Next Article in Journal
Change in Sunflower Oil Quality and Safety Depending on Number of Deodorisation Cycles Used
Previous Article in Journal
Development of Functional Foods: A Comparative Study on the Polyphenols and Anthocyanins Content in Chokeberry and Blueberry Pomace Extracts and Their Antitumor Properties
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Optimization of Pectin Extraction from Melon Peel as a New Source of Pectin and Pectin Hydrolysate with Prebiotic Potential

1
Environmental Biotechnology Laboratory, Faculty of Agro-Industry, Prince of Songkla University, Hat Yai, Songkhla 90110, Thailand
2
Center of Excellence of Microbial Diversity and Sustainable Utilization, Faculty of Science, Chiang Mai University, Chiang Mai 50200, Thailand
3
Center of Excellence in Innovative Biotechnology for Sustainable Utilization of Bioresources, Faculty of Agro-Industry, Prince of Songkla University, Hat Yai, Songkhla 90110, Thailand
4
Graduate School of Environmental and Life Sciences, Okayama University, 1-1-1 Thshimanaka, Kita-ku, Okayama 700-8530, Japan
*
Author to whom correspondence should be addressed.
Foods 2024, 13(16), 2554; https://doi.org/10.3390/foods13162554
Submission received: 18 July 2024 / Revised: 8 August 2024 / Accepted: 9 August 2024 / Published: 16 August 2024

Abstract

:
Food wastes have a large number of functional ingredients that have potential for valorization. Melon peels are increasingly produced as waste in food industries in Thailand. This study aimed to optimize pectin extraction conditions from melon peel for its prebiotic potential. Optimization was conducted using a response surface methodology and Box–Behnken experimental design. An analysis of variance indicated a significant interaction between the extraction conditions on extraction yield and degree of esterification (DE). These include pH and solvent-to-sample ratio. The conditions for the extraction of pectin with low DE (LDP), medium DE (MDP) and high DE (HDP) were optimized. Pectin hydrolysate from LDP, MDP and HDP was prepared by enzymatic hydrolysis into LPEH, MPEH and HPEH, respectively. LDP, MDP, HDP, LPEH, MPEH and HPEH were compared for their efficiency in terms of the growth of three probiotic strains, namely Lactobacillus plantarum TISTR 877, Lactobacillus casei TISTR 390 and Enterococcus faecium TISTR 1027. Among the samples tested, HPEH showed the highest ability as a carbon source to promote the growth and prebiotic activity score for these three probiotic strains. This study suggests that melon peel waste from agro-industry can be a novel source for prebiotic production.

Graphical Abstract

1. Introduction

Currently, people’s concerns on health and quality of life are increasing. The demand for food ingredients from natural products with functional and health properties has increased. Prebiotics are famous items due to their benefits on the promotion of human gut microbiota [1]. Prebiotics include dietary fibers that humans cannot digest and absorb in the digestive tract, although they can be degraded by gut the microflora [2]. Prebiotics include fructo-oligosaccharides, isomalto-oligosaccharides, lactilol, lactosucrose, lactulose and inulin. Pectin and pectin-derived oligosacharides (POSs) are a new source of prebiotics since they have been reported for their prebiotic properties and improve the growth of potential probiotics such as Lactobacillus spp. Bifidobacterium faecali, Bacterium prausnitzii and Roseburia spp. [3,4,5,6]. Several studies have demonstrated diverse effects, such as apoptosis in human colonic adenocarcinoma cells [7,8], their potential ability to protect heart health, mitigation of heavy metal-induced damage, anti-obesity impacts, antibacterial traits, and antioxidant properties [9].
The pectin extracted from fruit peel biomass displays potential as a promising source for producing pectic oligosaccharides or pectin hydrolysates. These forms demonstrate capabilities in aiding probiotic growth and presenting prebiotic potential [1,2,3]. Pectinase enzymes can hydrolyze pectin into monosaccharides or oligosaccharides, which can be used as carbon sources for probiotics [2]. In a previous study by Chatterjee et al. [10], pectin was extracted separately from the peels of various fruits, including orange, tomato, banana, guava and watermelon, to evaluate the prebiotic properties of the pectin derived from each fruit. The findings showed that tomato peel-derived pectin exhibited the highest growth stimulation for Lactobacillus casei [10]. POSs from mango peel were evaluated for their prebiotic effects on Lactobacillus reuteri DSM 17938 and Bifidobacterium animalis TISTR 2195. The results demonstrated valuable scores of prebiotic activities for both B. animalis TISTR 2195 (7.76) and L. reuteri DSM 17938 (6.87) [11]. Considering the degree of esterification (DE) of pectin between low-methoxyl pectin (LMP; DE < 50%) and high-methoxyl pectin (HMP; DE ≥ 50%), Olano-Martin et al. [12] reported that LMP has a higher prebiotic index (PI) than HMP. An evaluation of the prebiotic potential of POSs with a low or high DE of pectin from orange peel found that the PIs of low-DE POSs and high-DE POSs were 0.12 and 0.08, respectively [12].
The utilization of agricultural byproducts and wastes as sources of prebiotics is a useful way to valorize and eliminate waste. Therefore, many attempts have been made with various byproducts and wastes to comply with the government policies on zero-waste production. However, none of those attempts have investigated the prebiotic properties of pectin and pectic oligosaccharide components from melon peel. The production of melon in Thailand is popular nowadays. The average annual yield was 984,100 kg in 2018 [13]. During fruit processing, melon peels constitute approximately 20–30% of the total fresh weight, serving as byproducts that are abundant in pectin and dietary fiber [14,15]. According to Hartati and Subekti [16], melon peels are composed of approximately 20% cellulose, 23% hemicellulose, 10% lignin and 13% pectin. The extraction process for pectin involves various factors such as pH, temperature, time and solid-to-solvent ratio (SSR), which influence extract quality parameters like extraction yield and properties including methoxyl content (MeO), DE, and anhydro-galacturonic acid content (AUA), resulting in differences in prebiotic potentials between pectin and pectin hydrolysate [12,17]. Therefore, it is imperative to optimize the extraction variables and evaluate the prebiotic potentials of pectin and pectin hydrolysate.
This study aimed to optimize pectin extraction from melon peel for the production of pectin and pectin hydrolysate with prebiotic potential. This research introduces an approach to enhancing the value of byproducts sourced from Thai-fruit industrial wastes and suggests a framework conducive to sustainable industrial development.

