39. Fang, H.; Luo, F.; Li, P.; Zhou, Q.; Zhou, X.; Wei, B.; Cheng, S.; Zhou, H.; Ji, S. Potential of jasmonic acid (JA) in accelerating postharvest yellowing of broccoli by promoting its chlorophyll degradation. Food Chem. 2020, 309, 125737. [Google Scholar] [CrossRef]
  40. Bai, J.; Jin, K.; Qin, W.; Wang, Y.; Yin, Q. Proteomic Responses to Alkali Stress in Oats and the Alleviatory Effects of Exogenous Spermine Application. Front. Plant Sci. 2021, 12, 627129. [Google Scholar] [CrossRef]
  41. Zhang, X.; Zhao, Z.; Zhang, M.; Wang, J.; Cheng, T.; Zhang, Q.; Pan, H. FsHemF is involved in the formation of yellow Forsythia leaves by regulating chlorophyll synthesis in response to light intensity. Plant Physiol. Biochem. 2023, 200, 107746. [Google Scholar] [CrossRef] [PubMed]
  42. Jahan, M.S.; Hasan, M.M.; Alotaibi, F.S.; Alabdallah, N.M.; Alharbi, B.M.; Ramadan, K.M.A.; Bendary, E.S.A.; Alshehri, D.; Jabborova, D.; Al-Balawi, D.A.; et al. Exogenous Putrescine Increases Heat Tolerance in Tomato Seedlings by Regulating Chlorophyll Metabolism and Enhancing Antioxidant Defense Efficiency. Plants 2022, 11, 1038. [Google Scholar] [CrossRef] [PubMed]
  43. Yang, L.; Ren, Q.; Ma, X.; Wang, M.; Sun, J.; Wang, S.; Wu, X.; Chen, X.; Wang, C.; Li, Q.; et al. New insight into the effect of riluzole on cadmium tolerance and accumulation in duckweed (Lemna turionifera). Ecotoxicol. Environ. Saf. 2022, 241, 113783. [Google Scholar] [CrossRef]
  44. Fang, Z.; Bouwkamp, J.C.; Solomos, T. Chlorophyllase activities and chlorophyll degradation during leaf senescence in non-yellowing mutant and wild type of Phaseolus vulgaris L. J. Exp. Bot. 1998, 49, 503–510. [Google Scholar] [CrossRef]
  45. Lefebvre, T.; Talbi, A.; Atwi-Ghaddar, S.; Destandau, E.; Lesellier, E. Development of an analytical method for chlorophyll pigments separation by reversed-phase supercritical fluid chromatography. J. Chromatogr. A 2020, 1612, 460643. [Google Scholar] [CrossRef]
  46. Alabi, K.P.; Zhu, Z.; Sun, D.-W. Transport phenomena and their effect on microstructure of frozen fruits and vegetables. Trends Food Sci. Technol. 2020, 101, 63–72. [Google Scholar] [CrossRef]
  47. Tian, Y.; Zhang, P.; Zhu, Z.; Sun, D.-W. Development of a single/dual-frequency orthogonal ultrasound-assisted rapid freezing technique and its effects on quality attributes of frozen potatoes. J. Food Eng. 2020, 286, 110112. [Google Scholar] [CrossRef]
  48. Ma, W.; Li, J.; Murtaza, A.; Iqbal, A.; Zhang, J.; Zhu, L.; Xu, X.; Pan, S.; Hu, W. High-pressure carbon dioxide treatment alleviates browning development by regulating membrane lipid metabolism in fresh-cut lettuce. Food Control 2022, 134, 108749. [Google Scholar] [CrossRef]
  49. Bilbao-Sainz, C.; Sinrod, A.G.J.; Dao, L.; Takeoka, G.; Williams, T.; Wood, D.; Bridges, D.F.; Powell-Palm, M.J.; Ukpai, G.; Chiou, B.; et al. Preservation of spinach by isochoric (constant volume) freezing. Int. J. Food Sci. Technol. 2020, 55, 2141–2151. [Google Scholar] [CrossRef]
  50. Nida, S.; Moses, J.A.; Anandharamakrishnan, C. Isochoric Freezing and Its Emerging Applications in Food Preservation. Food Eng. Rev. 2021, 13, 812–821. [Google Scholar] [CrossRef]
  51. Bilbao-Sainz, C.; Zhao, Y.; Takeoka, G.; Williams, T.; Wood, D.; Chiou, B.; Powell-Palm, M.J.; Wu, V.C.; Rubinsky, B.; McHugh, T. Effect of isochoric freezing on quality aspects of minimally processed potatoes. J. Food Sci. 2020, 85, 2656–2664. [Google Scholar] [CrossRef] [PubMed]
  52. Yang, K.; Zhao, M.; Argyropoulos, D. Machine learning based framework for the detection of mushroom browning using a portable hyperspectral imaging system. Postharvest Biol. Technol. 2025, 219, 113247. [Google Scholar] [CrossRef]
  53. Shi, C.; Zhao, Z.; Jia, Z.; Hou, M.; Yang, X.; Ying, X.; Ji, Z. Artificial neural network-based shelf life prediction approach in the food storage process: A review. Crit. Rev. Food Sci. Nutr. 2024, 64, 12009–12024. [Google Scholar] [CrossRef]
  54. Wang, Y.; Zhang, L.; Yu, X.; Zhou, C.; Yagoub, A.E.A.; Li, D. A Catalytic Infrared System as a Hot Water Replacement Strategy: A Future Approach for Blanching Fruits and Vegetables to Save Energy and Water. Food Rev. Int. 2024, 40, 641–657. [Google Scholar] [CrossRef]
  55. You, Y.; Zhou, Y.; Duan, X.; Mao, X.; Li, Y. Research progress on the application of different preservation methods for controlling fungi and toxins in fruit and vegetable. Crit. Rev. Food Sci. Nutr. 2023, 63, 12441–12452. [Google Scholar] [CrossRef]
  56. Jiang, H.; Ling, B.; Zhou, X.; Wang, S. Effects of combined radio frequency with hot water blanching on enzyme inactivation, color and texture of sweet potato. Innov. Food Sci. Emerg. Technol. 2020, 66, 102513. [Google Scholar] [CrossRef]
  57. Wang, H.; Zhang, Q.; Mujumdar, A.S.; Fang, X.-M.; Wang, J.; Pei, Y.-P.; Wu, W.; Zielinska, M.; Xiao, H.-W. High-humidity hot air impingement blanching (HHAIB) efficiently inactivates enzymes, enhances extraction of phytochemicals and mitigates brown actions of chili pepper. Food Control 2020, 111, 107050. [Google Scholar] [CrossRef]
  58. Alenyorege, E.A. Recent applications of high-humidity hot air impingement blanching (HHAIB) on plant materials: A review. Food Humanit. 2025, 5, 100657. [Google Scholar] [CrossRef]
  59. Saini, R.; Kaur, S.; Aggarwal, P.; Dhiman, A. The influence of conventional and novel blanching methods on potato granules, phytochemicals, and thermal properties of colored varieties. Front. Nutr. 2023, 10, 1178797. [Google Scholar] [CrossRef]
  60. Bei, X.; Yu, X.; Li, D.; Sun, Q.; Yu, Y.; Wang, Y.; Okonkwo, C.E.; Zhou, C. Heat source replacement strategy using catalytic infrared: A future for energy saving drying of fruits and vegetables. J. Food Sci. 2023, 88, 4827–4839. [Google Scholar] [CrossRef]
  61. Mao, C.; Wu, J.; Zhang, X.; Ma, F.; Cheng, Y. Improving the Solubility and Digestibility of Potato Protein with an Online Ultrasound-Assisted PH Shifting Treatment at Medium Temperature. Foods 2020, 9, 1908. [Google Scholar] [CrossRef] [PubMed]
  62. Shinali, T.S.; Zhang, Y.; Altaf, M.; Nsabiyeze, A.; Han, Z.; Shi, S.; Shang, N. The Valorization of Wastes and Byproducts from Cruciferous Vegetables: A Review on the Potential Utilization of Cabbage, Cauliflower, and Broccoli Byproducts. Foods 2024, 13, 1163. [Google Scholar] [CrossRef] [PubMed]
