ARTICLE

Effects of Phosphate and Two-Stage Sous-Vide Cooking on Textural Properties of the Beef Semitendinosus

Nurul Nazirah Ruslan1https://orcid.org/0000-0003-2107-4343, John Yew Huat Tang1https://orcid.org/0000-0003-4608-9840, Nurul Huda2https://orcid.org/0000-0001-9867-6401, Mohammad Rashedi Ismail-Fitry3https://orcid.org/0000-0002-2280-5075, Ismail Ishamri1,*https://orcid.org/0000-0003-4820-8292
Author Information & Copyright
1Faculty of Bioresources and Food Industry, Universiti Sultan Zainal Abidin, Besut 22200, Terengganu, Malaysia
2Faculty of Sustainable Agriculture, Universiti Malaysia Sabah, Sandakan 90509, Sabah, Malaysia
3Department of Food Technology, Faculty of Food Science and Technology, Universiti Putra Malaysia, Serdang 43400, Selangor, Malaysia
*Corresponding author: Ismail Ishamri, Faculty of Bioresources and Food Industry, Universiti Sultan Zainal Abidin, Besut 22200, Terengganu, Malaysia, Tel: +60-9-6993129, Fax: +60-9-6993100, E-mail: ishamriismail@unisza.edu.my

© Korean Society for Food Science of Animal Resources. This is an Open-Access article distributed under the terms of the Creative Commons Attribution Non-Commercial License (http://creativecommons.org/licenses/by-nc/3.0) which permits unrestricted non-commercial use, distribution, and reproduction in any medium, provided the original work is properly cited.

Received: Jan 06, 2023 ; Revised: Mar 02, 2023 ; Accepted: Mar 06, 2023

Published Online: May 01, 2023

Abstract

Comparing the effects of sodium tripolyphosphate (STPP) concentrations of 0.2% and 0.4% on beef semitendinosus is the objective of the current investigation. The samples were cooked at varied temperatures (45+60°C and 45+70°C) and times (1.5+1.5 h and 3+3 h) using staged cooking. The colour properties, cooking loss, water retention, shear force, water-holding capacity, sarcoplasmic, and myofibrillar solubility, and total collagen were investigated. The cooking time and temperature affected the water-holding capacity, cooking loss, CIE L*, CIE a*, CIE b*, myofibrillar, and sarcoplasmic solubility, with lower temperature and short time having the lower detrimental effect. However, the significant effect can be intensified after the addition of STPP with higher water-holding capacity and tender meat obtained with 0.4% phosphate concentration at any cooking conditions. The STPP lowered the collagen content and increased the protein solubility of myofibrillar and sarcoplasmic, which this degradation is used as a good indicator of tenderness.

Keywords: sous-vide; phosphate; beef; collagen; protein solubility

Introduction

Among the several heat treatment methods for meat, the sous-vide approach is gaining popularity. This method has been utilised in the food sector since it was discovered to be effective for extending the shelf life of pasteurised meals and vacuum-packed goods in the 1960s. In the mid-1970s, French chef George Pralus developed the technique (Przybylski et al., 2021). Sous-vide cooking involves cooking meat at low temperatures (50°C–80°C) for a long time, however the method depends on the type of meat (del Pulgar et al., 2012). Studies conducted in the past have demonstrated that sous vide cooking lowers shear force, volatile flavour loss, and moisture loss, conserving sensory qualities linked to quality (Trbovich et al., 2018).

Semitendinosus is a large muscle located on the hind leg of mammals and it is categorized as tough meat (Ismail et al., 2019c). The tough cuts of semitendinosus are optimized by mild temperatures or effectively by staged cooking as proposed by Ismail et al. (2022). This optimization minimizes the shear toughness from connective tissue and myofibrillar proteins. Connective tissues are the leading cause of toughness in many cuts of meat which denaturation occurs between 53°C and 63°C, while contraction occurs around 65°C (Ismail et al., 2019b). Meanwhile, myofibrillar proteins are likely contributing to the subsequent rise in meat toughness (Baldwin, 2012) which denaturation of myosin occurs at 60°C while actin at 66°C–73°C (Ismail et al., 2019b).

Phosphates in meat products have a variety of functions and can affect pH, chelation, ionic strength, and antibacterial activity. Also, researchers enhanced the tenderness and juiciness of beef by adding salts and phosphate to the meat (Hoffman et al., 2008). Monophosphates are excellent buffers, yet they have little effect on muscle proteins. Tri- and polyphosphates are helpful in activating meat proteins because they partially chelate protein-bound magnesium and calcium ions. This results in increased myosin and actin solubilization and depolymerization of thick and thin filaments (Glorieux et al., 2017). Roldán et al. (2014) tested how phosphate brining affected the physical, chemical, and sensory qualities of lamb cooked in a vacuum bag. They found that phosphate solution improved sensory textural features and reduced the toughness of lamb loins. Thus, the goal of this investigation was to ascertain how varying phosphate concentrations affected the textural characteristics of beef semitendinosus subjected with two-stage sous vide cooking.

