Next Article in Journal
Current Research on Antioxidant, Anti-Inflammatory and Anti-Obesity Potential of Food Extracts
Previous Article in Journal
Electrospray-Assisted Fabrication of Dextran–Whey Protein Isolation Microcapsules for the Encapsulation of Selenium-Enriched Peptide
Previous Article in Special Issue
Antioxidant, Organoleptic and Physicochemical Changes in Different Marinated Oven-Grilled Chicken Breast Meat
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Investigating the Potential of Full-Fat Soy as an Alternative Ingredient in the Manufacture of Low- and High-Moisture Meat Analogs

Department of Food Science and Technology, Food and Feed Extrusion Research Center, Kongju National University, Yesan 32439, Republic of Korea
*
Author to whom correspondence should be addressed.
These authors contributed equally to this work.
Foods 2023, 12(5), 1011; https://doi.org/10.3390/foods12051011
Submission received: 10 December 2022 / Revised: 13 February 2023 / Accepted: 16 February 2023 / Published: 27 February 2023

Abstract

:
The increase in meat consumption could adversely affect the environment. Thus, there is growing interest in meat analogs. Soy protein isolate is the most common primary material to produce low- and high-moisture meat analogs (LMMA and HMMA), and full-fat soy (FFS) is another promising ingredient for LMMA and HMMA. Therefore, in this study, LMMA and HMMA with FFS were manufactured, and then their physicochemical properties were investigated. The water holding capacity, springiness, and cohesiveness of LMMA decreased with increasing FFS contents, whereas the integrity index, chewiness, cutting strength, degree of texturization, DPPH free radical scavenging activity, and total phenolic content of LMMA increased when FFS contents increased. While the physical properties of HMMA decreased with the increasing FFS content, its DPPH free radical scavenging activity and total phenolic contents increased. In conclusion, when full-fat soy content increased from 0% to 30%, there was a positive influence on the fibrous structure of LMMA. On the other hand, the HMMA process requires additional research to improve the fibrous structure with FFS.

Graphical Abstract

1. Introduction

The Food and Agriculture Organization of the United Nations (FAO) estimates that by 2030, per capita annual meat consumption will be 100 kg in industrialized countries [1]. Mass meat production induces ethical issues for animals, low land utilization, water usage, adverse environmental effects, etc [2]. Specifically, increasing meat products can boost carbon dioxide levels, increasing greenhouse gas formation [3,4]. Thus, consumers are starting to embrace plant-based meat analogs as an alternative source of real meat products because they are more environmentally-friendly and sustainable as a nutrient of protein as well as an aspect of animal welfare and animal rights [5,6,7].
At present, there are many technologies for producing meat analogs, such as extrusion cooking [8,9,10], freeze structuring [11], electrospinning [12], mechanical elongation [13], in vitro cultured meat [14], and the shear-cell technique [15]. Among the technologies, extrusion cooking is one of the most representative methods for the manufacture of meat analogs by using high temperatures, shear, and pressure [16,17,18]. Extrusion cooking is widely used for manufacturing many food products, such as puffed snacks, pasta products, noodles, breakfast cereals, meat analogs, gelatinized starch, dough, baby food, and others, because of its benefits of flexibility, low cost, high output, and quality control [8,10].
Extrusion cooking is classified into low- and high-moisture processing types, according to the degree of addition of moisture content into the extruder and the presence of a cooling die at the end of barrels. The low-moisture extrusion process for manufacturing meat analogs produces expanded meat analogs, and the expansion phenomenon causes sponge-like structures for the meat analogs [19]. Low-moisture meat analog (LMMA) requires hydration prior to making patties, chunks, and nuggets [20]. On the other hand, high-moisture extrusion cooking can produce non-expanded meat analogs that have denser and highly fibrous structures by using a cooling die [19]. More complex formulations are possible with high-moisture extrusion cooking, and it is not necessary to use highly soluble ingredients, making it a more economical technology [21,22].
Meat analogs, produced by extrusion cooking, are mainly composed of 50–95% (dry basis) plant-based proteins [23]. The most widely used plant-based protein in producing meat analogs using extrusion processing is soy protein due to its gelation, functional, and nutritional properties [22]. There are many types of soy protein in the meat analog market based on protein contents and extract processes, such as soy protein isolate (SPI), soy protein concentrate (SPC), full-fat soy (FFS), defatted soy flour (DSF), etc. [5]. Among them, SPI is the one mainly used in many research projects and industries for manufacturing plant-based meat analogs. In the extraction step for SPI, many steps are required to obtain concentrated and purified proteins, which cause environmental pollution and are expensive. In contrast, full-fat soy (FFS) is not only rich in nutrients such as complex carbohydrates, soluble fibers, and isoflavones but it is also easily accessible and cost-effective due to its minimal processing requirements [24,25].
Most research on extruded meat analogs (LMMA and HMMA) has centered around the use of SPI-based formulas due to their functional properties and smooth extrusion process for manufacturing [26,27]. Conversely, the impact of the manufacturing process on the texturization of FFS has not been explored, despite its potential as a substitute for SPI, due to the difficulties associated with its extrusion process. In this study, we aimed to fill this gap by investigating the texturization of FFS using two types of extrusion methods and examining the physical and antioxidant properties of LMMA and HMMA.

2. Materials and Methods

2.1. Materials

Soy protein isolate (SPI), full-fat soy (FFS), wheat gluten (WG), and corn starch (CS) were purchased from Plant Albumen Co., Ltd. (Pingdingshan, China), Korea Seed & Variety Service (Jecheon, Republic of Korea), Roquette Frères (Lestrem, France), and Samyang Co. (Ulsan, Republic of Korea), respectively. The crude protein and fat contents of SPI (84.87 ± 0.1% and 1.84 ± 0.2%), FFS (38.60 ± 0.3% and 20.07 ± 0.1%), and WG (77.81 ± 0.3% and 0.24 ± 0.2%) were measured using the Dumas method [28] and Danlami et al. [29], respectively. The formulation of raw materials for producing LMMA and HMMA is shown in Table 1.