  63. Chang, C.-K.; Yang, Y.-T.; Gavahian, M.; Cheng, K.-C.; Hou, C.-Y.; Chen, M.-H.; Santoso, S.P.; Hsieh, C.-W. Prolonging the shelf-life of atemoya (Annona cherimola × Annona squamosa) using pulsed electric field treatments. Innov. Food Sci. Emerg. Technol. 2023, 88, 103458. [Google Scholar] [CrossRef]
  64. Wang, L.; Li, Z.; Liu, D.; Fan, J. Effect of heat and pulsed electric field treatment on the physicochemical and nutritional properties of carrots. J. Sci. Food Agric. 2023, 103, 1514–1521. [Google Scholar] [CrossRef]
  65. Zhang, X.-j.; Zhang, M.; Law, C.L.; Guo, Z. High-voltage electrostatic field-assisted modified atmosphere packaging for long-term storage of pakchoi and avoidance of off-flavors. Innov. Food Sci. Emerg. Technol. 2022, 79, 103032. [Google Scholar] [CrossRef]
  66. Cukrov, D.; Brizzolara, S.; Tonutti, P. Chapter 20—Physiological and Biochemical Effects of Controlled and Modified Atmospheres. In Postharvest Physiology and Biochemistry of Fruits and Vegetables; Yahia, E.M., Ed.; Woodhead Publishing: Cambridge, UK, 2019; pp. 425–441. [Google Scholar]
  67. Shi, X.; Xiao, H.; Kanamori, K.; Yonezu, A.; Lackner, K.S.; Chen, X. Moisture-Driven CO2 Sorbents. Joule 2020, 4, 1823–1837. [Google Scholar] [CrossRef]
  68. Qu, P.; Zhang, M.; Fan, K.; Guo, Z. Microporous modified atmosphere packaging to extend shelf life of fresh foods: A review. Crit. Rev. Food Sci. Nutr. 2022, 62, 51–65. [Google Scholar] [CrossRef]
  69. Olatunde, O.O.; Benjakul, S.; Vongkamjan, K. Comparative study on nitrogen and argon-based modified atmosphere packaging on microbiological, chemical, and sensory attributes as well as on microbial diversity of Asian sea bass. Food Packag. Shelf Life 2019, 22, 100404. [Google Scholar] [CrossRef]
  70. Li, S.; Jiang, Y.; Wang, M.; Li, R.; Dai, J.; Yan, J.; Qin, W.; Liu, Y. 3D printing of essential oil/β-cyclodextrin/popping candy modified atmosphere packaging for strawberry preservation. Carbohydr. Polym. 2022, 297, 120037. [Google Scholar] [CrossRef]
  71. Shen, X.; Zhang, M.; Devahastin, S.; Guo, Z. Effects of pressurized argon and nitrogen treatments in combination with modified atmosphere on quality characteristics of fresh-cut potatoes. Postharvest Biol. Technol. 2019, 149, 159–165. [Google Scholar] [CrossRef]
  72. Tsikrika, K.; O’Brien, N.; Rai, D.K. The Effect of High Pressure Processing on Polyphenol Oxidase Activity, Phytochemicals and Proximate Composition of Irish Potato Cultivars. Foods 2019, 8, 517. [Google Scholar] [CrossRef] [PubMed]
  73. Woolf, A.B.; Wibisono, R.; Farr, J.; Hallett, I.; Richter, L.; Oey, I.; Wohlers, M.; Zhou, J.; Fletcher, G.C.; Requejo-Jackman, C. Effect of high pressure processing on avocado slices. Innov. Food Sci. Emerg. Technol. 2013, 18, 65–73. [Google Scholar] [CrossRef]
  74. Liu, J.; Bi, J.; McClements, D.J.; Liu, X.; Yi, J.; Lyu, J.; Zhou, M.; Verkerk, R.; Dekker, M.; Wu, X.; et al. Impacts of thermal and non-thermal processing on structure and functionality of pectin in fruit- and vegetable-based products: A review. Carbohydr. Polym. 2020, 250, 116890. [Google Scholar] [CrossRef]
  75. Wen, B.; Cui, S.; Suo, X.; Supapvanich, S. Stress response of fresh-cut potatoes to laser irradiation before processing can prevent discoloration and maintain overall quality. Postharvest Biol. Technol. 2023, 197, 112213. [Google Scholar] [CrossRef]
  76. Zhao, L.; Zhao, X.; Ma, Y.; Zhang, Y.; Wang, D. Reduction of enzymatic browning of fresh-cut Chinese yam (Dioscorea opposita) by UV-C treatment. Int. Food Res. J. 2021, 28, 207–216. [Google Scholar] [CrossRef]
  77. Singh, H.; Bhardwaj, S.K.; Khatri, M.; Kim, K.-H.; Bhardwaj, N. UVC radiation for food safety: An emerging technology for the microbial disinfection of food products. Chem. Eng. J. 2021, 417, 128084. [Google Scholar] [CrossRef]
  78. Manzocco, L.; Nicoli, M.C. Surface Processing: Existing and Potential Applications of Ultraviolet Light. Crit. Rev. Food Sci. Nutr. 2015, 55, 469–484. [Google Scholar] [CrossRef]
  79. Zhang, J.; Yu, X.; Xu, B.; Yagoub, A.E.A.; Mustapha, A.T.; Zhou, C. Effect of intensive pulsed light on the activity, structure, physico-chemical properties and surface topography of polyphenol oxidase from mushroom. Innov. Food Sci. Emerg. Technol. 2021, 72, 102741. [Google Scholar] [CrossRef]
  80. Punia Bangar, S.; Trif, M.; Ozogul, F.; Kumar, M.; Chaudhary, V.; Vukic, M.; Tomar, M.; Changan, S. Recent developments in cold plasma-based enzyme activity (browning, cell wall degradation, and antioxidant) in fruits and vegetables. Compr. Rev. Food Sci. Food Saf. 2022, 21, 1958–1978. [Google Scholar] [CrossRef]
  81. Wang, J.; Jiang, J.; Wang, J.; Wang, Z.; Yang, X.; Jia, L. The influence of gamma irradiation on the storage quality of bamboo shoots. Radiat. Phys. Chem. 2019, 159, 124–130. [Google Scholar] [CrossRef]
  82. Otero, L.; Pozo, A. Effects of the application of static magnetic fields during potato freezing. J. Food Eng. 2022, 316, 110838. [Google Scholar] [CrossRef]
  83. Zhu, Z.; Luo, W.; Sun, D.-W. Effects of liquid nitrogen quick freezing on polyphenol oxidase and peroxide activities, cell water states and epidermal microstructure of wolfberry. LWT 2020, 120, 108923. [Google Scholar] [CrossRef]
  84. Qing, S.; Long, Y.; Wu, Y.; Shu, S.; Zhang, F.; Zhang, Y.; Yue, J. Hot-air-assisted radio frequency blanching of broccoli: Heating uniformity, physicochemical parameters, bioactive compounds, and microstructure. J. Sci. Food Agric. 2023, 103, 2664–2674. [Google Scholar] [CrossRef]
  85. Zhang, J.; Yagoub, A.E.A.; Sun, Y.; Arun, M.S.; Ma, H.; Zhou, C. Role of thermal and non-thermal drying techniques on drying kinetics and the physicochemical properties of shiitake mushroom. J. Sci. Food Agric. 2022, 102, 214–222. [Google Scholar] [CrossRef]
  86. Wang, Y.; Li, T.; Pan, Z.; Ye, X.; Ma, H. Effectiveness of combined catalytic infrared radiation and holding time for decontamination Aspergillus niger on dried shiitake mushrooms (Lentinus edodes) with different moisture contents. LWT 2023, 176, 114503. [Google Scholar] [CrossRef]
  87. Nguyen, T.V.L.; Tran, T.Y.N.; Lam, D.T.; Bach, L.G.; Nguyen, D.C. Effects of microwave blanching conditions on the quality of green asparagus (Asparagus officinalis L.) butt segment. Food Sci. Nutr. 2019, 7, 3513–3519. [Google Scholar] [CrossRef]
  88. Sicari, V.; Romeo, R.; Leporini, M.; Pellicanò, T.M.; Tundis, R.; Loizzo, M.R. Comparison of traditional hot water and vacuum assisted blanching methods on the physico-chemical quality parameters and antioxidant activity of zucchini (Cucurbita pepo L.) slices. J. Food Meas. Charact. 2022, 16, 281–294. [Google Scholar] [CrossRef]