2. Materials and Methods

2.1. Preparation of Melon Peel

The commercial Hami melon cultivars (Cucumis melo var. saccharinus) utilized in this research were obtained from a nearby fruit market. To ensure homogeneity, all samples were collected in the same week of the year. After collection, approximately 10 kg of the fruit peels was removed, cleaned, and dried in an oven at 60 °C for 48 h. Then, the dried peels were ground into powder by grinder and sieved through a 40-mesh sieve. The processed powder was stored in a dry environment before being used for experimental analyses [18].

2.2. Extraction of Pectin from Melon Peel

The pectin extraction from melon peel was conducted in accordance with the methodology described by Khan et al. [19]. Pectin was extracted under varied conditions: temperatures of 75, 85, and 95 °C, extraction times of 60, 120, and 180 min, pH levels at 1.0, 2.0, and 3.0, and different solid-to-solvent ratios (SSR) of 10, 30, and 50 v/w. Following the filtration process with two layers of gauze, the pectin extract underwent centrifugation at 8000× g for 15 min to remove residual particles. Subsequently, the resulting supernatant was precipitated using a 95% ethanol solution (at a ratio of 1:2% v/v) for 24 h at 4 °C. The polysaccharides underwent two ethanol washes. Finally, the resultant precipitated pectin was vacuum-dried at 60 °C until reaching a constant weight. The extraction yield was then determined using the following formula:
%   Yield   = weight   of   dried   pectin   extract   ( g ) weight   of   dried   peel   powder g   ×   100
The determination of DE followed the titrimetric approach by Raji et al. [18]. Initially, 0.2 g of dried pectin extract was amalgamated with ethanol and then dissolved in 20 mL of deionized water at 40 °C. Following the thorough dissolution of the pectin extract, five drops of phenolphthalein were introduced into the solution. After dissolving the pectin extract, five drops of phenolphthalein were added. Following this, the solution was subjected to titration utilizing 0.5 M NaOH, and the volume of NaOH solution causing the color change was recorded as the initial titer (v1). Subsequently, 10 mL of 0.5 M NaOH was added, and intense agitation of the solution for 15 min was initiated; then 10 mL of 0.5 M HCl was introduced, and the solution was stirred until the pink color disappeared. Titration with 0.5 M NaOH was performed, and the volume used was recorded as the saponification titer or final titer (v2). DE was calculated based on the following formula:
%   DE   = v 2 v 1 + v 2 ×   100

2.3. Optimization of the Pectin Extraction Conditions of Melon Peel Using the Response Surface Methodology (RSM)

The RSM was used to anticipate the optimal pectin extraction conditions for melon peel, which were yield and DE. The optimization process was carried out following Raji et al. [18] and Prakash Maran et.al. [20] with a slight modification. Twenty-nine hydrolysis trials were randomly run per Box–Behnken (BBD) experimental design with independent variables, including temperature (X1: 75, 85, 95 °C), extraction time (X2: 60, 120, 180 min), pH (X3: 1, 2, 3) and SSR (X4: 10, 30, 50 mL/g) employed at three equally spaced tiers (−1, 0, +1). The conditions for the extraction of pectin with a low (LDP), medium (MDP) and high (HDP) degree of esterification were optimized. Pectin hydrolysate from LDP, MDP and HDP was subjected to enzymatic hydrolysis; the hydrolysates from LDP, MDP and HDP were LPEH, MPEH and HPEH, respectively.

2.4. Properties of Pectin Extract from Melon Peel

Pectin from the models (LDP, MDP and HDP) was characterized by some properties such as equivalent weight (Eq.wt), MeO, DE⁠ and AUA using a modification designed by Mohamed [21].

2.5. Prebiotic Properties of Extracted Pectin and Pectin Hydrolysate (PEH)

2.5.1. Preparation of Pectin and Pectin Hydrolysate from Melon Peel

Pectin samples from melon peel extraction in optimized conditions with three levels of DE, low (DE 40–50%), medium (DE 60–70%) and high (DE 80–90%), were used to study prebiotic potentials. PEH was prepared through the hydrolysis of pectin with pectinase. The pectin extract at 2% w/v was used to react with commercial pectinase (30,000 U, Reach biotechnology, Thailand) at pH 4 and 45 °C for 6 h in an incubator shaker. Then, the reaction was stopped by boiling for 10 min. The hydrolyzed biomass was centrifuged at 5000× g for 15 min. The three levels of PEH were named LPEH (PEH hydrolyzed from pectin with 40–50% DE), MPEH (PEH hydrolyzed from pectin with 60–70% DE) and HPEH (PEH hydrolyzed from pectin with 80–90% DE).

2.5.2. Preparation of Probiotics

Three strains of probiotic microorganisms were used for testing the prebiotic properties: L. plantarum TISTR 877, L. casei TISTR 390 and E. faecium TISTR 1027. Microbial strains were cultured in MRS broth and incubated at 37 °C for 48 h. The amount of each microorganism after incubation at 106 CFU/mL was used for further testing of the prebiotic properties of pectin and PEH.

2.5.3. Effect of the DE of Pectin and PEH on the Growth of Probiotics

Probiotic cultures, comprising 0.1 mL with 106 CFU/mL, were added to 10 mL of MRS broth containing 2% pectin or 2% PEH and incubated at 37 °C for 0, 12, 24 and 48 h. After inoculation in the test medium, the final amount was 104 CFU/mL. Then, serial dilution and pour plate on MRS agar were conducted to count viable cells as CFU/mL. The controls included a positive control (with glucose) and a negative control (without glucose).

2.5.4. Evaluation of Prebiotic Activity Score (PAS)

The assessment of prebiotic activity was conducted following a method previously reported [22]. L. plantarum TISTR 877, L. casei TISTR 390 and E. faecium TISTR 1027 were represented as probiotic cultures (P), while Escherichia coli DMST 4212 was chosen to represent the enteric species (E). The experiment was conducted by measuring the number of CFU/mL prior (P0 and E0) and subsequent (P24 and E24) to incubation for 24 h at 37 °C on 2% (w/v) glucose (PG and EG), with 2% (w/v) of the test pectin or pectin hydrolysate as the prebiotic (Px and Ex) as reported in [23]. The experiment was performed three times for each assay. Subsequently, the prebiotic activity score was determined using Equation (3):
Prebiotic   activity   score   ( PAS ) = [ ( L o g P x 24 L o g P x 0 ) / ( L o g P G 24 l o g P G 0 ) ] [ ( L o g E x 24 L o g E x 0 ) / ( L o g E G 24 l o g E G 0 ) ] .