2.2. Manufacturing of Meat Analogs by Low- and High-Moisture Extrusion Cooking

Low- and high-moisture extrusion cookings were performed using a co-rotating intermeshing twin screw extruder (THK31T-No.5, Incheon Machinery Co., Incheon, Republic of Korea) with a 3 cm screw diameter and a 69 cm length (L/D: 23:1). The extrusion conditions—40% feed moisture, 160 °C barrel temperature, 250 rpm screw speed, and a slit die with dimensions of 1 cm (W) × 0.45 cm (H) × 8 cm (L)—were set for the low-moisture extrusion process to manufacture LMMA (Figure 1A). High-moisture extrusion cooking for HMMA was performed with 60% feed moisture, 160 °C barrel temperature, 150 rpm screw speed, and a long cooling die (dimensions of 7 cm (W) × 1 cm (H) × 50 cm (L)) cooled by 20 °C water with a water circulator (Duksan Cotran Co., Ltd., Daegu, Republic of Korea) (Figure 1B). At least 30 extrudates were collected for each sample after the condition was stable. The extrusion process for LMMA was performed until 30% of the FFS inclusion level because of the collapse of the texturized structure at the FFS level above 40%.
After the extrusion cooking process, the extruded samples were cut into 1 cm × 1 cm pieces for subsequent texture profile analysis and to determine the cutting strength of both low-moisture meat analogs (LMMA) and high-moisture meat analogs (HMMA). The cut LMMAs were dried at 50 °C for 12 h and stored at room temperature, and the cut HMMAs were stored in a −18 °C freezer and thawed in a 4 °C refrigerator for 24 h (FR-S690FXB, Klasse Auto Co., Ltd., Seoul, Republic of Korea) before the following analysis. Chemical properties were measured by using the dried LMMA and freeze-dried HMMA samples as ground samples (50–70 mesh particles).

2.3. Texture Profile Analysis and Cutting Strength

Texture profile analysis (TPA) was performed with a texture analyzer (Compac-100, Sun Science Co., Ltd., Tokyo, Japan) using a 2.5 cm diameter cylinder probe with a 10 kg maximum peak stress. The cutting strength (CS) was determined using a cutting probe (0.75 × 3.83 cm) with a 2 kg maximum peak stress. The LMMA sample was hydrated in a water bath at 70 °C for 1 h and then equilibrated at room temperature, and the HMMA sample was equilibrated at room temperature before the determination of the TPA and CS. The meat analogs for CS were cut along the vertical direction (transversal strength, Fv) and parallel direction (longitudinal strength, Fp) of the samples’ flow direction. The calculation of springiness, cohesiveness, chewiness, and CS was expressed according to the equation below [30]. The results were averaged from 6 measurements.
Springiness (%) = D2/D1 × 100
D1: distance reached for the first maximum stress
D2: distance reached for the second maximum stress
Cohesiveness (%) = A2/A1 × 100
A1: area underneath the first compression curve
A2: area underneath the second compression curve
Chewiness (g) = cohesiveness × springiness × highest peak force (g)
Cutting strength (g/cm2) = highest peak force (g)/cross-sectional area (cm2)

2.4. Water Holding Capacity

The water holding capacity (WHC) of LMMA and HMMA was determined by the modified methods of Gu and Ryu [31] and Diaz et al. [32], respectively. Approximately 5 g (on a dry basis) of samples were weighed. The LMMA was hydrated at 70 °C in a water bath for 1 h as per Gu and Ryu [31], and the HMMA was hydrated at 50 °C for 16 h as per Diaz et al. [32]. After that, the samples were drained for 15 min. WHC was expressed as grams of water retained per gram of dried sample using Equation (5).
Water holding capacity (g/g) = (weight of wet sample − weight of dry sample)/(weight of dry sample) × 100

2.5. Degree of Texturization

Degree of texturization was calculated by Equation (6) [33].
Degree of texturization = Fv/Fp
where Fv is the transversal strength, and Fp is the longitudinal strength.

2.6. Integrity Index

The integrity index was determined by a modified method of Gu and Ryu [31]. A 5 g (dry basis) sample that was sunk in 100 mL of distilled water was autoclaved at 121 °C for 15 min, and then the sample was cooled and drained for 30 s. 100 mL of distilled water was added into the sample and homogenized at 17,450 rpm for 1 min, and then they were dried on a 20-mesh sieve at 105 °C for 12 h. The integrity index was calculated according to Equation (7).
Integrity index (%) = (weight of dry residue on sieve)/(weight of sample) × 100

2.7. DPPH Free Radical Scavenging Activity

DPPH free radical scavenging activity was determined as per Brand-Williams et al. [34] with some modifications. A 2 g sample was extracted at room temperature by ethanol (80%) for 2 h, and then the extracted sample was centrifuged at 3000 rpm for 0.5 h. 3.9 mL of DPPH solution (0.0024 g DPPH reagent/100 mL of methanol) was mixed with the supernatant (0.1 mL). The absorbance (at 515 nm) of the solution was read after the mixture was incubated at room temperature in a dark room for 0.5 h. The DPPH free radical scavenging was calculated by Equation (8).
DPPH free radical scavenging activity (%) = (A0 − A)/A0 × 100
where A0 is the absorbance of the blank, and A is the absorbance of the sample.

2.8. Total Phenolic Contents

Total phenolic contents (TPC) were determined as per Slinkard and Singleton’s method [35]. A 2 g sample was extracted with 10 mL of ethanol (80%) for 2 h, and then the extract was centrifuged at 3000 rpm for 0.5 h. A 1.5 mL of 10% (v/v) Folin-Ciocalteu reagent was mixed with 0.3 mL supernatant for 5 min with a vortexer, and then the mixture was mixed with another solution (1.5 mL of Na2CO3 (60 g/L)) and incubated at room temperature for 2 h. After incubation, a wavelength of 765 nm was utilized to measure the absorbance of the solution. TPC was calculated as mg/g of gallic acid equivalents in milligrams per gram of dried sample (mg GAE/g).

2.9. Macrostructure

The macrostructure of cross-sectional LMMA and HMMA was observed after the samples were cut into 1 cm (W) × 1 cm (D). For the fibrous structure of LMMA and HMMA, the extruded samples (LMMA and HMMA) were cut into 3 cm (W) × 5 cm (D) and 5 cm (W) × 7 cm (D), respectively. The cut samples were opened in the longitudinal direction and then photographed for observation of the fibrous structure.