Materials and Methods

Sampling and thermal treatment

At 24 hours postmortem, the semitendinosus muscle (from native cow) was procured from the local market. All of the muscles were chosen at random on the day of collection (30–32 months old). The collection of paired semitendinosus muscles was carried out and repeated every week until week 8. The fresh cuts were immediately placed in an ice box and brought to the muscle laboratory of University Sultan Zainal Abidin. Muscles were refrigerated to 36 h postmortem at 4°C.

The muscle was cut to 1.5 cm thickness (90–105 g) and divided according to the treatment in Table 1. Prior to vacuum packaging, each steak was injected with sodium tripolyphosphate (STPP) at different concentrations (Table 1) and subsequently was tumbled intermittently for 1 h at 8 rpm and at 4°C to ensure the STPP spread evenly within the muscle. The steaks were tumbled, vacuum sealed, then cooked in water baths on an Anova Precision stove (Anova, San Francisco, CA, USA). For a temperature of 45+60°C, steaks were heated to 45°C at first and 60°C at second. The same applies to 45+70°C. Both treatments were cooked for 3 and 6 h (Table 1). Each block of samples was cooked and then cooled at 4°C for 15 h before analysis.

Table 1. Two-stage sous-vide with and without addition of phosphate
Treatment (°C)1) Temperature 1 (°C) Temperature 2 (°C) Time 1 (h) Time 2 (h) STPP concentration (%)
45+60 45 60 1.5 1.5 0.0
0.2
0.4
45 60 3.0 3.0 0.0
0.2
0.4
45+70 45 70 1.5 1.5 0.0
0.2
0.4
45 70 3.0 3.0 0.0
0.2
0.4

1) Beef was first cooked at 45°C for 1.5 h or 3.0 h and then at 60°C or 70°C for 1.5 h or 3.0 h.

STPP, sodium tripolyphosphate.

Download Excel Table
Cooking loss, water-holding capacity (WHC), and water content

After removing the steaks from the vacuum bag and wiping away excess moisture on the meat surface with a clean wipe, the cooking loss was measured. The proportion of cooking loss was determined by comparing the steaks’ pre- and post-cooking weights (Ismail et al., 2019c). Based on AOAC (2002), the water content was carried out by drying a 4 g sample in an oven for 16 hours at 105°C. WHC was conducted based on Joo (2018) using the filter paper press method of a 3 g sample with a 2.5 kg load.

Colour properties

A colorimeter (Chroma metre, CR-300, Konica Minolta, Osaka, Japan) outfitted with a pulse xenon lamp, 8 mm of reading surface area, a standard observer in 2° position, and standard illuminant D65 was used to test colour attributes at different locations on the steak cut surface. The instrument was first calibrated to y=0.3198, X=0.3132, Y=93.5, using a white ceramic plate. The CIE L*, CIE a*, and CIE b* of steak samples were recorded.

Shear force

A double arm texture analyser was used to analyse the shear force (Stable Micro System, London, UK). From each sample, beef cubes of the same size and shape were cut perpendicular to the myofiber, measuring about 1 cm in diameter. Only muscle with no discernible fat and connective tissue were used for this study. A 3-mm thick steel blade (HDP/BSW) with a 73°V cut into its lower edge was used for the test, which was fitted through a 4-mm wide slot in the platform. The sample was tested and placed onto a platform under the blade and cut perpendicular to muscle fibre with constant speed (parameter: test speed, 2.00 mm/sec; post-test speed, 10.00 mm/sec; travel distance, 50 mm). The highest peak force that was measured throughout the test was shear force. The Warner-Bratzler shear force value in N was calculated using the average of five measurements for each sample (Newtons).

Total collagen

Total collagen was measured according to Ismail et al. (2019c). In 30 mL of 3.5 M H2SO4, 4 g of beef were hydrolysed at 105°C for 16 h. The hydrolysate was filtered and diluted with distilled water to a volume of 500 mL. A graduated cylinder with a volume of 100 mL was filled with distilled water after pipetting the diluent (1 mL) into it. The oxidation solution (100 mL of buffer solutions with a pH of 6 and 1.4 g of the chloramines-T reagent) was then mixed to the final diluent, which had a volume of 2 mL. The buffer solution was made by combining 29 mL of 1-propanol with 1.5 g of sodium hydroxide, 9 g of sodium acetate trihydrate, and 3 g of citric acid monohydrate. From there, it was diluted to 100 mL with distilled water. A freshly prepared 1 mL of colour reagent (produced by dissolving 10 g of 4-(dimethylamino) benzaldehyde in 65 mL of 2-propanol and 35 mL of HCl) was then added to the oxidised sample and vortexed. The test tube was sealed, placed in a water bath set at 60°C for 15 min, and then swiftly cooled with running water. A UV-Vis spectrophotometer operating at 560 nm was used to measure the absorbance of solutions after cooling. Hydroxyproline at concentrations of 0, 1.2, 2.4, 3.6, and 4.8 g hydroxyproline/mL was used to generate a standard calibration curve. Using coefficient factor 8, the amount of collagen was calculated from the amount of hydroxyproline.