  89. Zhang, L.; Yu, X.; Majumdar, A.S.; Zhou, C. Effect of freeze-thaw pretreatment combined with variable temperature on infrared and convection drying of lotus root. LWT 2022, 154, 112804. [Google Scholar] [CrossRef]
  90. Guo, Y.; Wu, B.; Guo, X.; Liu, D.; Wu, P.; Ma, H.; Pan, Z. Ultrasonication and thermosonication blanching treatments of carrot at varying frequencies: Effects on peroxidase inactivation mechanisms and quality characterization evaluation. Food Chem. 2021, 343, 128524. [Google Scholar] [CrossRef]
  91. Guo, Y.; Wu, B.; Guo, X.; Liu, D.; Qiu, C.; Ma, H. Effect of thermosonication on texture degradation of carrot tissue in relation to alterations in cell membrane and cell wall structure. Food Chem. 2022, 393, 133335. [Google Scholar] [CrossRef]
  92. Zhang, X.; Guo, L.; Tuly, J.; Hong, C.; Ma, H. Multi-frequency thermosonication washing treatments of fresh-cut lotus root: Effect on the accumulation of phenolics and storage quality characterization evaluation. Innov. Food Sci. Emerg. Technol. 2024, 93, 103641. [Google Scholar] [CrossRef]
  93. Wu, B.; Ma, Y.; Guo, X.; Guo, Y.; Qiu, C.; Gao, K.; Ma, H.; Pan, Z. Catalytic infrared blanching and drying of carrot slices with different thicknesses: Effects on surface dynamic crusting and quality characterization. Innov. Food Sci. Emerg. Technol. 2023, 88, 103444. [Google Scholar] [CrossRef]
  94. Feng, Y.; Xu, B.; Yagoub, A.E.A.; Ma, H.; Sun, Y.; Xu, X.; Yu, X.; Zhou, C. Role of drying techniques on physical, rehydration, flavor, bioactive compounds and antioxidant characteristics of garlic. Food Chem. 2021, 343, 128404. [Google Scholar] [CrossRef] [PubMed]
  95. Yan, J.-K.; Wu, L.-X.; Qiao, Z.-R.; Cai, W.-D.; Ma, H. Effect of different drying methods on the product quality and bioactive polysaccharides of bitter gourd (Momordica charantia L.) slices. Food Chem. 2019, 271, 588–596. [Google Scholar] [CrossRef]
  96. Guo, Y.; Wu, B.; Guo, X.; Ding, F.; Pan, Z.; Ma, H. Effects of power ultrasound enhancement on infrared drying of carrot slices: Moisture migration and quality characterizations. LWT 2020, 126, 109312. [Google Scholar] [CrossRef]
  97. Waseem, M.; Akhtar, S.; Ahmad, N.; Ismail, T.; Lazarte, C.E.; Hussain, M.; Manzoor, M.F.; Al-Farga, A. Effect of Microwave Heat Processing on Nutritional Indices, Antinutrients, and Sensory Attributes of Potato Powder-Supplemented Flatbread. J. Food Qual. 2022, 2022, 2103884. [Google Scholar] [CrossRef]
  98. Xu, J.; Wang, B.; Wang, Y. Electromagnetic fields assisted blanching—Effect on the dielectric and physicochemical properties of cabbage. J. Food Process Eng. 2019, 42, e13294. [Google Scholar] [CrossRef]
  99. Osae, R.; Zhou, C.; Xu, B.; Tchabo, W.; Tahir, H.E.; Mustapha, A.T.; Ma, H. Effects of ultrasound, osmotic dehydration, and osmosonication pretreatments on bioactive compounds, chemical characterization, enzyme inactivation, color, and antioxidant activity of dried ginger slices. J. Food Biochem. 2019, 43, e12832. [Google Scholar] [CrossRef]
  100. Mashabela, M.; Mahajan, P.V.; Sivakumar, D. Influence of different types of modified atmosphere packaging films and storage time on quality and bioactive compounds in fresh-cut cauliflower. Food Packag. Shelf Life 2019, 22, 100374. [Google Scholar] [CrossRef]
  101. Li, X.; Zhang, C.; Wang, X.; Liu, X.; Zhu, X.; Zhang, J. Integration of Metabolome and Transcriptome Profiling Reveals the Effect of Modified Atmosphere Packaging (MAP) on the Browning of Fresh-Cut Lanzhou Lily (Lilium davidii var. unicolor) Bulbs during Storage. Foods 2023, 12, 1335. [Google Scholar] [CrossRef]
  102. Jin, S.; Ding, Z.; Xie, J. Modified Atmospheric Packaging of Fresh-Cut Amaranth (Amaranthus tricolor L.) for Extending Shelf Life. Agriculture 2021, 11, 1016. [Google Scholar] [CrossRef]
  103. Sun, Y.; Zhao, X.; Ma, Y.; Guan, H.; Liang, H.; Wang, D. Inhibitory effect of modified atmosphere packaging on Escherichia coli O157:H7 in fresh-cut cucumbers (Cucumis sativus L.) and effectively maintain quality during storage. Food Chem. 2022, 369, 130969. [Google Scholar] [CrossRef] [PubMed]
  104. Barbosa, C.; Machado, T.B.; Alves, M.R.; Oliveira, M.B.P.P. Fresh-Cut Bell Peppers in Modified Atmosphere Packaging: Improving Shelf Life to Answer Food Security Concerns. Molecules 2020, 25, 2323. [Google Scholar] [CrossRef]
  105. Hu, H.; Li, P.; Shen, W. Preharvest application of hydrogen-rich water not only affects daylily bud yield but also contributes to the alleviation of bud browning. Sci. Hortic. 2021, 287, 110267. [Google Scholar] [CrossRef]
  106. Elwan, M.W.M.; Nasef, I.N.; El-Seifi, S.K.; Hassan, M.A.; Ibrahim, R.E. Storability, shelf-life and quality assurance of sugar snap peas (cv. super sugar snap) using modified atmosphere packaging. Postharvest Biol. Technol. 2015, 100, 205–211. [Google Scholar] [CrossRef]
  107. Yang, T.-D.; Chen, Y.-l.; Zeng, F.-K.; Ye, M.-Q.; Wang, L.; Luo, Z.; Qi, Y.-W.; Chen, F.-P. Effects of modified atmosphere packaging on the postharvest quality of mulberry leaf vegetable. Sci. Rep. 2022, 12, 10893. [Google Scholar] [CrossRef]
  108. Tang, J.; Cheng, J.; Li, Z.; Zhang, J.; Pan, Y. Ethanol fumigation combined with modified atmosphere packaging delays potato greening under light. Plant Physiol. Biochem. 2023, 202, 107962. [Google Scholar] [CrossRef]
  109. Almuhayawi, S.M.; Almuhayawi, M.S.; Al Jaouni, S.K.; Selim, S.; Hassan, A.H.A. Effect of Laser Light on Growth, Physiology, Accumulation of Phytochemicals, and Biological Activities of Sprouts of Three Brassica Cultivars. J. Agric. Food Chem. 2021, 69, 6240–6250. [Google Scholar] [CrossRef]
  110. Ufuk Kasım, M.; Kasım, R. Yellowing of fresh-cut spinach (Spinacia oleracea L.) leaves delayed by UV-B applications. Inf. Process. Agric. 2017, 4, 214–219. [Google Scholar] [CrossRef]
  111. Valerga, L.; González, R.E.; Pérez, M.B.; Concellón, A.; Cavagnaro, P.F. Differential and Cultivar-Dependent Antioxidant Response of Whole and Fresh-Cut Carrots of Different Root Colors to Postharvest UV-C Radiation. Plants 2023, 12, 1297. [Google Scholar] [CrossRef]
  112. Zhang, J.; Yagoub, A.E.A.; Sun, Y.; Mujumdar, A.S.; Ma, H.; Wahia, H.; Zhou, C. Intensive pulsed light pretreatment combined with controlled temperature and humidity for convection drying to reduce browning and improve quality of dried shiitake mushrooms. J. Sci. Food Agric. 2021, 101, 5608–5617. [Google Scholar] [CrossRef] [PubMed]