2.6. Statistical Analysis

The optimization of the pectin extraction conditions was performed using RSM Design-Expert 7.0 software (Stat-Ease 2005). Each analysis was represented by mean values (±SD). An ANOVA analysis through Minitab 14.0 was conducted to determine potential significant differences among the treatments.

3. Results

3.1. Optimization of Pectin Extraction from Melon Peel

3.1.1. Model Fitting and Statistical Analysis

The RSM was applied to enhance the conditions for the extraction of pectin from melon peel. Data from 29 experimental trials conducted using central composite design (CCD) were obtained, involving four distinct independent factors: temperature (X1), extraction time (X2), pH (X3) and solvent-to-sample ratio (X4). In this study, the summary model indicated the importance of a two-factor interaction model, offering valuable insights for pectin yield and DE. The predicted model for pectin yield and DE was a quadratic model. The pectin yield of melon peel obtained from 29 experimental trials ranged from 2.77 to 22.44%, which was similar to the pectin yield obtained from orange peels by Marin et al. [24] but higher than that from the yield of passion fruit peel (2.25–14.60%) [25]. Khan et al. [26], Hashmi et al. [27] and Dehankar et al. [28] have reported the highest pectin yields of 20%, 20.12% and 21%, respectively. Meanwhile, the DE of extracted pectin ranged from 41.96 to 91.80%, which was broader than the extraction of passion fruit peel (41.67–67.31%) [25]. The quantity, quality and characteristics of pectin differ depending on plant origins, extraction techniques and parameters. [29]. The following are the second-order equations associated with the extraction yield and DE:
pectin yield (%) = 20.49460 − 1.04076X1 + 0.09640X2 + 23.98358X3 + 0.65615X4 + 0.00714X12 − 0.000355833 X22 − 4.58975X32 − 0.00638688X42 + 0.0003625X1X2 − 0.083750X1X3 + 0.0019625X1X4 − 0.015542X2X3 − 0.0000645833X2X4 − 0.13375X3X4
de (%) = 149.86462 − 2.23924X1 + 0.26224X2 − 48.02965X3 − 0.29087X4 + 0.018179X12 + 0.000640764X22 + 17.03438X32 + 0.00867173X42 − 0.0043934X1X2 − 0.022362X1X3 − 0.00445979X1X4 − 0.010865X2X3 − 0.000684182X2X4 + 0.15893X3X4
where Xi represents the coded independent factor (X1 = temperature, X2 = extraction time, X3 = pH, X4= solvent-to-sample ratio).
In our investigation, an analysis of variance (ANOVA) was employed to validate the significance of the models, as shown in Table 1. Significantly low p-values for both yield and DE (less than 0.001) and an insignificant lack of fit (higher than 0.05) indicate the suitability of the model in representing the responses for these parameters [30]. On the other hand, the adequacy of the models was examined by evaluating the determination coefficients for yield and DE [31,32] with the value of the determination coefficient (95.70% and 98.44%, respectively), the adjusted determination coefficient (91.40% and 96.88%, respectively) and the predicted determination coefficient (76.50% and 92.21%, respectively). The significant R2 and Adj-R2 values attest to the high level of precision and reliability of the conducted experiments [33].

3.1.2. The Relationship between Factor Variables and Extraction Yield

The experiments were conducted over different pH levels (1–3), and the outcomes are depicted in Figure 1C,D,F. The results notably demonstrate an increase in the extraction yield of pectin as the pH levels decrease. The acidic solvent at a lower pH effectively solubilized insoluble pectin, converting it into a soluble form and consequently enhancing the extraction yield of pectin [18]. It should also be noted that at high pH levels, pectin accumulates, preventing its release and thereby lowering the extraction yield of pectin [34]. The results of this study align with research on extracting pectin from apple pomace, sugar beet pulp, pumpkin and pomegranate peel [35,36,37,38]. Temperature is considered one of the critical factors influencing pectin extraction yield. The findings demonstrate an increase in pectin yield as temperature rises (Figure 1A,D,E). Increased temperature enhances solvent solubility and penetration into plant tissue, leading to an increased pectin extraction yield [39]. This outcome was consistent with pectin extraction from peach pomace [40] and melon peel [18]. Regarding the impact of extraction time, longer extraction times boosted pectin yield (Figure 1B,C,E). This observation aligned with those previously reported by Chen et al. [41] and Zheng et al. [42]. Also, a high value of the solvent-to-sample ratio enhances pectin yield from melon peel extraction (Figure 1A,B,F), possibly due to the enlarged contact surface area between the melon peels and the solvent [33].

3.1.3. The Relationship between Factor Variables and DE

An increase in the DE was observed as pH and the SSR were increased while extraction time and temperature were decreased, as illustrated in Figure 2. This occurrence is attributed to polygalacturonic chain de-esterification occurring under harsh conditions such as high temperature, extended duration and low pH [43]. Similar patterns were observed in pectin extraction from dragon fruit peel by Tang et al. [44] and from melon peel by Raji et al. [18]. In addition, the values of DE reduced with a decrease in extraction temperature, a decrease in time, an increase in the SSR and a decrease in the pH value. The pH of the extraction solution and the SSR significantly influenced the DE, consistent with previous findings [45]. Significant interactive effects included a linear pH effect (p < 0.0001), a cubic pH effect (p < 0.0001), an SSR effect (p = 0.0096) and a quadratic pH-SSR effect (p = 0.0488) on the DE. Similar findings were reported by Li et al. [45].