2.10. Statistical Analysis

All experiments were performed in triplicate unless otherwise stated and analyzed using SPSS (version 27.0, IBM-SPSS, Thornwood, NY, USA). Analysis of variance (ANOVA) and comparison of means were performed using Duncan’s multiple range tests at p  <  0.05. Correlation coefficients among the data were determined using Pearson’s correlation coefficient (r).

3. Results and Discussion

3.1. Texture Profile Analysis, Cutting Strength, and Water Holding Capacity

The textural properties (springiness, cohesiveness, and chewiness) and cutting strength of the vertical (V-CS) and parallel (P-CS) directions of LMMA and HMMA are summarized in Table 2. Springiness means how quickly an extruded meat analog is recovered after deformation by a probe; cohesiveness indicates the internal strength of bonds; and chewiness shows the energy requirement for masticating the food [36,37]. Significant differences in the springiness and cohesiveness of both LMMA and HMMA were observed, showing that higher FFS content in LMMA and HMMA caused a decrease in springiness and cohesiveness. Springiness is connected to protein content, and the addition of FFS decreases protein content comparatively, resulting in a decrease in protein cross-linking strength [36]. Ma and Ryu [38] also reported that the springiness and cohesiveness of meat analogs were mainly affected by cross-linking formation in the internal structure. The springiness of both LMMA and HMMA showed a positive correlation with cohesiveness (r = 0.987 and 0.990, respectively, p < 0.01) (Table 3 and Table 4), and the FFS caused a more negative effect on the springiness of LMMA than that of HMMA, resulting in 97.43 ± 3.4 (FFS 0%) to 79.32 ± 5.1% (FFS 30%) for LMMA and 93.03 ± 0.8 (FFS 0%) to 88.58 ± 1.2% (FFS 30%) for HMMA. The result could be due to the fact that high-moisture extrusion cooking is a more effective method of making fibrous structures compared to low-moisture extrusion cooking, maintaining molecular interactions in spite of higher FFS content [39]. The intermolecular protein cross-linking occurs in the cooling die, resulting in high-fibrous structures [40]. Additionally, increasing moisture content can increase the interactions between hydrogen bonds-disulfide bonds, and hydrophobic interactions-disulfide bonds [41].
The chewiness and cutting strength (in both vertical and parallel directions) of LMMA and HMMA had contrary results, showing that, with an increase in FFS content, those of LMMA were increased but those of HMMA were decreased. The increase in LMMA’s texture properties could be due to the interaction of lipids and proteins during the extrusion cooking process [42]. The reason for this is due to the reduction of the expansion ratio of LMMA with increasing FFS content due to the lubricant behavior of the lipids in FFS in the extruder barrel. Then, the behavior resulted in a decrease in the buildup of pressure for the vaporization of water, increasing both non-covalent and covalent interactions to build protein networks with suitable fibrous structures [43,44]. A water-phase change that causes the expansion phenomenon by the vaporization of water in the mixture of raw materials can increase the distance between the substances, resulting in weak structures in meat analogs [45,46]. On the other hand, the HMMA process, in general, did not involve the expansion phenomenon, and thus the vaporization of water by pressure drop did not occur during the process that extends between the protein, starch, and lipid molecules. Lipids are expected to interact with proteins with a saturation limit determined by the number of hydrophobic sites [41], but the decrease in the texture properties of HMMA might be due to the excessive lipid content that disturbs the molecular bonds during the extrusion cooking. According to Van Hoan et al. [47], a high moisture content can reduce the die pressure, thus reducing the amount of lipid lost during the extrusion. In other words, high-moisture extrusion cooking, which has a higher moisture content than low-moisture extrusion cooking with a low die pressure, could cause excessive lipid content that hinders molecular bonding. Alzagtat and Alli [48] reported that lipids might be coated on the surface of the protein aggregates, which prevented the protein molecules from cross-linking.
The water holding capacity (WHC) of LMMA and HMMA is shown in Figure 2. As shown, the lowest WHCs of 2.55 ± 0.1 g/g at FFS 30% for LMMA and 2.48 ± 0.0 g/g at FFS 50% for HMMA were observed, while the highest WHCs were 5.09 ± 0.21 g/g for LMMA and 3.16 ± 0.19 g/g for HMMA at FFS 0%. This is because the WHC is directly affected by the porosity of the meat analogs [49]. The empty spaces are filled by water that builds intra- and inter-hydrogen bonds caused by hydroxyl groups [50]. On the other hand, the reduction of porous structure could be due to the increase in lipid contents from FFS. Ottoboni et al. [51] reported that high lipid content might lead to a decrease in back-pressure during extrusion cooking, resulting in poor expansion.
WHC had positive correlations with springiness (r = 0.934, p < 0.01) and cohesiveness (r = 0.950, p < 0.01) of LMMA and negative correlations with chewiness (r = −0.910, p < 0.01), V-CS (r = −0.972, p < 0.01), and P-CS (r = −0.911, p < 0.01) (Table 3). It was consistent with the results of Gu and Ryu [52] that the WHC of LMMA is positively correlated with springiness and cohesiveness. In contrast, those of HMMA had positive correlations, springiness (r = 0.778, p < 0.01), cohesiveness (r = 0.764, p < 0.01), chewiness (r = 0.821, p < 0.01), V-CS (r = 0.807, p < 0.01), and P-CS (r = 0.767, p < 0.01) (Table 4).