  113. Zhu, Y.; Li, C.; Cui, H.; Lin, L. Antimicrobial mechanism of pulsed light for the control of Escherichia coli O157:H7 and its application in carrot juice. Food Control 2019, 106, 106751. [Google Scholar] [CrossRef]
  114. Lin, L.; Wang, X.; He, R.; Cui, H. Action mechanism of pulsed magnetic field against E. coli O157:H7 and its application in vegetable juice. Food Control 2019, 95, 150–156. [Google Scholar] [CrossRef]
  115. Guo, Y.; Wu, B.; Lu, D.; Pan, Z.; Ma, H. Tri-frequency ultrasound as pretreatment to infrared drying of carrots: Impact on enzyme inactivation, color changes, nutrition quality parameters and microstructures. Int. J. Food Eng. 2021, 17, 275–284. [Google Scholar] [CrossRef]
  116. Osae, R.; Zhou, C.; Tchabo, W.; Xu, B.; Bonah, E.; Alenyorege, E.A.; Ma, H. Optimization of osmosonication pretreatment of ginger (Zingiber officinale Roscoe) using response surface methodology: Effect on antioxidant activity, enzyme inactivation, phenolic compounds, and physical properties. J. Food Process Eng. 2019, 42, e13218. [Google Scholar] [CrossRef]
  117. Sun, Y.; Ji, D.; Ma, H.; Chen, X. Ultrasound accelerated γ-aminobutyric acid accumulation in coffee leaves through influencing the microstructure, enzyme activity, and metabolites. Food Chem. 2022, 385, 132646. [Google Scholar] [CrossRef]
  118. Faisal Manzoor, M.; Ahmed, Z.; Ahmad, N.; Karrar, E.; Rehman, A.; Aadil, R.M.; Al-Farga, A.; Iqbal, M.W.; Rahaman, A.; Zeng, X. Probing the combined impact of pulsed electric field and ultra-sonication on the quality of spinach juice. J. Food Process. Preserv. 2021, 45, e15475. [Google Scholar] [CrossRef]
  119. Wang, Y.; Li, L.; Wang, B.; Xu, J. Selective enzyme inactivation in a simulated system and in cabbage juice using electrospray technology. Innov. Food Sci. Emerg. Technol. 2022, 75, 102875. [Google Scholar] [CrossRef]
  120. Alenyorege, E.A.; Ma, H.; Aheto, J.H.; Agyekum, A.A.; Zhou, C. Effect of sequential multi-frequency ultrasound washing processes on quality attributes and volatile compounds profiling of fresh-cut Chinese cabbage. LWT 2020, 117, 108666. [Google Scholar] [CrossRef]
  121. Gong, W.; Shi, B.; Zeng, F.K.; Dong, N.; Lei, Z.; Liu, J. Evaluation of cooking, nutritional, and quality characteristics of fresh-cut potato slice pretreated with acetic acid. J. Food Sci. 2022, 87, 427–437. [Google Scholar] [CrossRef]
  122. Roy, S.; Rhim, J.-W. Anthocyanin food colorant and its application in pH-responsive color change indicator films. Crit. Rev. Food Sci. Nutr. 2021, 61, 2297–2325. [Google Scholar] [CrossRef] [PubMed]
  123. Gérard, V.; Ay, E.; Graff, B.; Morlet-Savary, F.; Galopin, C.; Mutilangi, W.; Lalevée, J. Ascorbic Acid Derivatives as Potential Substitutes for Ascorbic Acid To Reduce Color Degradation of Drinks Containing Ascorbic Acid and Anthocyanins from Natural Extracts. J. Agric. Food Chem. 2019, 67, 12061–12071. [Google Scholar] [CrossRef] [PubMed]
  124. Yu, A.-N.; Zhou, Y.-Y.; Yang, Y.-N. Kinetics of browning and correlations between browning degree and pyrazine compounds in l-ascorbic acid/acidic amino acid model systems. Food Chem. 2017, 221, 1678–1684. [Google Scholar] [CrossRef] [PubMed]
  125. Njus, D.; Kelley, P.M.; Tu, Y.-J.; Schlegel, H.B. Ascorbic acid: The chemistry underlying its antioxidant properties. Free Radic. Biol. Med. 2020, 159, 37–43. [Google Scholar] [CrossRef]
  126. Šamec, D.; Pavlović, I.; Radojčić Redovniković, I.; Salopek-Sondi, B. Comparative analysis of phytochemicals and activity of endogenous enzymes associated with their stability, bioavailability and food quality in five Brassicaceae sprouts. Food Chem. 2018, 269, 96–102. [Google Scholar] [CrossRef]
  127. Feng, L.; Jiang, X.; Kitazawa, H.; Wang, X.; Guo, Y.; Li, L.; Liu, H.; Wang, Y.; Wang, J. Characterization of bioactive films loaded with melatonin and regulation of postharvest ROS scavenging and ascorbate-glutathione cycle in Agaricus bisporus. Postharvest Biol. Technol. 2022, 194, 112107. [Google Scholar] [CrossRef]
  128. Sruthi, S.; Millot, N.; Mohanan, P.V. Zinc oxide nanoparticles mediated cytotoxicity, mitochondrial membrane potential and level of antioxidants in presence of melatonin. Int. J. Biol. Macromol. 2017, 103, 808–818. [Google Scholar] [CrossRef]
  129. Fan, X. Chemical inhibition of polyphenol oxidase and cut surface browning of fresh-cut apples. Crit. Rev. Food Sci. Nutr. 2023, 63, 8737–8751. [Google Scholar] [CrossRef]
  130. Zhang, W.; Pan, Y.; Jiang, Y.; Zhang, Z. Advances in control technologies and mechanisms to treat peel browning in postharvest fruit. Sci. Hortic. 2023, 311, 111798. [Google Scholar] [CrossRef]
  131. Dawood, M.F.A.; Azooz, M.M. Concentration-dependent effects of tungstate on germination, growth, lignification-related enzymes, antioxidants, and reactive oxygen species in broccoli (Brassica oleracea var. italica L.). Environ. Sci. Pollut. Res. 2019, 26, 36441–36457. [Google Scholar] [CrossRef]
  132. Wang, S.-J.; Zhai, S.; Xu, X.-T.; Lu, Y.-T.; Yuan, T.-T. Hydrogen peroxide participates in leaf senescence by inhibiting CHLI1 activity. Plant Cell Rep. 2024, 43, 258. [Google Scholar] [CrossRef] [PubMed]
  133. Miwa, S.; Kashyap, S.; Chini, E.; von Zglinicki, T. Mitochondrial dysfunction in cell senescence and aging. J. Clin. Investig. 2022, 132, e158447. [Google Scholar] [CrossRef] [PubMed]
  134. Pochodylo, A.L.; Aristilde, L. Molecular dynamics of stability and structures in phytochelatin complexes with Zn, Cu, Fe, Mg, and Ca: Implications for metal detoxification. Environ. Chem. Lett. 2017, 15, 495–500. [Google Scholar] [CrossRef]
  135. Meng, Q.; Jiang, H.; Tu, J.; He, Y.; Zhou, Z.; Wang, R.; Jin, W.; Han, J.; Liu, W. Effect of pH, protein/polysaccharide ratio and preparation method on the stability of lactoferrin-polysaccharide complexes. Food Chem. 2024, 456, 140056. [Google Scholar] [CrossRef]
  136. Ali, S.; Khan, A.S.; Anjum, M.A.; Nawaz, A.; Naz, S.; Ejaz, S.; Hussain, S. Effect of postharvest oxalic acid application on enzymatic browning and quality of lotus (Nelumbo nuciferaGaertn.) root slices. Food Chem. 2020, 312, 126051. [Google Scholar] [CrossRef]
  137. Ali, S.; Anjum, M.A.; Nawaz, A.; Naz, S.; Hussain, S.; Ejaz, S.; Sardar, H. Effect of pre-storage ascorbic acid and Aloe vera gel coating application on enzymatic browning and quality of lotus root slices. J. Food Biochem. 2020, 44, e13136. [Google Scholar] [CrossRef]