3.2. Optimization Conditions of Factor Variables for the Yield and DE

To explore the impact of varying pectin esterification levels on prebiotic properties, the factor variables were optimized to maximize pectin yield at low, medium and high DE levels by solving Equations (4) and (5). The optimal conditions for the highest yield for the three levels of methylated pectin are shown in Table 2. Also, all of this methylated pectin was characterized in terms of equivalent weight (Eq.wt), methoxyl content (MeO) and anhydrouronic acid content (AUA) (Table 2). Based on the model, three conditions of pectin extraction that produced LDP, MDP and HDP were optimized to compare their prebiotic properties. The optimal extraction conditions for the LDP were identified as an 85 °C temperature, a 166 min duration, a pH of 1, and an SSR of 50 mL/g. The MDP displayed optimal conditions at an 85 °C temperature, a 120 min duration, a pH of 2.2, and an SSR of 50 mL/g, while the HDP exhibited ideal conditions at an 85 °C temperature, a 90 min duration, a pH of 2.7, and an SSR of 50 mL/g. The corresponding responses of the LDP, MDP and HDP were 46.17%, 56.65% and 88.38% DE, respectively. From Table 2, the extraction temperature of pectin was specified for a maximum of 85 °C to reduce the energy consumption required for extraction for model optimization. The optimal temperature for the effect of LDP, MDP and HDP extraction on pectin yield was the same at 85 °C, which is the maximum temperature condition. It was also demonstrated by Chan and Choo [17] and Vriesmann et al. [46] that increasing the temperature significantly enhanced the extraction of pectin from cocoa husks with the assistance of citric acid. The heated conditions facilitated the dissolution of pectin and other pectic substances enclosed in the cell wall (protopectin), thereby elevating the overall pectin yield. The optimal SSR for the effect of LDP, MDP and HDP extraction on pectin yield was the same at 50 mL/g, which is the maximum. Increasing the temperature and SSR showed an increase in pectin yield and %DE. The DE value decreased with increasing extraction time and decreasing extraction pH, consistent with a report by Li et al. [45], which emphasized the significant impact of pH and time on the DE of mango peel pectin, whereas temperature showed no considerable effect. To achieve both maximum yield and DE, the extraction conditions should involve shorter extraction periods with a high pH value. In the case that the highest yield but lowest DE were required, the conditions for extraction should be a higher extraction duration with a low pH. Since a lower extraction duration with a high pH were mild conditions, some of the methyl group still remained in the galacturonic acid residues. The yield of pectin under the conditions used to extract HDP was lower than the yield of pectin in the conditions of LDP extraction, but the purity of the HDP was better (78.43% of AUA) than that of the LDP (13.05% of AUA) when the temperature and SSR were the same. Finally, the three pectin types LDP, MDP and HDP were used to study prebiotic properties.

3.3. Prebiotic Properties of Extracted Pectin and PEH on Growth of Probiotics

The growth of three probiotic strains, namely E. faecium TISTR 1027, L. plantarum TISTR 877 and L. casei TISTR 390, in the glucose-free MRS media, containing three types of pectin (LDP, MDP and HDP) or PEHs (LPEH, MPEH and HPEH), is shown in Table 3. During the early stages of incubation, the growth of L. casei TISTR 390 was promoted by HDP, LPEH, MPEH and HPEH, but only pectin hydrolysate promoted its growth when the incubation time increased to 36 h, similar to the results for E. faecium TISTR 1027. In contrast, the growth of L. plantarum TISTR 877 was only promoted by PEHs at the beginning of incubation, but both pectin and pectin enzyme hydrolysate promoted their growth after 36 h of incubation time. The fact that the probiotic counts were lower in the media containing pectin at the beginning of incubation might result from the acidity following the pattern of LDP > MDP > HDP. The presence of HPEH in the culture media as a carbon source resulted in the highest growth for all three probiotic strains. Despite glucose in the MRS media displaying comparable growth, it is crucial to recognize that glucose is absorbable and provides calories. On the other hand, pectin oligosaccharides in HPEH cannot be absorbed by the body, offering no caloric value [2]. Figure 3 illustrates the prebiotic activity scores of pectin and pectin hydrolysates from melon peel for L. casei TISTR 390, L. plantarum TISTR 877 and E. faecium TISTR 1027, calculated from cfu/mL data after 12 and 24 h of incubation using Equation (3). L. plantarum TISTR 877 exhibited higher prebiotic activity when grown in MRS broth with added PEH compared to pectin, and the highest score was recorded for HPEH (1.10). After 12 h of fermentation, HPEH exhibited prebiotic scores 4.6-fold and 5.2-fold greater than inulin for L. plantarum TISTR 877 and E. faecium TISTR 1027, respectively. Furthermore, the prebiotic activity score for HPEH obtained from melon peel was higher than the prebiotic activity score of citrus-derived POSs for L. paracasei LPC-37 (0.41) and Bifidobacterium bifidum ATCC 29521 (0.92) [5]. The pectin enzyme hydrolysates demonstrate significant prebiotic potential, especially for L. plantarum TISTR 877, as pectinase is expected to break down pectin into galacturonic acid and pectic oligosaccharides [47].
Our results clearly showed that PEH from melon pectin offers growth-enhancing ability for the tested probiotics. The data support the prebiotic potential of POSs or pectin hydrolysate from citrus peel [2,5] and sugar beet pulp [9]. Thus, PEH is a highly beneficial component for food supplements, offering prebiotic properties.

4. Conclusions

In conclusion, both the yield and the degree of esterification of pectin were significantly affected by the pH value and SSR used in the extraction process. These conditions encompassed an extraction temperature of 85 °C, an SSR of 50 mL/g for 166 min and a pH of 1. Both pectin and pectin hydrolysate from melon peel have the potential to enhance the growth of Lactobacillus plantarum TISTR 877, Lactobacillus casei TISTR 390 and Enterococcus faecium TISTR 1027. The prebiotic activity score of HPEH was the highest. This is the first time the feasibility of a high value-added product such as prebiotics from melon peel has been shown in the field of utilizing fruit industry wastes.

Author Contributions

Conceptualization, U.T., S.B. and S.S.; methodology, S.B.; formal analysis, S.B.; resources, U.T.; data curation, S.B. and U.T.; writing original draft, S.B. and U.T.; writing—review and editing, U.T., F.K. and B.C.; visualization, U.T.; supervision, U.T. and A.U.; funding acquisition, U.T. All authors have read and agreed to the published version of the manuscript.

Funding

This research was supported by the National Science, Research and Innovation Fund (NSRF), Prince of Songkla University (Grant No. AGR6505051j) and Thailand Research Fund (Grant No. RTA6280014).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

The original contributions presented in the study are included in the article, further inquiries can be directed to the corresponding author.