3.2. Degree of Texturization and Integrity Index

The degree of texturization (DT) can be used as an indicator for fibrous structure formation [53], and the integrity index indicates the residue of meat analogs after hydrating, autoclaving, homogenizing, and drying [31,33]. The higher integrity index means the better the texturization, the higher the integrity index, since the meat analogs’ texture remained strong and unweakened after the harsh process steps [19,26,31,54]. Therefore, the DT and integrity index of LMMA and HMMA generally showed a positive correlation (r = 0.955 and r = 0.768, respectively) (p < 0.01) (Table 3 and Table 4) [26,55].
With increasing FFS content from 0% to 30%, the DT and integrity index of LMMA increased, but those of HMMA decreased up to 50% of FFS content (Figure 3). The highest DT and integrity index were observed at LMMA with 30% FFS content (3.04 ± 0.1 and 70.28 ± 1.6%, respectively), and the lowest DT and integrity index were determined at HMMA with 50% FFS content (1.09 ± 0.0 and 74.46 ± 1.4%, respectively). The increase in DT and integrity index of LMMA could be a result of the interactions among phenolic compounds in FFS, protein, starch, and lipid molecules. Polyphenols and protein molecules are known to interact with each other reversibly (hydrogen bonding, hydrophobic bonding, and van der Waals forces) and irreversibly (covalent bonds) [27,53,56]. Alzagtat and Alli [47] also reported that lipids play a role as a plasticizer by forming complexes with starch, protein, and lipid during the extrusion cooking process, which can contribute to the stabilized fibrous structure. Conversely, the reason for the decrease in DT and integrity index of HMMA could be because an excessive amount of lipid hindered protein-protein, protein-starch, protein-lipid, and protein-polyphenol interactions during extrusion cooking.

3.3. DPPH Free Radical Scavenging Activity and Total Phenolic Contents

DPPH free radical scavenging activity and total phenolic contents (TPC) were conducted to find out the antioxidant properties of meat analogs with FFS. A significant change in DPPH free radical scavenging activity and TPC of LMMA and HMMA was observed when FFS was added. The lowest values of DPPH free radical scavenging activity and TPC for both LMMA and HMMA were observed at FFS 0%, and DPPH free radical scavenging and TPC were increased by increasing FFS content from 0 to 30% for LMMA and 0 to 50% for HMMA (Figure 4). FFS content had a significant positive effect on the values of DPPH free radical scavenging activity and TPC that are related to antioxidant activity, showing the highest values of DPPH free radical scavenging activity were 18.14 ± 0.4% for LMMA and 28.99 ± 0.8% for HMMA [57]. This could be because FFS originally contained more phenolic compounds such as tannins, polyphenols, flavonoids, and phenolic terpenes compared to SPI [58]. Aludatt et al. [59] also reported that the phenolic content of the full-fat meal was higher than that of defatted ones, showing higher antioxidant activity.

3.4. Macrostructure

Digital photographs of LMMAs and HMMAs with different FFS content are shown in Table 5. The LMMA was not texturized when the FFS content was higher than 40%. In contrast, the HMMA was texturized to 50% FFS content. The fibrous structure is one of the key factors in judging whether meat analogs have a meat-like texture [60]. The fibrous structure of LMMA increased as the FFS content increased, and LMMA with a FFS content of 30% had the most fibrous structure and well-arranged layers than other LMMAs. The fibrous structure of HMMA was decreased by increasing the FFS content, but HMMA with an FFS of 50% did not exhibit definite layers or fibrous structure. The sponge-like structure of LMMA decreased with an increase in the FFS content, but HMMA showed a dense and layered structure. The structure of LMMA is due to the many pores caused by the expansion phenomenon that resulted from the pressure drop caused by the pressure difference between the inside and outside of the extruder. However, the HMMA was formed with a dense and fibrous structure because of the aggregation effect in the cooling die [26].

4. Conclusions

Overall, FFS was determined to be a promising ingredient for manufacturing meat analogs as an alternative source to SPI in this study. FFS content had a significant effect on the physical and antioxidant properties of LMMA and HMMA (p < 0.05). In addition, LMMA with 30% FFS content showed the most fibrous structures and the highest texture properties (chewiness, cutting strength, integrity index, and degree of texturization). However, texturization was not possible with over 40% FFS content in the LMMA process, but the HMMA process could manufacture texturized proteins with up to 50% FFS content, which completely replaced SPI in this study. The antioxidant properties of LMMA and HMMA increased as the FFS content increased. Further research is needed to increase FFS content for the LMMA process and enhance the texture properties of the HMMA by optimizing the independent process variables of the extrusion process. This study will contribute to enhancing the quality of meat analogs in terms of nutrition and texture aspects and removing the complex extraction steps for producing soy protein isolate, resulting in a reduction in cost and environmental pollution for the manufacture of meat analogs.

Author Contributions

Conceptualization, G.-H.R. and B.-J.G.; methodology, G.-H.R.; formal analysis, Y.-H.J.; investigation, Y.-H.J.; resources, G.-H.R.; writing—original draft preparation, Y.-H.J. and B.-J.G.; writing—review and editing, B.-J.G.; visualization, Y.-H.J.; supervision, B.-J.G.; project administration, G.-H.R.; funding acquisition, G.-H.R. All authors have read and agreed to the published version of the manuscript.

Funding

This work was carried out with the support of the “Cooperative Research Program for Agriculture Science and Technology Development (Project No. PJ016188012022),” Rural Development Administration, Republic of Korea. This work was supported by a research grant from Kongju National University in 2022.