  138. Chung, Y.B.; Song, H.; Jo, K.; Suh, H.J. Effect of ascorbic acid and citric acid on the quality of salted Chinese cabbage during storage. Food Sci. Biotechnol. 2021, 30, 227–234. [Google Scholar] [CrossRef]
  139. Ali, S.; Khan, A.S.; Nawaz, A.; Naz, S.; Ejaz, S.; Ullah, S. Glutathione application delays surface browning of fresh-cut lotus (Nelumbo nucifera Gaertn.) root slices during low temperature storage. Postharvest Biol. Technol. 2023, 200, 112311. [Google Scholar] [CrossRef]
  140. Li, F.; Han, Q.; Wang, W.; Wu, S. Carboxymethyl chitosan-based coatings loaded with glutathione extend the shelf-life of harvested enoki mushrooms (Flammulina velutipes). LWT 2022, 166, 113807. [Google Scholar] [CrossRef]
  141. Sohail, M.; Wills, R.B.H.; Bowyer, M.C.; Pristijono, P. Beneficial impact of exogenous arginine, cysteine and methionine on postharvest senescence of broccoli. Food Chem. 2021, 338, 128055. [Google Scholar] [CrossRef]
  142. Wen, B.; Li, D.; Tang, D.; Huang, Z.; Kedbanglai, P.; Ge, Z.; Du, X.; Supapvanich, S. Effects of simultaneous ultrasonic and cysteine treatment on antibrowning and physicochemical quality of fresh-cut lotus roots during cold storage. Postharvest Biol. Technol. 2020, 168, 111294. [Google Scholar] [CrossRef]
  143. Jiang, W.; Zhu, D.; Zhao, L.; Liu, Y.; Wang, C.; Farid, M.S.; Gu, Y.; Li, J.; Li, T.; Sun, Y.; et al. l-Cysteine Treatment Delayed the Quality Deterioration of Fresh-Cut Button Mushrooms by Regulating Oxygen Metabolism, Inhibiting Water Loss, and Stimulating Endogenous H2S Production. J. Agric. Food Chem. 2023, 71, 974–984. [Google Scholar] [CrossRef] [PubMed]
  144. Shekari, A.; Hassani, R.N.; Aghdam, M.S.; Rezaee, M.; Jannatizadeh, A. The effects of melatonin treatment on cap browning and biochemical attributes of Agaricus bisporus during low temperature storage. Food Chem. 2021, 348, 129074. [Google Scholar] [CrossRef] [PubMed]
  145. Luo, S.; Hu, H.; Wang, Y.; Zhou, H.; Zhang, Y.; Zhang, L.; Li, P. The role of melatonin in alleviating the postharvest browning of lotus seeds through energy metabolism and membrane lipid metabolism. Postharvest Biol. Technol. 2020, 167, 111243. [Google Scholar] [CrossRef]
  146. Wang, M.; Xu, J.; Ding, Z.; Xie, J. Prolong the postharvest shelf life of spinach through the antioxidative ability of melatonin. Food Chem. X 2023, 19, 100769. [Google Scholar] [CrossRef]
  147. Sun, L.; Luo, S.; Huali, H.; Zhou, H.; Zhang, Y.; An, R.; Ling, J.; Li, P. Melatonin promotes the normal cellular mitochondrial function of lotus seeds through stimulating nitric oxide production. Postharvest Biol. Technol. 2022, 185, 111814. [Google Scholar] [CrossRef]
  148. Hu, H.; Luo, S.; An, R.; Li, P. Endogenous melatonin delays sepal senescence and extends the storage life of broccoli florets by decreasing ethylene biosynthesis. Postharvest Biol. Technol. 2022, 188, 111894. [Google Scholar] [CrossRef]
  149. Alenyorege, E.A.; Ma, H.; Ayim, I.; Aheto, J.H.; Hong, C.; Zhou, C. Reduction of Listeria innocua in fresh-cut Chinese cabbage by a combined washing treatment of sweeping frequency ultrasound and sodium hypochlorite. LWT 2019, 101, 410–418. [Google Scholar] [CrossRef]
  150. He, M.; Fan, M.; Liu, W.; Li, Y.; Wang, G. Design, synthesis, molecular modeling, and biological evaluation of novel kojic acid derivatives containing bioactive heterocycle moiety as inhibitors of tyrosinase and antibrowning agents. Food Chem. 2021, 362, 130241. [Google Scholar] [CrossRef]
  151. Feng, Y.; Zhang, D.; Sun, Y.; Meng, Z.; Malik, A.U.; Zhang, S.; Yan, H.; Wang, Q. A novel anti-browning agent S-furfuryl thioacetate for fresh-cut potato screened from food-derived sulfur compounds. Postharvest Biol. Technol. 2022, 192, 112007. [Google Scholar] [CrossRef]
  152. Feng, Y.; Sun, Y.; Meng, Z.; Sui, X.; Zhang, D.; Yan, H.; Wang, Q. S-Ethyl thioacetate as a natural anti-browning agent can significantly inhibit the browning of fresh-cut potatoes by decreasing polyphenol oxidase activity. Sci. Hortic. 2022, 305, 111427. [Google Scholar] [CrossRef]
  153. Wang, G.; He, M.; Huang, Y.; Peng, Z. Synthesis and biological evaluation of new kojic acid-1,3,4-oxadiazole hybrids as tyrosinase inhibitors and their application in the anti-browning of fresh-cut mushrooms. Food Chem. 2023, 409, 135275. [Google Scholar] [CrossRef] [PubMed]
  154. Xu, H.; Li, X.; Mo, L.; Zou, Y.; Zhao, G. Tyrosinase inhibitory mechanism and the anti-browning properties of piceid and its ester. Food Chem. 2022, 390, 133207. [Google Scholar] [CrossRef] [PubMed]
  155. Zhang, Z.; Peng, Y.; Meng, W.; Pei, L.; Zhang, X. Browning inhibition of seabuckthorn leaf extract on fresh-cut potato sticks during cold storage. Food Chem. 2022, 389, 133076. [Google Scholar] [CrossRef]
  156. Jiang, H.; Zhou, L.; Sun, Y.; Yu, K.; Yu, W.; Tian, Y.; Liu, J.; Zou, L.; Liu, W. Polyphenol oxidase inhibited by 4-hydroxycinnamic acid and naringenin: Multi-spectroscopic analyses and molecular docking simulation at different pH. Food Chem. 2022, 396, 133662. [Google Scholar] [CrossRef]
  157. Kaya, E.D.; Türkhan, A.; Gür, F.; Gür, B. A novel method for explaining the product inhibition mechanisms via molecular docking: Inhibition studies for tyrosinase from Agaricus bisporus. J. Biomol. Struct. Dyn. 2022, 40, 7926–7939. [Google Scholar] [CrossRef]
  158. Cheng, D.; Wang, G.; Tang, J.; Yao, C.; Li, P.; Song, Q.; Wang, C. Inhibitory effect of chlorogenic acid on polyphenol oxidase and browning of fresh-cut potatoes. Postharvest Biol. Technol. 2020, 168, 111282. [Google Scholar] [CrossRef]
  159. Liu, X.; Lu, Y.; Yang, Q.; Yang, H.; Li, Y.; Zhou, B.; Li, T.; Gao, Y.; Qiao, L. Cod peptides inhibit browning in fresh-cut potato slices: A potential anti-browning agent of random peptides for regulating food properties. Postharvest Biol. Technol. 2018, 146, 36–42. [Google Scholar] [CrossRef]
  160. Rivero-Pino, F. Bioactive food-derived peptides for functional nutrition: Effect of fortification, processing and storage on peptide stability and bioactivity within food matrices. Food Chem. 2023, 406, 135046. [Google Scholar] [CrossRef]
  161. Arias, E.; González, J.; Peiró, J.M.; Oria, R.; Lopez-Buesa, P. Browning Prevention by Ascorbic Acid and 4-Hexylresorcinol: Different Mechanisms of Action on Polyphenol Oxidase in the Presence and in the Absence of Substrates. J. Food Sci. 2007, 72, C464–C470. [Google Scholar] [CrossRef]
  162. de OG Silva, C.; Sun, P.; Barrett, K.; Sanders, M.G.; van Berkel, W.J.; Kabel, M.A.; Meyer, A.S.; Agger, J.W. Polyphenol oxidase activity on guaiacyl and syringyl lignin units. Angew. Chem. 2024, 136, e202409324. [Google Scholar] [CrossRef]