Acknowledgments

The authors gratefully acknowledge the support of the National Science, Research and Innovation Fund (NSRF) and Prince of Songkla University (Grant No. AGR6505051j) as well as the Thailand Research Fund (Grant No. RTA6280014). Additionally, we thank KritSamai Plantation, Songkhla, Thailand, for kindly providing materials for the experiment.

Conflicts of Interest

The authors declare no conflicts of interest.

References

  1. Davani-Davari, D.; Negahdaripour, M.; Karimzadeh, I.; Seifan, M.; Mohkam, M.; Masoumi, S.J.; Berenjian, A.; Ghasemi, Y. Prebiotics: Definition, types, sources, mechanisms, and cinical applications. Food 2019, 8, 92. [Google Scholar] [CrossRef]
  2. Ho, Y.-Y.; Lin, C.-M.; Wu, M.-C. Evaluation of the prebiotic effects of citrus pectin hydrolysate. J. Food Drug Anal. 2017, 25, 550–558. [Google Scholar] [CrossRef]
  3. Gomez, B.; Pelaez, C.; Martinez-cuesta, M.C.; Parajo, J.C.; Alonso, J.L.; Requena, T. Emerging prebiotics obtained from lemon and sugar beet byproducts: Evaluation of their in vitro fermentability by probiotic bacteria. LWT Food Sci. Technol. 2019, 109, 17–25. [Google Scholar] [CrossRef]
  4. Larsen, N.; de Souza, C.B.; Krych, L.; Cahu, T.B.; Wiese, M.; Kot, W.; Hansen, K.M.; Blennow, A.; Venema, K.; Jespersen, L. Potential of Pectins to beneficially modulate the gut microbiota depends on their structural properties. Front. Microbiol. 2019, 10, 223. [Google Scholar] [CrossRef]
  5. Zhang, S.; Hu, H.; Wang, L.; Liu, F.; Pan, S. Preparation and prebiotic potential of pectin oligosaccharides obtained from citrus peel pectin. Food Chem. 2018, 244, 232–237. [Google Scholar] [CrossRef]
  6. Pradi, B.; Baldassarre, S.; Babbar, N.; Elena, B.; Vandezande, P.; Hermans, D.; Bruggeman, G.; Gatti, M.; Elst, K.; Sforza, S. Pectin oligosaccharides from sugar beet pulp: Molecular characterization and potential prebiotic activity. Food Funct. 2018, 10, 223. [Google Scholar] [CrossRef]
  7. Huang, Y.; Wang, T. Pectin Oligosaccharides enhance α2,6-sialylation modification that promotes apoptosis of bladder cancer cells by targeting the hedgehog pathway. Cell Biochem. Biophys. 2021, 79, 719–728. [Google Scholar] [CrossRef]
  8. Tan, H.; Chen, W.; Liu, Q.; Yang, G.; Li, K. Pectin oligosaccharides ameliorate colon cancer by regulating oxidative stress- and inflammation-activated signaling pathways. Front. Immunol. 2018, 9, 1504. [Google Scholar] [CrossRef]
  9. Gomez, B.; Gullon, B.; Yanez, R.; Schols, H.; Alonso, J.L. Prebiotic potential of pectins and pectic oligosaccharides derived from lemon peel wastes and sugar beet pulp: A comparative evaluation. J. Funct. Foods 2016, 20, 108–121. [Google Scholar] [CrossRef]
  10. Chatterjee, E.; Manuel, G.A.S.; Hassan, S. Effect of fruit pectin on growth of lactic acid bacteria. J. Prob. Health 2016, 4, 147–150. [Google Scholar] [CrossRef]
  11. Wongkaew, M.; Tinpovong, B.; Sringarm, K.; Leksawasdi, N.; Jantanasakulwong, K.; Rachtanapun, P.; Hanmoungjai, P.; Sommano, S.R. Crude pectic oligosaccharide recovery from Thai Chok Anan mango peel using pectinolytic enzyme hydrolysis. Foods 2021, 10, 627. [Google Scholar] [CrossRef] [PubMed]
  12. Olano-Martin, E.; Gibson, G.R.; Rastall, R.A. Comparison of the in vitro bifidogenic properties of pectins and pectic-oligosaccharides. J. Appl. Microbiol. 2002, 93, 505–511. [Google Scholar] [CrossRef] [PubMed]
  13. Department of Agriculture Extension, 2017, Cantaloupe in 2016. Available online: http://www.agriinfo.doae.go.th/year60/plant/rortor/veget/18.pdf (accessed on 12 January 2019).
  14. Mallek-Ayadi, S.; Bahloul, N.; Kechaou, N. Characterization, phenolic compounds and functional properties of Cucumis melo L. peels. Food Chem. 2017, 221, 1691–1697. [Google Scholar] [CrossRef] [PubMed]
  15. Gomez-Garcia, R.; Camposa, D.A.; Oliveiraa, A.; Aguilarb, C.N.; Madureiraa, A.R.; Pintado, M. A chemical valorisation of melon peels towards functional food ingredients: Bioactives profile and antioxidant properties. Food Chem. 2021, 335, 1–11. [Google Scholar] [CrossRef] [PubMed]
  16. Gomez-Garcia, R.; Sanchez-Gutierrez, M.; Freitas-Costa, C.; Vilas-Boas, A.A.; Campos, D.A.; Aguilar, C.N.; Madureira, A.R.; Pintado, M. Prebiotic effect, bioactive compounds and antioxidant capacity of melon peel (Cucumis melo L. inodorus) flour subjected to in vitro gastrointestinal digestion and human faecal fermentation. Food Res. Int. 2022, 154, 1–13. [Google Scholar] [CrossRef] [PubMed]
  17. Chan, S.Y.; Choo, W.S. Effect of extraction conditions on the yield and chemical properties of pectin from cocoa husks. Food Chem. 2013, 141, 3752–3758. [Google Scholar] [CrossRef] [PubMed]
  18. Raji, Z.; Khodaiyan, F.; Rezaei, K.; Kiani, H.; Hosseini, S.S. Extraction optimization and physicochemical properties of pectin from melon peel. Int. J. Biol. Macromol. 2017, 98, 709–716. [Google Scholar] [CrossRef] [PubMed]