Data Availability Statement

Data is contained within the article.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Bruinsma, J. World Agriculture: Towards 2015/2030: An FAO Study; Routledge: London, UK, 2017; p. 89. [Google Scholar]
  2. Errickson, F.; Kuruc, K.; McFadden, J. Animal-based foods have high social and climate costs. Nat. Food 2021, 2, 274–281. [Google Scholar] [CrossRef]
  3. Vatansever, S.; Tulbek, M.C.; Riaz, M.N. Low-and high-moisture extrusion of pulse proteins as plant-based meat ingredients: A review. Cereal. Foods World 2020, 65, 12–14. [Google Scholar]
  4. Davis, H.; Magistrali, A.; Butler, G.; Stergiadis, S. Nutritional benefits from fatty acids in organic and grass-fed beef. Foods 2022, 11, 646. [Google Scholar] [CrossRef] [PubMed]
  5. Kyriakopoulou, K.; Dekkers, B.; Van Der Goot, A.J. Plant-Based Meat Analogues. In Sustainable Meat Production and Processing; Elsevier: Amsterdam, The Netherlands, 2019; pp. 103–126. [Google Scholar]
  6. Zhou, H.; Hu, Y.; Tan, Y.; Zhang, Z.; McClements, D.J. Digestibility and gastrointestinal fate of meat versus plant-based meat analogs: An in vitro comparison. Food Chem. 2021, 364, 130439. [Google Scholar] [CrossRef]
  7. Yang, C.; Chen, X.; Sun, J.; Gu, C. The Impact of Alternative Foods on Consumers’ Continuance Intention from an Innovation Perspective. Foods 2022, 11, 1167. [Google Scholar] [CrossRef]
  8. Pietsch, V.L.; Emin, M.A.; Schuchmann, H.P. Process conditions influencing wheat gluten polymerization during high moisture extrusion of meat analog products. J. Food Eng. 2017, 198, 28–35. [Google Scholar] [CrossRef]
  9. Smetana, S.; Larki, N.A.; Pernutz, C.; Franke, K.; Bindrich, U.; Toepfl, S.; Heinz, V. Structure design of insect-based meat analogs with high-moisture extrusion. J. Food Eng. 2018, 229, 83–85. [Google Scholar] [CrossRef]
  10. do Carmo, C.S.; Knutsen, S.H.; Malizia, G.; Dessev, T.; Geny, A.; Zobel, H.; Myhere, K.S.; Varela, P.; Sahlstrøm, S. Meat analogues from a faba bean concentrate can be generated by high moisture extrusion. Future Foods 2021, 3, 100014. [Google Scholar] [CrossRef]
  11. Dekkers, B.L.; Boom, R.M.; van der Goot, A.J. Structuring processes for meat analogues. Trends Food Sci. Technol. 2018, 81, 25–36. [Google Scholar] [CrossRef]
  12. Nieuwland, M.; Geerdink, P.; Brier, P.; van den Eijnden, P.; Henket, J.T.M.M.; Langelaan, M.L.P.; Stroeks, N.; van Deventer, H.C.; Martin, A.H. Food-grade electrospinning of proteins. Innov. Food Sci. Emerg. Technol. 2013, 20, 269–275. [Google Scholar] [CrossRef]
  13. Chiang, J.H.; Tay, W.; Ong, D.S.M.; Liebl, D.; Ng, C.P.; Henry, C.J. Physicochemical, textural and structural characteristics of wheat gluten-soy protein composited meat analogues prepared with the mechanical elongation method. Food Struct. 2021, 28, 100183. [Google Scholar] [CrossRef]
  14. Langelaan, M.L.P.; Boonen, K.J.M.; Polak, R.B.; Baaijens, F.P.T.; Post, M.J.; van der Schaft, D.W.J. Meet the new meat: Tissue engineered skeletal muscle. Trends Food Sci. Technol. 2010, 21, 59–66. [Google Scholar] [CrossRef]
  15. Dekkers, B.L.; de Kort, D.W.; Grabowska, K.J.; Tian, B.; Van As, H.; van der Goot, A.J. A combined rheology and time domain NMR approach for determining water distributions in protein blends. Food Hydrocoll. 2016, 60, 525–532. [Google Scholar] [CrossRef]
  16. Krintiras, G.A.; Göbel, J.; Van der Goot, A.J.; Stefanidis, G.D. Production of structured soy-based meat analogues using simple shear and heat in a Couette Cell. J. Food Eng. 2015, 160, 34–41. [Google Scholar] [CrossRef]
  17. Martín-Diana, A.B.; Blanco Espeso, B.; Jimenez Pulido, I.J.; Martínez, P.J.A.; Rico, D. Twin-Screw Extrusion as Hydrothermal Technology for the Development of Gluten-Free Teff Flours: Effect on Antioxidant, Glycaemic Index and Techno-Functional Properties. Foods 2022, 11, 3610. [Google Scholar] [CrossRef]
  18. Zahari, I.; Ferawati, F.; Purhagen, J.K.; Rayner, M.; Ahlström, C.; Helstad, A.; Östbring, K. Development and characterization of extrudates based on rapeseed and pea protein blends using high-moisture extrusion cooking. Foods 2021, 10, 2397. [Google Scholar] [CrossRef]
  19. Lin, S.; Huff, H.E.; Hsieh, F. Extrusion process parameters, sensory characteristics, and structural properties of a high moisture soy protein meat analog. J. Food Sci. 2002, 67, 1066–1072. [Google Scholar] [CrossRef]
  20. Kołodziejczak, K.; Onopiuk, A.; Szpicer, A.; Poltorak, A. Meat Analogues in the Perspective of Recent Scientific Research: A Review. Foods 2021, 11, 105. [Google Scholar] [CrossRef]
  21. Sun, C.; Ge, J.; He, J.; Gan, R.; Fang, Y. Processing, quality, safety, and acceptance of meat analogue products. Engineering 2021, 7, 674–678. [Google Scholar] [CrossRef]
  22. Zhang, T.; Dou, W.; Zhang, X.; Zhao, Y.; Zhang, Y.; Jiang, L.; Sui, X. The development history and recent updates on soy protein-based meat alternatives. Trends Food Sci. Technol. 2021, 109, 702–710. [Google Scholar] [CrossRef]
  23. Osen, R.; Schweiggert-Weisz, U. High-Moisture Extrusion: Meat Analogues. In Reference Module in Food Science; Smithers, G.W., Ed.; Elsevier: Amsterdam, The Netherlands, 2016; ISBN 9780081005965. [Google Scholar]