  163. Zhou, M.; Chen, J.; Bi, J.; Li, X.; Xin, G. The roles of soluble poly and insoluble tannin in the enzymatic browning during storage of dried persimmon. Food Chem. 2022, 366, 130632. [Google Scholar] [CrossRef] [PubMed]
  164. de Chiara, M.L.V.; Castagnini, J.M.; Capozzi, V. Cutting-edge physical techniques in postharvest for fruits and vegetables: Unveiling their power, inclusion in ‘hurdle’ approach, and latest applications. Trends Food Sci. Technol. 2024, 151, 104619. [Google Scholar] [CrossRef]
  165. Paudel, P.; Seong, S.H.; Wagle, A.; Min, B.S.; Jung, H.A.; Choi, J.S. Antioxidant and anti-browning property of 2-arylbenzofuran derivatives from Morus alba Linn root bark. Food Chem. 2020, 309, 125739. [Google Scholar] [CrossRef] [PubMed]
  166. Lim, W.Y.; Cheun, C.F.; Wong, C.W. Inhibition of enzymatic browning in sweet potato (Ipomoea batatas (L.)) with chemical and natural anti-browning agents. J. Food Process. Preserv. 2019, 43, e14195. [Google Scholar] [CrossRef]
  167. Zhu, Y.; Du, X.; Zheng, J.; Wang, T.; You, X.; Liu, H.; Liu, X. The effect of ultrasonic on reducing anti-browning minimum effective concentration of purslane extract on fresh-cut potato slices during storage. Food Chem. 2021, 343, 128401. [Google Scholar] [CrossRef]
  168. Liu, X.; Chen, T.; Wang, Q.; Liu, J.; Lu, Y.; Shi, Y. Structure Analysis and Study of Biological Activities of Condensed Tannins from Bruguiera gymnorhiza (L.) Lam and Their Effect on Fresh-Cut Lotus Roots. Molecules 2021, 26, 1369. [Google Scholar] [CrossRef]
  169. Tanhaş, E.; Martin, E.; Korucu, E.N.; Dirmenci, T. Effect of aqueous extract, hydrosol, and essential oil forms of some endemic Origanum L. (Lamiaceae) taxa on polyphenol oxidase activity in fresh-cut mushroom samples. J. Food Process. Preserv. 2020, 44, e14726. [Google Scholar] [CrossRef]
  170. Xiao, Y.; He, J.; Zeng, J.; Yuan, X.; Zhang, Z.; Wang, B. Application of citronella and rose hydrosols reduced enzymatic browning of fresh-cut taro. J. Food Biochem. 2020, 44, e13283. [Google Scholar] [CrossRef]
  171. Jirasuteeruk, C.; Theerakulkait, C. Ultrasound-Assisted Extraction of Phenolic Compounds from Mango (Mangifera indica cv. Chok Anan) Peel and Its Inhibitory Effect on Enzymatic Browning of Potato Puree. Food Technol. Biotechnol. 2019, 57, 350–357. [Google Scholar] [CrossRef]
  172. Zhang, L.; Deng, N.; Yagoub, A.E.A.; Chen, L.; Mustapha, A.T.; Yu, X.; Zhou, C. Ultrasound-assisted probiotics fermentation suspension treatment under mild heat to improve the storage quality of freshly cut lotus root. Food Chem. 2022, 397, 133823. [Google Scholar] [CrossRef] [PubMed]
  173. Zhang, L.; Yu, X.; Yagoub, A.E.A.; Xia, G.; Zhou, C. Effect of vacuum impregnation assisted probiotics fermentation suspension on shelf life quality of freshly cut lotus root. Food Chem. 2022, 381, 132281. [Google Scholar] [CrossRef] [PubMed]
  174. Shen, C.; Chen, W.; Li, C.; Chen, X.; Cui, H.; Lin, L. Pickering emulsion stabilized by gliadin/soybean polysaccharide composite colloidal nanoparticle: Physicochemical properties and its application on washing of fresh-cut cabbage. Food Res. Int. 2022, 161, 111886. [Google Scholar] [CrossRef]
  175. Cui, H.; Bai, M.; Yuan, L.; Surendhiran, D.; Lin, L. Sequential effect of phages and cold nitrogen plasma against Escherichia coli O157:H7 biofilms on different vegetables. Int. J. Food Microbiol. 2018, 268, 1–9. [Google Scholar] [CrossRef]
  176. Lin, L.; Wang, X.; Cui, H. Synergistic efficacy of pulsed magnetic fields and Litseacubeba essential oil treatment against Escherichia coli O157:H7 in vegetable juices. Food Control 2019, 106, 106686. [Google Scholar] [CrossRef]
  177. Duan, L.; Jiang, T.; Zhou, Y.; Bai, X.; Wang, Y.; Lü, X.; Xia, X.; Lin, L.; Shi, C. The inactivation of Shigella flexneri by synergistic effect of ultrasound combined with basil essential oil nanoemulsion and application in cabbage cleaning. Food Control 2024, 156, 110142. [Google Scholar] [CrossRef]
  178. Dai, J.; Li, C.; Cui, H.; Lin, L. Unraveling the anti-bacterial mechanism of Litsea cubeba essential oil against E. coli O157:H7 and its application in vegetable juices. Int. J. Food Microbiol. 2021, 338, 108989. [Google Scholar] [CrossRef]
  179. Song, L.; Yang, H.; Cheng, S.; Zhang, Z.; Zhang, L.; Su, R.; Li, Y.; Zhan, X.; Yang, B.; Lin, L.; et al. Combination effects of ultrasound and citral nanoemulsion against Shigella flexneri and the preservation effect on fresh-cut carrots. Food Control 2024, 155, 110069. [Google Scholar] [CrossRef]
  180. Yang, Z.; Li, M.; Li, Y.; Huang, X.; Li, Z.; Zhai, X.; Shi, J.; Zou, X.; Xiao, J.; Sun, Y.; et al. Sodium alginate/guar gum based nanocomposite film incorporating β-Cyclodextrin/persimmon pectin-stabilized baobab seed oil Pickering emulsion for mushroom preservation. Food Chem. 2024, 437, 137891. [Google Scholar] [CrossRef]
  181. Rashid, A.; Qayum, A.; Bacha, S.A.S.; Liang, Q.; Liu, Y.; Kang, L.; Chi, Z.; Chi, R.; Han, X.; Ekumah, J.-N.; et al. Preparation and functional characterization of pullulan-sodium alginate composite film enhanced with ultrasound-assisted clove essential oil Nanoemulsions for effective preservation of cherries and mushrooms. Food Chem. 2024, 457, 140048. [Google Scholar] [CrossRef]
  182. Cui, H.; Zhang, C.; Li, C.; Lin, L. Inhibition of Escherichia coli O157:H7 biofilm on vegetable surface by solid liposomes of clove oil. LWT 2020, 117, 108656. [Google Scholar] [CrossRef]
  183. Zhang, C.; Li, C.; Abdel-Samie, M.A.; Cui, H.; Lin, L. Unraveling the inhibitory mechanism of clove essential oil against Listeria monocytogenes biofilm and applying it to vegetable surfaces. LWT 2020, 134, 110210. [Google Scholar] [CrossRef]
  184. Lin, L.; Liao, X.; Li, C.; Abdel-Samie, M.A.; Siva, S.; Cui, H. Cold nitrogen plasma modified cuminaldehyde/β-cyclodextrin inclusion complex and its application in vegetable juices preservation. Food Res. Int. 2021, 141, 110132. [Google Scholar] [CrossRef] [PubMed]
  185. Zhang, X.; Xin, Y.; Wang, J.; Dhanasekaran, S.; Yue, Q.; Feng, F.; Gu, X.; Li, B.; Zhao, L.; Zhang, H. Characterization of a Bacillus velezensis strain as a potential biocontrol agent against soft rot of eggplant fruits. Int. J. Food Microbiol. 2024, 410, 110480. [Google Scholar] [CrossRef]