  19. Khan, A.A.; Butt, M.S.; Randhawa, M.A.; Karim, R.; Sultan, M.T.; Ahmed, W. Extraction and characterization of pectin from grapefruit (Duncan cultivar) and its utilization as gelling agent. Int. Food Res. J. 2014, 21, 2195–2199. [Google Scholar]
  20. Prakash Maran, J.; Manikandan, S.; Thirugnanasambandham, K.; Vigna Nivetha, C.; Dinesh, R. Box-Behnken design based statistical modeling for ultrasound-assisted extraction of corn silk polysaccharide. Carbohydr. Polym. 2013, 92, 604–611. [Google Scholar] [CrossRef]
  21. Mohamed, H. Extraction and characterization of pectin from grapefruit peels. MOJ Food Process. Technol. 2016, 2, 31–38. [Google Scholar] [CrossRef]
  22. Thitiratsakul, B.; Anprung, P. Prebiotic activity score and bioactive compounds in longan (Dimocarpus longan Lour.): Influence of pectinase in enzyme-assisted extraction. J. Food Sci. Technol. 2014, 51, 1947–1955. [Google Scholar] [CrossRef] [PubMed]
  23. Anprung, P.; Sangthawan, S. Prebiotic activity and bioactive compounds of the enzymatically depolymerized Thailand-grown mangosteen aril. J. Food Res. 2012, 1, 268–276. [Google Scholar] [CrossRef]
  24. Marin, F.R.; Soler-Rivas, C.; Benavente-Garcia, O.; Castillo, J.; Perez-Alvarez, J.A. By-products from different citrus processes as a source of customized functional fibres. Food Chem. 2007, 100, 736–741. [Google Scholar] [CrossRef]
  25. Liew, S.O.; Chin, N.L.; Yusof, Y.A. Extraction and characterization of pectin from passion fruit peels. Agric. Agric. Sci. Procedia. 2014, 2, 231–236. [Google Scholar] [CrossRef]
  26. Khan, M.; Bibi, N.; Zeb, A. Optimization of process conditions for pectin extraction from citrus peel. Sci. Tech. Dev. 2015, 34, 9–15. [Google Scholar] [CrossRef]
  27. Hashmi, S.H.; Ghatge, P.; Machewad, G.M.; Pawar, S. Studies on extraction of essential oil and pectin from sweet orange. J. Food Proc. Tech. 2012, 1, 1–3. [Google Scholar] [CrossRef]
  28. Dehankar, P.B.; Patil, N.S.; Mane, P.; Mane, J. Recycling of orange peel as pectin—A valuable product. Int. J. Inf. Fut. Res. 2015, 2, 3041–3048. [Google Scholar]
  29. Hundie, K.B.; Akuma, D.A. Extraction and Characterization of pectin from lemon waste for commercial applications. J. Text. Cloth. Sci. Appl. 2021, 8, 1111–1120. [Google Scholar]
  30. Homayoonfal, M.; Khodaiyan, F.; Mousavi, M. Modelling and optimizing physicochemical features of walnut-oil beverage emulsions by implementation of response surface methodology: Effect of preparation conditions on emulsion stability. Food Chem. 2015, 174, 649–659. [Google Scholar] [CrossRef] [PubMed]
  31. Balasubramani, P.; Palaniswamy, P.T.; Visvanathan, R.; Thirupathi, V.; Subbarayan, A.; Prakash Maran, J. Microencapsulation of garlic oleoresin using maltodextrin as wall material by spray drying technology. Int. J. Biol. Macromol. 2015, 72, 210–217. [Google Scholar] [CrossRef]
  32. Chaharbaghi, E.; Khodaiyan, F.; Hosseini, S.S. Optimization of pectin extraction from pistachio green hull as a new source. Carbohydr. Polym. 2017, 173, 107–113. [Google Scholar] [CrossRef] [PubMed]
  33. Prakash Maran, J.; Sivakumar, V.; Thirugnanasambandham, K.; Sridhar, R. Optimization of microwave assisted extraction of pectin from orange peel. Carbohydr. Polym. 2013, 97, 703–709. [Google Scholar] [CrossRef]
  34. Shivamathi, C.S.; Gunaseelan, S.; Soosai, M.R.; Vignesh, N.S.; Varalakshmi, P.; Kumar, R.S.; Ganesh Moorthy, I.M. Process optimization and characterization of pectin derived from underexploited pineapple peel biowaste as a value-added product. Food Hydrocoll. 2022, 123, 107141. [Google Scholar] [CrossRef]
  35. Canteri-Schemin, M.H.; Fertonani, H.C.R.; Waszczynskyj, N.; Wosiacki, G. Extraction of pectin from apple pomace. Braz. Arch. Biol. Technol. 2005, 48, 259–266. [Google Scholar] [CrossRef]
  36. Yapo, B.M.; Robert, C.; Etienne, I.; Wathelet, B.; Paquot, M. Effect of extraction conditions on the yield, purity and surface properties of sugar beet pulp pectin extracts. Food Chem. 2007, 100, 1356–1364. [Google Scholar] [CrossRef]
  37. Prakash Maran, J.; Mekala, V.; Manikandan, S. Modeling and optimization of ultrasound-assisted extraction of polysaccharide from Cucurbita moschata. Carbohydr. Polym. 2013, 92, 2018–2026. [Google Scholar] [CrossRef]
  38. Moorthy, I.G.; Maran, J.P.; Surya, S.M.; Naganyashree, S.; Shivamathi, C.S. Response surface optimization of ultrasound assisted extraction of pectin from pomegranate peel. Int. J. Biol. Macromol. 2015, 72, 1323–1328. [Google Scholar] [CrossRef] [PubMed]
  39. Yang, Z.; Zhai, W. Optimization of microwave-assisted extraction of anthocyanins from purple corn (Zea mays L.) cob and identification with HPLC-MS. Innov. Food Sci. Emerg. Technol. 2010, 11, 470–476. [Google Scholar] [CrossRef]