  24. Cho, S.Y.; Ryu, G.H. Effects of Oyster Mushroom Addition on Physicochemical Properties of Full Fat Soy-based Meat Analog by Extrusion Process. Food Eng. Prog. 2021, 25, 85–94. [Google Scholar] [CrossRef]
  25. Islam, M.; Huang, Y.; Islam, M.S.; Lei, N.; Shan, L.; Fan, B.; Tong, L.; Wang, F. Effect of high-moisture extrusion on soy meat analog: Study on its morphological and physiochemical properties. Ital. J. Food Sci. 2022, 34, 9–20. [Google Scholar] [CrossRef]
  26. Samard, S.; Gu, B.Y.; Ryu, G.H. Effects of extrusion types, screw speed and addition of wheat gluten on physicochemical characteristics and cooking stability of meat analogues. J. Sci. Food Agric. 2019, 99, 4922–4931. [Google Scholar] [CrossRef] [PubMed]
  27. Cho, S.Y.; Ryu, G.H. Effects of mushroom composition on the quality characteristics of extruded meat analog. Korean J. Food Sci. Technol. 2020, 52, 357–362. [Google Scholar]
  28. AOAC. AOAC Official Method 968.06: Protein (crude) in animal feed. Dumas method. In Official Methods of Analysis of AOAC International; AOAC International: Gaithersburg, MD, USA, 2005. [Google Scholar]
  29. Danlami, J.M.; Arsad, A.; Zaini, M.A.A. Characterization and process optimization of castor oil (Ricinus communis L.) extracted by the soxhlet method using polar and non-polar solvents. J. Taiwan Inst. Chem. Eng. 2015, 47, 99–104. [Google Scholar] [CrossRef]
  30. Trinh, T.; Glasgow, S. On the Texture Profile Analysis Test. In Proceedings of the Chemeca 2012: Quality of Life through Chemical Engineering, Wellington, New Zealand, 23–26 September 2012; Available online: https://www.researchgate.net/publication/316093466 (accessed on 9 October 2020).
  31. Gu, B.Y.; Ryu, G.H. Effects of moisture content and screw speed on physical properties of extruded soy protein isolate. J. Korean Soc. Food Sci. Nutr. 2017, 46, 751–758. [Google Scholar]
  32. Diaz, J.R.; Kantanen, K.; Edelmann, J.M.; Suhonen, H.; Sontag-Strohm, T.; Jouppila, K.; Piironen, V. Fibrous meat analogues containing oat fiber concentrate and pea protein isolate: Mechanical and physicochemical characterization. Innov. Food Sci. Emerg. Technol. 2022, 77, 102954. [Google Scholar] [CrossRef]
  33. Chen, F.L.; Wei, Y.M.; Zhang, B.; Ojokoh, A.O. System parameters and product properties response of soybean protein extruded at wide moisture range. J. Food Eng. 2010, 96, 208–213. [Google Scholar] [CrossRef]
  34. Brand-Williams, W.; Cuvelier, M.E.; Berset, C. Use of a free radical method to evaluate antioxidant activity. LWT 1995, 28, 25–30. [Google Scholar] [CrossRef]
  35. Krishnaiah, D.; Sarbatly, R.; Nithyanandam, R. A review of the antioxidant potential of medicinal plant species. Food Bioprod. Process. 2011, 89, 217–233. [Google Scholar] [CrossRef]
  36. Wee, M.S.M.; Goh, A.T.; Stieger, M.; Forde, C.G. Correlation of instrumental texture properties from textural profile analysis (TPA) with eating behaviours and macronutrient composition for a wide range of solid foods. Food Funct. 2018, 9, 5301–5312. [Google Scholar] [CrossRef]
  37. Chandra, M.V.; Shamasundar, B.A. Texture profile analysis and functional properties of gelatin from the skin of three species of fresh water fish. Int. J. Food Prop. 2015, 18, 572–584. [Google Scholar] [CrossRef]
  38. Ma, X.; Ryu, G.H. Effects of green tea contents on the quality and antioxidant properties of textured vegetable protein by extrusion-cooking. Food Sci. Biotechnol. 2019, 28, 67–74. [Google Scholar] [CrossRef]
  39. Yao, G.; Liu, K.S.; Hsieh, F. A new method for characterizing fiber formation in meat analogs during high-moisture extrusion. J. Food Sci. 2004, 69, 303–307. [Google Scholar] [CrossRef]
  40. Chen, F.L.; Wei, Y.M.; Zhang, B. Chemical cross-linking and molecular aggregation of soybean protein during extrusion cooking at low and high moisture content. LWT 2011, 44, 957–962. [Google Scholar] [CrossRef]
  41. Liu, K.S.; Hsieh, H.S. Protein–protein interactions during high-moisture extrusion for fibrous meat analogues and comparison of protein solubility methods using different solvent systems. J. Agric. Food Chem. 2008, 56, 2681–2687. [Google Scholar] [CrossRef]
  42. Mitchell, J.R.; Areas, J.A.G. Structural changes in biopolymers during extrusion. In Food Extrusion Science and Technology; Kokini, J.L., Ho, C.T., Karwe, M.V., Eds.; Marcel Dekker: New York, NY, USA, 1992; pp. 345–371. [Google Scholar]
  43. Choton, S.; Gupta, N.; Bandral, J.; Anjum, N.; Choudary, A. Extrusion technology and its application in food processing: A review. Pharma. Innov. 2020, 9, 162–168. [Google Scholar] [CrossRef]
  44. Kumar, L.; Brennan, M.A.; Mason, S.L.; Zheng, H.; Brennan, C.S. Rheological, pasting and microstructural studies of dairy protein-starch interactions and their application in extrusion-based products: A review. Starch-Stärke 2017, 69, 1600273. [Google Scholar] [CrossRef]
  45. Alam, M.S.; Kaur, J.; Khaira, H.; Gupta, K. Extrusion and extruded products: Changes in quality attributes as affected by extrusion process parameters: A review. Crit. Rev. Food Sci. Nutr. 2016, 56, 445–473. [Google Scholar] [CrossRef]
  46. de Mesa, N.J.E.; Alavi, S.; Singh, N.; Shi, Y.C.; Dogan, H.; Sang, Y. Soy protein-fortified expanded extrudates: Baseline study using normal corn starch. J. Food Eng. 2009, 90, 262–270. [Google Scholar] [CrossRef]