  186. Rocchetti, G.; Galimberti, S.; Callegari, M.L.; Lucini, L. Metabolomics and proteomics approaches provide a better understanding of non-enzymatic browning and pink discoloration in dairy products: A mini review. Food Biosci. 2023, 56, 103328. [Google Scholar] [CrossRef]
  187. Huang, J.; Li, M.; Han, C.; Zhang, Z.; Liu, X.; Ying, Z.; Yin, P.; Yang, L. Structural and mechanistic insights into the anti-tyrosinase, anti-melanogenesis, and anti-browning effect of proanthocyanidins from seed coats of Acer truncatum Bunge. Int. J. Biol. Macromol. 2025, 284, 138246. [Google Scholar] [CrossRef]
  188. Shi, J.; Xie, W.; Li, S.; Wang, Y.; Wang, Q.; Li, Q. Prohibitin StPHB3 affects the browning of fresh-cut potatoes via influencing antioxidant capacity and polyphenol oxidase activation. Postharvest Biol. Technol. 2024, 207, 112598. [Google Scholar] [CrossRef]
Figure 1. The Browning mechanism and protection mechanism of vegetables.
Figure 1. The Browning mechanism and protection mechanism of vegetables.
Foods 14 02222 g001
Figure 2. Mechanism of enzymatic browning.
Figure 2. Mechanism of enzymatic browning.
Foods 14 02222 g002
Table 1. Recent innovations in physical methods for delaying vegetable browning.
Table 1. Recent innovations in physical methods for delaying vegetable browning.
MethodsApplicationsConclusionsReferences
Cold treatmentOrthogonal ultrasound-assisted freezingPotatoSample tissue structure integrity, and significant increase in phenolic content [47]
Static magnetic fields freezingPotatoSMF does not significantly affect enzymatic browning [82]
Isometric freezingSpinachThawed samples had similar cellular integrity and lysis as fresh spinach [49]
PotatoThe total phenolic content and antioxidant capacity of the samples increased and browning was delayed by 1 week [51]
Liquid nitrogen spray freezingChinese matrimony vinePPO activity was significantly reduced [83]
Heat treatmentHigh-humidity hot air impingement blanchingChili peppersReduce the residual activity (0.52%) and Browning index (7.09) of POD [57]
Hot air-assisted radio-frequency heating Red bell pepperHigher ascorbic acid retention and free radical scavenging activity [84]
Infrared hot bleachingMushroomIRHAD provides samples with the highest oxidation resistance and the lowest total color difference. [85]
Catalytic infrared heating coupled with holding time (CIRH)MushroomCIRH enhanced the maintenance of microstructure and increased the retention of total phenols [86]
Microwave ironing and bleachingAsparagusThe highest free radical scavenging capacity and total phenol content were found in samples blanched at 300 W for 4 min [87]
Vacuum-assisted blanchingZucchini sliceLess color difference with hot water rinsing, TFC reduced to 39.91%. [88]
Infrared and convection dryingLotus rootInhibits the activity of PPO and POD enzymes [89]
Thermosonication blanchingCarrot22/40 kHz downregulated POD genes expression [90]
CarrotThermal sonication at 60 °C significantly increased the hardness of the sample tissue. [91]
Lotus rootTS decreased the browning degree, POD activity and MDA content [92]
Catalytic infrared heatingCarrot90% inactivation of POD enzyme activity by short-term CIR blanching [93]
Infrared hot air dryingGarlicIRHAD retains nutrients and antioxidant activity. [94]
Bitter
Gourd
BPS-I exhibits enhanced antioxidant activity in vitro [95]
Ultrasound infrared dryingCarrotUS-IR drying does not adversely affect the color of the sample [96]
Microwave heatPotato1.1 kW microwave heating can reduce the anti-nutrient load of potato [97]
Electromagnetic field-assisted blanchingCabbageThe dielectric properties of blanched cabbage were significantly improved and color retention was increased [98]
Osmosonication dryingGingerSignificantly improves the efficiency of enzyme inactivation [99]
Gas-conditioning treatmentMapCauliflower6.06% O2; 11.43% CO2 can inhibit browning most effectively [100]
Lily10% O2, 5% CO2, 85% N2 inhibits the synthesis and accumulation of phenolics, as well as oxidative reactions [101]
Amaranth10% O2, 10% CO2 and 80% N2 are effective in maintaining chlorophyll and ascorbic acid content and antioxidant enzyme activities [102]
Cucumber2% O2, 7% CO2, and 91% N2 reduced the ability of bacterial biofilm formation [103]
Bell pepper10% O2; 45% CO2 slows down metabolism [104]
Daylily bud yieldHRW treatment maintained redox balance by inhibiting O2•− and H2O2 accumulation [105]
Microporous MAPBean podsMPPP12 (14% O2 and 4% CO2) effectively preserved chlorophyll content [106]
Mulberry leaf lettuceModified polyethylene packaging reduced the respiration intensity of the samples, and the inhibition intensity was positively correlated with the enzyme activity inhibition rate [107]
Ethanol fumigation MAPPotatoEffectively prevents the conversion of branched-chain starch to chloroplasts [108]
High-pressure treatmentHigh pressure inert gasesPotato4 MPa high pressure effectively inhibited respiration rate and biofilm oxidation [71]
Irradiation treatmentDiode laser irradiation (450 nm)PotatoInhibition of PPO and POD expression and enhancement of antioxidant activity [75]
He-Ne laser irradiationShepherd’s purse
Cauliflower
Turnip
Enhanced antioxidant, antibacterial, anti-inflammatory, and anticancer activities [109]
UV-BSpinachHigh-dose UV-B irradiation reduced leaf yellowing [110]
UV-CCarrotIncreased total phenolic content and antioxidant capacity [111]
Intensive pulsed lightMushroomLPL altered the secondary and tertiary structure of PPO leading to inactivation[79]
Shiitake mushroomAfter 25 pulses of IPL energy of 400  J, the PPO activity was reduced by 42.83% [112]
Carrot juiceThe production of hydroxyl radicals accelerates the death of bacteria [113]
Other methodsPulsed magnetic fieldCucumber Juice
Lettuce Juice
Carrot Juice
Tomato Juice
Maintain the color and flavor of vegetable juice while controlling pathogenic bacteria [114]
Tri-frequency ultrasoundCarrotPOD activity decreased by 81.43% [115]
Osmosonication GingerThe highest phenol content and antioxidant activity were obtained at 50 kHz [116]
UltrasoundCoffee leafUltrasonic regulation of phenolic metabolism [117]
Pulsed electric field and ultra-sonicationSpinach juiceThe inactivation rates of POD and PPO increased from 0.85 and 0.025 Abs/min to 0.18 and 0.011 Abs/min [118]
ElectrosprayCabbage juiceThe secondary and tertiary structures of PPO and POD are destroyed [119]
Multi-frequency ultrasoundChinese cabbageDual-frequency ultrasound showed more positive sensory properties [120]
Table 2. Recent innovations in chemical methods for delaying vegetable browning.