  40. Pagan, J.; Ibarz, A.; Llorca, M.; Pagan, A.; Barbosa-Canovas, G.V. Extraction and characterization of pectin from stored peach pomace. Food Res. Int. 2001, 34, 605–612. [Google Scholar] [CrossRef]
  41. Chen, Y.; Xie, M.Y.; Gong, X.F. Microwave-assisted extraction used for the isolation of total triterpenoid saponins from Ganoderma atrum. J. Food Eng. 2007, 81, 162–170. [Google Scholar] [CrossRef]
  42. Zheng, X.; Yin, F.; Liu, C.; Xu, X. Effect of process parameters of microwave assisted extraction (MAE) on polysaccharides yield from pumpkin. J. Northeast Agric. Univ. 2011, 18, 79–86. [Google Scholar] [CrossRef]
  43. Hosseini, S.S.; Khodaiyan, F.; Yarmand, M.S. Aqueous extraction of pectin from sour orange peel and its preliminary physicochemical properties. Int. J. Biol. Macromol. 2016, 82, 920–926. [Google Scholar] [CrossRef] [PubMed]
  44. Yang, P.Y.; Wong, C.J.; Woo, K.K. Optimization of pectin extraction from peel of dragon fruit (Hylocereus polyrhizus). Asian J. Biol. Sci. 2011, 4, 189–195. [Google Scholar]
  45. Li, W.J.; Fan, Z.G.; Wu, Y.Y.; Jiang, Z.G.; Shi, R.C. Eco-friendly extraction and physicochemical properties of pectin from jackfruit peel waste with subcritical water. J. Sci. Food Agric. 2019, 99, 5283–5292. [Google Scholar] [CrossRef] [PubMed]
  46. Vriesmann, L.C.; Teofilo, R.F.; Perkowicz, C.L. Extraction and characterization of pectin from cacao pod husks (Theobroma cacao L.) with citric acid. LWT Food Sci. Technol. 2012, 49, 108–116. [Google Scholar] [CrossRef]
  47. Gullon, B.; Gomez, B.; Martínez-Sabajanes, M.; Yanez, R.; Parajó, J.C.; Alonso, J.L. Pectic oligosaccharides: Manufacture and functional properties. Trends Food Sci. Technol. 2013, 30, 153–161. [Google Scholar] [CrossRef]
Figure 1. Three-dimensional response surface representations illustrating the impact of temperature (°C), extraction time (min), pH and the solvent-to-solid ratio (ml/g) on the extraction yield (%). The interactions are depicted in subfigures: (A) solvent-to-solid ratio vs. temperature; (B) solvent-to-solid ratio vs. extraction time; (C) pH vs. extraction time; (D) pH vs. temperature; (E) extraction time vs. temperature; and (F) solvent-to-solid ratio vs. pH.
Figure 1. Three-dimensional response surface representations illustrating the impact of temperature (°C), extraction time (min), pH and the solvent-to-solid ratio (ml/g) on the extraction yield (%). The interactions are depicted in subfigures: (A) solvent-to-solid ratio vs. temperature; (B) solvent-to-solid ratio vs. extraction time; (C) pH vs. extraction time; (D) pH vs. temperature; (E) extraction time vs. temperature; and (F) solvent-to-solid ratio vs. pH.
Foods 13 02554 g001
Figure 2. Three-dimensional response surface representations illustrating the impact of temperature (°C), extraction time (min), pH, and solvent-to-solid ratio (mL/g) on the extraction DE (%). The interactions are depicted in subfigures: (A) extraction time vs. temperature; (B) pH vs. temperature; (C) solvent-to-solid ratio vs. temperature; (D) pH vs. extraction time; (E) solvent-to-solid ratio vs. extraction time; and (F) solvent-to-solid ratio vs. pH.
Figure 2. Three-dimensional response surface representations illustrating the impact of temperature (°C), extraction time (min), pH, and solvent-to-solid ratio (mL/g) on the extraction DE (%). The interactions are depicted in subfigures: (A) extraction time vs. temperature; (B) pH vs. temperature; (C) solvent-to-solid ratio vs. temperature; (D) pH vs. extraction time; (E) solvent-to-solid ratio vs. extraction time; and (F) solvent-to-solid ratio vs. pH.
Foods 13 02554 g002
Figure 3. The prebiotic activity scores of bacterial cultures grown in MRS supplemented with 2% (w/v) pectin (LDP, MDP, and HDP) or pectin enzyme hydrolysates (PEHs: LPEH, MPEH, and HPEH), as well as inulin, are presented. Data are shown from three independent repeats.
Figure 3. The prebiotic activity scores of bacterial cultures grown in MRS supplemented with 2% (w/v) pectin (LDP, MDP, and HDP) or pectin enzyme hydrolysates (PEHs: LPEH, MPEH, and HPEH), as well as inulin, are presented. Data are shown from three independent repeats.
Foods 13 02554 g003
Table 1. The statistical analysis of variance using ANOVA evaluates the variability and significance within the chosen model for pectin yield and degree of esterification (DE).
Table 1. The statistical analysis of variance using ANOVA evaluates the variability and significance within the chosen model for pectin yield and degree of esterification (DE).
SourceSum of SquaresDFMean SquareF-Valuep-ValueSourceSum of SquaresDFMean SquareF-Valuep-Value
(A) Yield (B) DE
Model1015.571472.5422.26<0.0001 *Model7673.7914548.1363.15<0.0001 *
X113.98113.984.290.0573X125.42125.422.930.1091
X23.3413.341.020.3286X20.004610.00460.00050.9819
X3652.101652.10200.09<0.0001 *X35635.7715635.77649.33<0.0001 *
X4129.891129.8939.86<0.0001 *X435.59135.594.100.0624