  47. Van Hoan, N.; Mouquet-Rivier, C.; Trèche, S. Effects of starch, lipid and moisture contents on extrusion behavior and extrudate characteristics of rice-based blends prepared with a very-low-cost extruder. J. Food Process Eng. 2010, 33, 519–539. [Google Scholar] [CrossRef]
  48. Alzagtat, A.A.; Alli, I. Protein-lipid interactions in food systems: A review. Int. J. Food Sci. Nutr. 2002, 53, 249–260. [Google Scholar] [CrossRef] [PubMed]
  49. Ning, L.; Villota, R. Influence of 7S and 11S globulins on the extrusion performance of soy protein concentrates. J. Food Proc. Preserv. 1994, 18, 421. [Google Scholar] [CrossRef]
  50. Lee, J.S.; Choi, I.; Han, J. Construction of rice protein-based meat analogues by extruding process: Effect of substitution of soy protein with rice protein on dynamic energy, appearance, physicochemical, and textural properties of meat analogues. Food Res. Int. 2022, 161, 111840. [Google Scholar] [CrossRef] [PubMed]
  51. Ottoboni, M.; Spranghers, T.; Pinotti, L.; Baldi, A.; De Jaeghere, W.; Eeckhout, M. Inclusion of Hermetia Illucens larvae or prepupae in an experimental extruded feed: Process optimisation and impact on in vitro digestibility. Ital. J. Anim. Sci. 2018, 17, 418–427. [Google Scholar] [CrossRef] [Green Version]
  52. Gu, B.Y.; Ryu, G.H. Effects of extrusion variables and die configuration on physicochemical characteristics of texturized soy protein isolate. J. Korean Soc. Food Sci. Nutr. 2019, 48, 1405–1411. [Google Scholar] [CrossRef]
  53. Palanisamy, M.; Töpfl, S.; Aganovic, K.; Berger, R.G. Influence of iota carrageenan addition on the properties of soya protein meat analogues. LWT 2018, 87, 546–552. [Google Scholar] [CrossRef]
  54. Ma, X.; Gu, B.Y.; Ryu, G.H. Optimization of extrusion variables for improving the qualities of textured vegetable protein with green tea using response surface methodology. Food Eng. Prog. 2018, 22, 1–8. [Google Scholar] [CrossRef]
  55. Rajendra, A.; Ying, D.; Warner, R.D.; Ha, M.; Fang, Z. Effect of Extrusion on the Functional, Textural and Colour Characteristics of Texturized Hempseed Protein. Food Bioproc. Tech. 2022, 16, 98–110. [Google Scholar] [CrossRef]
  56. Ozdal, T.; Capanoglu, E.; Altay, F. A review on protein–phenolic interactions and associated changes. Food Res. Int. 2013, 51, 954–970. [Google Scholar] [CrossRef]
  57. Dawidowicz, A.L.; Wianowska, D.; Olszowy, M. On practical problems in estimation of antioxidant activity of compounds by DPPH method (Problems in estimation of antioxidant activity). Food Chem. 2012, 131, 1037–1043. [Google Scholar] [CrossRef]
  58. Rahman, M.A.A.; Moon, S.S. Antioxidant polyphenol glycosides from the Plant Draba nemorosa. Bull. Korean Chem. Soc. 2007, 28, 827–831. [Google Scholar] [CrossRef]
  59. Aludatt, M.H.; Rababah, T.; Ereifej, K.; Alli, I. Distribution, antioxidant, and characterisation of phenolic compounds in soybeans, flaxseed and olives. Food Chem. 2019, 139, 93–99. [Google Scholar] [CrossRef]
  60. Zhang, X.; Zhao, Y.; Zhao, X.; Sun, P.; Zhao, D.; Jiang, L.; Sui, X. The texture of plant protein-based meat analogs by high moisture extrusion: A review. J. Texture Stud. 2022, 1–14. [Google Scholar] [CrossRef]
Figure 1. Low- (A) and high-moisture (B) extrusion processes and screw configurations used in this experiment.
Figure 1. Low- (A) and high-moisture (B) extrusion processes and screw configurations used in this experiment.
Foods 12 01011 g001
Figure 2. Water holding capacity of meat analogs with various full-fat soy (FFS) contents. LMMA: low-moisture meat analog; HMMA: high-moisture meat analog. Different letters (a–d) above the bars indicate the significantly different for each extrusion type (LMMA and HMMA) (p ≤ 0.05) by Duncan’s multiple range tests.
Figure 2. Water holding capacity of meat analogs with various full-fat soy (FFS) contents. LMMA: low-moisture meat analog; HMMA: high-moisture meat analog. Different letters (a–d) above the bars indicate the significantly different for each extrusion type (LMMA and HMMA) (p ≤ 0.05) by Duncan’s multiple range tests.
Foods 12 01011 g002
Figure 3. Degree of texturization (A) and integrity index (B) of extruded low- and high-moisture meat analogs with different full-fat soy (FFS) content. LMMA: low-moisture meat analog; HMMA: high-moisture meat analog. Different letters (a–e) above the bars indicate the significantly different for each extrusion type (LMMA and HMMA) (p ≤ 0.05) by Duncan’s multiple range tests.
Figure 3. Degree of texturization (A) and integrity index (B) of extruded low- and high-moisture meat analogs with different full-fat soy (FFS) content. LMMA: low-moisture meat analog; HMMA: high-moisture meat analog. Different letters (a–e) above the bars indicate the significantly different for each extrusion type (LMMA and HMMA) (p ≤ 0.05) by Duncan’s multiple range tests.
Foods 12 01011 g003
Figure 4. DPPH free radical scavenging activity (A) and total phenolic contents (B) of meat analogs with various by full-fat soy (FFS) contents. LMMA: low-moisture meat analog; HMMA: high-moisture meat analog. Different letters (a–e) above the bars indicate the significantly different for each extrusion type (LMMA and HMMA) (p ≤ 0.05) by Duncan’s multiple range tests.