Table 2. Recent innovations in chemical methods for delaying vegetable browning.
MethodApplicationsConclusionsReferences
Acidifying agentsAcetic acidPotatoesDelayed ascorbic acid degradation and enhanced cell wall integrity [121]
Oxalic acidLotus RootInhibited H2O2 and O2− production and reduced POD and PPO activities [136]
Citric acidLotus rootReduced production of H2O2 and O2− while decreasing malondialdehyde, PPO and POD levels [137]
AntioxidantsCitric acidCabbageReduced residual PPO activity [138]
GlutathioneLotus rootIncreasing PAL activity to stimulate total phenol accumulation while inhibiting PPO and POD activity [139]
Flammulina velutipesIncrease soluble solids content [140]
L-arginine
L-cysteine
L-methionine
BroccoliReduced metabolism by inhibiting endogenous ethylene production led to a lower browning rate [141]
L-cysteineLotus rootConversion of phenolic compounds to cysteine adducts as competitive inhibitors of PPO [142]
Flammulina velutipesRegulates ROS metabolism and stimulates endogenous H2S production [143]
MelatoninSweet potatoInduces the expression of genes related to antioxidant pathway and reduces enzyme activity, ROS content and membrane lipid peroxidation [91]
Mushroom100 μM melatonin decreased PPO gene expression and enzyme activity [144]
Lotus seedInhibited oxidase activity and increased endogenous MT [145]
Spinach0.20 mg/mL melatonin retarded chlorophyll degradation but increased POD activity [146]
Lotus seedsReduction in membrane oxidative damage in mitochondria by stimulating antioxidant enzymes to scavenge ROS [147]
BroccoliEndogenous melatonin homeostasis led to downregulation of the expression of chlorophyll degradation-related and ethylene synthesis-related genes [148]
NacloCabbageThe synergistic effect of multi-frequency ultrasound in a sweeping mode combined with naclo is effective in sterilization [149]
Chelating agentsOxalic acidLotus rootInhibition of H2O2 and O2− production and reduction in POD and PPO activity [136]
Kojic acid derivativesPotatoCompetitive inhibitor of tyrosinase (IC50 = 3.23 ± 0.26 μM) [150]
S-Furfuryl thioacetatePotato
Eggplant
Lettuce
Changing the conformation of PPO by chelating Cu2+ [151]
PotatoDecreased PPO activity by chelating Cu2+ and acting on the residues of PPO [152]
Kojic acid-1,3,4-oxadiazole derivativesMushroomsBinds to Cu2+ in the active region, altering the secondary structure of tyrosinase [153]
Fern 6’-O-azelatePotatoThe inhibition of monophenolase was 31.4 ± 1.36% [154]
Table 3. Recent innovations in natural anti-browning method for delaying vegetable browning.
Table 3. Recent innovations in natural anti-browning method for delaying vegetable browning.
Natural AdditivesApplicationsConclusionsReferences
Γ-aminobutyric acidStem lettuceInhibition of PPO activity by delaying the expression of lsppo[7]
Sea buckthorn leaf extractPotatoCompetitive inhibitor of PPO (IC50 = 0.7 mg/mL)[155]
Chlorogenic acidPotatoRearrangement of PPO secondary structure by hydrophobic interaction[158]
4-Hydroxycinnamic acid Interaction with PPO through hydrogen bonding and hydrophobic interactions converted α-helix into β-sheet[156]
Cod peptidePotatoHigh concentration of Cod peptide decreased the total phenol content, PPO, POD and PAL activities[159]
Mulberry root bark 2-arylbenzofuran derivatives The active site interacted with Cu2+ and peroxide ions to enhance antioxidant activity and inhibit tyrosinase activity[165]
Pineapple extract
Onion extract
Pepper extract
Honey
Sweet potatoAll of them inhibited PPO, among which honey had the highest inhibition rate of 41.39–48.0%[166]
Purslane extractPotatoUltrasound-coupled extract treatment is more effective in maintaining cell membrane integrity, inhibiting PPO and POD activities, and improving antioxidant capacity[167]
Golden ginkgo tanninsLotus rootActed as a reversible mixed competitive inhibitor of tyrosinase (IC50 = 123.90 ± 0.140 μg/mL)[168]
Beef oregano extractMushroomThe greatest inhibitory effect was observed on PPO with 64.50% inhibition[169]
Citronella hydrosol
Rose hydrosol
TaroTerpenoids were effective in reducing PAL, POD and PPO activities[170]
Mango peel extractMashed potatoCompetitive inhibition of PPO (IC50 = 0.3 mg/mL)[171]
Probiotics fermentation suspensionLotus rootDecreased activities of PPO, POD and PAL, reduced TPC and soluble quinones[172]
Lotus rootSlowed down physiological responses and inhibits enzymatic browning-related enzyme activities[173]
Eucalyptus citriodora essential oilCabbageEssential oil-based Pickering emulsion maintained color, chlorophyll content[174]
BacteriophagesLettuce, Cucumber
Carrot
Improvement of organoleptic quality of vegetables by sterilization (Disrupts the cellular structural properties of bacteria)[175]
Litseacubeba essential oilCucumber juice
Carrot juice
Spinach juice
The optimal synergistic effects were found using PMF (3 times under 8 T, 60 pulses) treatments combined with 1.5 mg/ml[176]
Basil essential oilCabbageUltrasound and basil essential oil nanoemulsion disrupted the increase in intracellular ROS and extracellular MDA leading to bacterial sterilization[177]
Litsea cubeba essential oilBitter gourd Juice
Carrot juice
Cucumber juice
Spinach juice
Inhibit bacterial respiratory tract metabolism, hinder bacterial nucleic acid replication[178]
CitralCarrotUS and CLON reduced the amount of Sh. Flexneri adhering to the sample surface while retaining important quality attributes[179]
Baobab seed oilMushroomNot only improved the freshness of mushrooms but also maintained the structural stability[180]
Clove essential oilMushroomUltrasound-assisted clove essential oil nanoemulsions effectively maintained the quality characteristics of mushroom[181]
Cucumber LettuceEnhances vegetable color by killing bacteria[182]
[183]
CuminCucumber juice
Tomato juice
Cold nitrogen plasma-modified cumin aldehyde/β-cyclodextrin inclusion complexes for effective color retention in vegetable juices[184]
Bacillus velezensisEggplantThe activity of ROS scavenging enzyme was enhanced and the antioxidant capacity was improved[185]
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

Zhang, J.; Zhang, J.; Zhang, L.; Xue, Y.; Zhang, K. Mechanistic Insights into Vegetable Color Stability: Discoloration Pathways and Emerging Protective Strategies. Foods 2025, 14, 2222. https://doi.org/10.3390/foods14132222

AMA Style

Zhang J, Zhang J, Zhang L, Xue Y, Zhang K. Mechanistic Insights into Vegetable Color Stability: Discoloration Pathways and Emerging Protective Strategies. Foods. 2025; 14(13):2222. https://doi.org/10.3390/foods14132222

Chicago/Turabian Style

Zhang, Jianing, Junjun Zhang, Lidan Zhang, Yuhong Xue, and Ke Zhang. 2025. "Mechanistic Insights into Vegetable Color Stability: Discoloration Pathways and Emerging Protective Strategies" Foods 14, no. 13: 2222. https://doi.org/10.3390/foods14132222

APA Style

Zhang, J., Zhang, J., Zhang, L., Xue, Y., & Zhang, K. (2025). Mechanistic Insights into Vegetable Color Stability: Discoloration Pathways and Emerging Protective Strategies. Foods, 14(13), 2222. https://doi.org/10.3390/foods14132222

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

Article Metrics

Back to TopTop