X123.3113.311.010.3309X1221.44121.442.470.1384
X2210.64110.643.270.0923X2234.52134.523.980.0660
X32136.641136.6441.93<0.0001 *X321882.1811882.18216.86<0.0001 *
X4242.34142.3412.990.0029 *X4278.04178.048.990.0096 *
X1 X20.1910.190.0580.8131X1 X227.79127.793.200.0952
X1 X32.8112.810.860.3692X1 X30.2010.200.0230.8815
X1 X40.6210.620.190.6703X1 X43.1813.180.370.5545
X2 X33.4813.481.070.3191X2 X31.7011.700.200.6648
X2 X40.02410.0240.00740.9328X2 X42.7012.700.310.5861
X3 X428.62128.628.780.0103 *X3 X440.42140.424.660.0488 *
Residual45.63143.26 Residual121.51148.68
Lack of Fit42.43104.245.320.0608Lack of Fit99.46109.951.800.2992
Pure Error3.1940.80 Pure Error22.0545.51
Total1061.1928 Total7795.3028
Std. Dev.1.81R-Squared0.9570 Std. Dev.2.95R-Squared0.9844
Mean12.89Adj R-Squared0.9140 Mean57.40Adj R-Squared0.9688
C.V.14.00Pred R-Squared0.7650 C.V.5.13Pred R-Squared0.9221
PRESS249.40Adeq Precision16.424 PRESS607.35Adeq Precision24.341
* Significance level at 95%.
Table 2. Recommended solutions for optimal pectin extraction and their characteristics.
Table 2. Recommended solutions for optimal pectin extraction and their characteristics.
Optimal ConditionResponseCharacteristics
Predicted ValueMeasure Value
Temp (°C)Time (min)pHRatio (mL/g)Yield (%)DE (%)Yield (%)DE (%)Eq.wt (g)MeO (%)AUA (%)
LDP8516615022.8345.2021.42 ± 1.4546.17 ± 0.71210.55 ± 3.1312.63 ± 0.5513.05 ± 1.41
MDP851202.25014.2260.0012.02 ± 2.7456.65 ± 1.03303.10 ± 6.4913.37 ± 0.2755.30 ± 12.90
HDP85902.7507.2880.006.53 ± 1.0688.38 ± 0.712678.57 ± 252.548.84 ± 0.2278.43 ± 5.48
LDP = pectin with low DE; MDP = pectin with medium DE; HDP = pectin with high DE; DE = degree of esterification; Eq.wt = equivalent weight; MeO = methoxyl content; AUA = anhydrouronic acid.
Table 3. The log CFU/mL levels of Lactobacillus casei TISTR 390, Lactobacillus plantarum TISTR 877 and Enterococcus faecium were assessed in glucose-free MRS media with 2% supplementation of various types of pectin or PEHs during incubation at 37 °C for 0, 24, and 36 h. Statistically significant differences (p < 0.05) are indicated by distinct letters on bars within the same incubation period.
Table 3. The log CFU/mL levels of Lactobacillus casei TISTR 390, Lactobacillus plantarum TISTR 877 and Enterococcus faecium were assessed in glucose-free MRS media with 2% supplementation of various types of pectin or PEHs during incubation at 37 °C for 0, 24, and 36 h. Statistically significant differences (p < 0.05) are indicated by distinct letters on bars within the same incubation period.
SupplementsL. plantarum TISTR 877L. casei TISTR 390E. faecium TISTR 1027
Time (h)Time (h)Time (h)
024360243602436
LDP4.28 ± 0.05 a6.84 ± 0.03 e8.53 ± 0.07 d4.36 ± 0.12 bc6.58 ± 0.33 d7.16 ± 0.08 c4.31 ± 0.03 bc6.07 ± 0.06 f7.43 ± 0.09 d
MDP4.25 ± 0.02 a7.18 ± 0.03 d8.67 ± 0.01 c4.38 ± 0.03 abc7.36 ± 0.09 bc7.20 ± 0.17 cd4.22 ± 0.03 d6.46 ± 0.11 de7.79 ± 0.07 c
HDP4.25 ± 0.02 a7.15 ± 0.09 d8.66 ± 0.07 c4.49 ± 0.02 ab7.44 ± 0.09 b7.39 ± 0.12 c4.32 ± 0.07 bc7.47 ± 0.12 c8.39 ± 0.06 b
LPEH4.26 ± 0.05 a7.90 ± 0.03 c8.85 ± 0.03 b4.39 ± 0.02 abc6.99 ± 0.27 c8.34 ± 0.04 b4.31 ± 0.05 bc6.33 ± 0.14 e7.73 ± 0.03 cd
MPEH4.27 ± 0.01 a8.42 ± 0.02 a9.21 ± 0.07 a4.48 ± 0.00 ab7.61 ± 0.01 b8.46 ± 0.05 b4.44 ± 0.00 a7.78 ± 0.05 b9.05 ± 0.04 a
HPEH4.25 ± 0.05 a8.30 ± 0.01 b9.11 ± 0.03 a4.51 ± 0.07 a8.29 ± 0.01 a8.72 ± 0.02 a4.39 ± 0.00 ab8.17 ± 0.07 a9.16 ± 0.06 a
Positive control4.30 ± 0.02 a8.28 ± 0.01 b9.18 ± 0.01 a4.33 ± 0.06 c7.99 ± 0.04 a8.73 ± 0.16 a4.43 ± 0.05 a8.28 ± 0.02 a9.09 ± 0.12 a
Negative control4.22 ± 0.03 a7.20 ± 0.03 d8.29 ± 0.07 e4.35 ± 0.02 bc7.50 ± 0.12 b8.37 ± 0.08 b4.26 ± 0.00 cd6.64 ± 0.04 d8.35 ± 0.00 b
LDP = pectin with low DE; MDP = pectin with medium DE; HDP = pectin with high DE; LPEH = pectin enzyme hydrolysate from LDP; MPEH = pectin enzyme hydrolysate from MDP; HPEH = pectin enzyme hydrolysate from HDP.
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

Bilraheem, S.; Srinuanpan, S.; Cheirsilp, B.; Upaichit, A.; Kawai, F.; Thumarat, U. Optimization of Pectin Extraction from Melon Peel as a New Source of Pectin and Pectin Hydrolysate with Prebiotic Potential. Foods 2024, 13, 2554. https://doi.org/10.3390/foods13162554

AMA Style

Bilraheem S, Srinuanpan S, Cheirsilp B, Upaichit A, Kawai F, Thumarat U. Optimization of Pectin Extraction from Melon Peel as a New Source of Pectin and Pectin Hydrolysate with Prebiotic Potential. Foods. 2024; 13(16):2554. https://doi.org/10.3390/foods13162554

Chicago/Turabian Style

Bilraheem, Saroya, Sirasit Srinuanpan, Benjamas Cheirsilp, Apichat Upaichit, Fusako Kawai, and Uschara Thumarat. 2024. "Optimization of Pectin Extraction from Melon Peel as a New Source of Pectin and Pectin Hydrolysate with Prebiotic Potential" Foods 13, no. 16: 2554. https://doi.org/10.3390/foods13162554

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