Figure 4. DPPH free radical scavenging activity (A) and total phenolic contents (B) of meat analogs with various by full-fat soy (FFS) contents. LMMA: low-moisture meat analog; HMMA: high-moisture meat analog. Different letters (a–e) above the bars indicate the significantly different for each extrusion type (LMMA and HMMA) (p ≤ 0.05) by Duncan’s multiple range tests.
Foods 12 01011 g004
Table 1. Formulation of extruded low- and high-moisture meat analog with different full-fat soy content.
Table 1. Formulation of extruded low- and high-moisture meat analog with different full-fat soy content.
Extrusion TypeFull-Fat Soy (%)Soy Protein Isolate (%)Wheat Gluten (%)Corn Starch (%)
LMMA0504010
10404010
20304010
30204010
HMMA0504010
10404010
20304010
30204010
40104010
5004010
Table 2. Texture profile analysis and cutting strength (vertical and parallel) of meat analogs with various full-fat soy contents.
Table 2. Texture profile analysis and cutting strength (vertical and parallel) of meat analogs with various full-fat soy contents.
Extrusion TypeFFS Content (%)Springiness (%)Cohesiveness (%)Chewiness (g)Cutting Strength (g/cm2)
VerticalParallel
LMMA097.43 ± 3.4 a94.15 ± 3.5 a263.42 ± 98.6 c551.3 ± 38.9 d351.2 ± 30.4 c
1084.57 ± 1.3 b79.74 ± 2.3 b399.16 ± 82.2 b999.7 ± 69.3 c403.4 ± 64.7 bc
2083.87 ± 5.3 bc74.95 ± 3.5 b436.72 ± 43.5 b1193.4 ± 68.2 b427.5 ± 45.1 b
3079.32 ± 5.1 c67.58 ± 6.0 c537.67 ± 34.5 a1677.5 ± 66.2 a552.0 ± 31.1 a
HMMA093.03 ± 0.8 a79.07 ± 1.2 a4295.64 ± 130.9 a1236.8 ± 128.8 a711.7 ± 114.8 a
1093.78 ± 0.6 a79.15 ± 0.7 a3967.45 ± 114.2 b1020.7 ± 52.8 b667.9 ± 30.1 ab
2090.01 ± 1.0 b74.45 ± 2.0 b3559.24 ± 175.4 c977.7 ± 57.5 b625.5 ± 37.5 bc
3088.58 ± 1.2 b73.11 ± 1.8 b3170.47 ± 172.5 d947.2 ± 63.8 b626.0 ± 41.6 bc
4086.26 ± 1.5 c69.73 ± 4.1 c2550.08 ± 206.1 e821.1 ± 70.6 c562.3 ± 47.4 c
5075.09 ± 3.6 d54.51 ± 2.8 d1183.40 ± 115.6 f477.9 ± 35.0 d438.6 ± 39.3 d
Different letters (a–f) in the same column for each extrusion type indicate the significantly different (p ≤ 0.05) by Duncan’s multiple range tests. FFS: Full-fat soy content; LMMA: low-moisture meat analog; HMMA: high-moisture meat analog.
Table 3. Pearson correlation matrix of physical properties for low-moisture meat analogs.
Table 3. Pearson correlation matrix of physical properties for low-moisture meat analogs.
WHCSpringinessCohesivenessChewinessV-CSP-CSDTIntegrity Index
WHC1
Springiness0.934 **1
Cohesiveness0.950 **0.987 **1
Chewiness−0.910 **−0.905 **−0.917 **1
V-CS−0.972 **−0.949 **−0.961 **0.928 **1
P-CS−0.911 **−0.875 **−0.891 **0.917 **0.968 **1
DT−0.925 **−0.964 **−0.957 **0.841 **0.910 **0.790 **1
Integrity index−0.962 **−0.949 **−0.952 **0.866 **0.965 **0.881 **0.955 **1
Water holding capacity (WHC); vertical cutting strength (V-CS); parallel cutting strength (P-CS); degree of texturization (DT). Values with ** were significantly different (p < 0.01).
Table 4. Pearson correlation matrix of physical properties for high-moisture meat analogs.
Table 4. Pearson correlation matrix of physical properties for high-moisture meat analogs.
WHCSpringinessCohesivenessChewinessV-CSP-CSDTIntegrity Index
WHC1
Springiness0.778 **1
Cohesiveness0.764 **0.990 **1
Chewiness0.821 **0.976 **0.964 **1
V-CS0.807 **0.916 **0.900 **0.950 **1
P-CS0.767 **0.864 **0.840 **0.889 **0.958 **1
DT0.715 **0.889 **0.882 **0.907 **0.902 **0.752 **1
Integrity index0.777 **0.749 **0.739 **0.837 **0.809 **0.720 **0.768 **1
Water holding capacity (WHC); vertical cutting strength (V-CS); parallel cutting strength (P-CS); degree of texturization (DT). Values with ** were significantly different (p < 0.01).
Table 5. Fibrous and cross-sectional structures of extruded high-moisture meat analogs with different full-fat soy (FFS) content.
Table 5. Fibrous and cross-sectional structures of extruded high-moisture meat analogs with different full-fat soy (FFS) content.
Extrusion TypesStructureFull-Fat Soy Content (%)
01020304050
LMMACross-sectionalFoods 12 01011 i001Foods 12 01011 i002Foods 12 01011 i003Foods 12 01011 i004
FibrousFoods 12 01011 i005Foods 12 01011 i006Foods 12 01011 i007Foods 12 01011 i008
HMMACross-sectionalFoods 12 01011 i009Foods 12 01011 i010Foods 12 01011 i011Foods 12 01011 i012Foods 12 01011 i013Foods 12 01011 i014
FibrousFoods 12 01011 i015Foods 12 01011 i016Foods 12 01011 i017Foods 12 01011 i018Foods 12 01011 i019Foods 12 01011 i020
LMMA: low-moisture meat analog; HMMA: high-moisture meat analog.
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

Jeon, Y.-H.; Gu, B.-J.; Ryu, G.-H. Investigating the Potential of Full-Fat Soy as an Alternative Ingredient in the Manufacture of Low- and High-Moisture Meat Analogs. Foods 2023, 12, 1011. https://doi.org/10.3390/foods12051011

AMA Style

Jeon Y-H, Gu B-J, Ryu G-H. Investigating the Potential of Full-Fat Soy as an Alternative Ingredient in the Manufacture of Low- and High-Moisture Meat Analogs. Foods. 2023; 12(5):1011. https://doi.org/10.3390/foods12051011

Chicago/Turabian Style

Jeon, Yung-Hee, Bon-Jae Gu, and Gi-Hyung Ryu. 2023. "Investigating the Potential of Full-Fat Soy as an Alternative Ingredient in the Manufacture of Low- and High-Moisture Meat Analogs" Foods 12, no. 5: 1011. https://doi.org/10.3390/foods12051011

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