Actomyosin preparation

The process of extraction of actomyosin was described Mac Donald and Lanier (1994). A homogeniser (Model IKA® T25, IKA-Werke, Staufen, Germany) was used to homogenise 1 g of beef sample for 4 min in 10 mL of cold (4°C) 0.6 M KCl, pH 7.0. The sample beaker was placed in an ice bath, and every 20 s of blending was followed by a 20 s rest period to prevent overheating during the extraction process. At 0°C, the extract was centrifuged for 30 min at 2,000×g using a Sigma 3-18K centrifuge. Three volumes of cold, deionized water were added, followed by a 2-min vortex, to precipitate the actomyosin. The water-soluble sarcoplasmic protein was obtained by 20 min of centrifugation at 0°C and 2,000×g. In an equal volume of cold, 1.2 M KCl, pH 7.0, the pellet was dissolved overnight in an incubator shaker at 120 rpm at 1°C to obtain actomyosin. Any undissolved material was removed from the preparation by centrifugation for 20 min at 0°C at 2,000×g.

Protein solubility

A 1 mL sample of soluble myofibrillar and sarcoplasmic was pipette into a Kjeldahl tube to measure protein solubility. The AOAC official method 981.10 was used for digestion, distillation, and titration (AOAC, 2000). The solubility of proteins was determined using a conversion factor of 6.25.

Statistical analysis

SPSS v23.0 was used to conduct all statistical analyses. The interaction between cooking times (3 and 6 h), cooking temperatures (45+60°C and 45+70°C) and STPP concentrations (0%, 0.2%, and 0.4%) were analysed using the general linear model and Duncan test for multiple mean comparisons with level of significance at 0.05.

Results and Discussion

Cooking loss, water content, and water-holding capacity (WHC)

Table 2 illustrates the influence of different times, temperatures, and STPP concentrations on cooking loss, water content, and WHC. Cooking loss of beef semitendinosus was affected by time (p=0.002), temperature (p<0.001), STPP concentration (p<0.001), and interaction between cooking temperature×time (p=0.022). Lower cooking temperatures and phosphate added resulted in greater water content values. STPP concentration (p=0.007) and temperature (p<0.001) both had an impact on these values. Similarly, for WHC of sous-vide cooked beef was affected by STPP (p<0.001), time (p=0.009), and temperature (p=0.002). Despite being cooked for a longer period of time and at a higher temperature, the STPP inclusion in the current study successfully increased the water retention of sous-vide beef. According to changes in lamb loins cooked sous-vide by Roldán et al. (2014), adding phosphate had an effect on the moisture content and cooking loss but not the WHC.

Table 2. Cooking loss, water content, and water holding capacity (WHC) of cooked sous-vide at different time, temperature, and sodium tripolyphosphate (STPP) concentration
Time (h) Temperature (°C) STPP concentration (%) Cooking loss (%) Water content (%) WHC (%)
3 45+60 0 33.11±0.85cd 65.50±0.46abc 85.90±0.81d
0.2 30.06±0.40de 65.88±0.66abc 92.38±0.91bc
0.4 28.53±0.42e 67.82±0.39a 96.64±0.21a
45+70 0 46.09±0.74a 59.14±0.53e 82.11±0.82ef
0.2 41.95±0.75b 66.48±0.90abc 90.99±0.13c
0.4 40.57±0.20b 63.87±0.26bcd 95.86±0.01a
6 45+60 0 35.11±0.10c 65.54±1.05abc 82.73±0.69e
0.2 34.81±0.86c 66.23±0.30abc 92.70±0.86bc
0.4 33.43±0.10cd 67.53±0.08ab 94.80±0.85ab
45+70 0 48.20±0.29a 59.61±0.84e 79.63±1.02f
0.2 42.34±0.77b 60.31±0.85de 90.27±0.78c
0.4 40.78±0.90b 62.64±0.15cde 94.08±0.59ab
P time 0.002 0.106 0.009
P temperature <0.001 <0.001 0.002
P concentration <0.001 0.007 <0.001
P time×P temp 0.022 0.097 0.928
P time×P temp×P conc. 0.406 0.096 0.250

Mean±SD.

a–f Means value within different letters in same column referring to significant different (p<0.05).

Download Excel Table

The incorporation of phosphate was thought to positively affect meat pH as described by Roldán et al. (2014). This is an unarguable finding as adding alkaline phosphate (such as pyrophosphate or tripolyphosphate) to manufactured meat products will increase the pH. As a result, the meat proteins’ electrostatic repulsion with one another or inside the meat causes a rise in WHC (Glorieux et al., 2017). In addition, the change in the ionic strength will also be related to an increase in WHC. Phosphate addition enhances actomyosin solubility by forming polyelectrolytes in water, causing the protein filaments to swell more (Glorieux et al., 2017; Roldán et al., 2014). However, our results did not support the first evidence because phosphate has no significant effect on pH (p=0.170, data not shown). The present study was related to the second evidence as shown in Table 3. Proteins solubility of myofibrillar were significantly shown an effect (p<0.001) and a strong correlation to WHC (r=0.817) with the STPP concentration. Meanwhile, the trend of cooking loss was consistent with that found in the water content.

Table 3. Shear force, total collagen, myofibrillar, and sarcoplasmic solubility of cooked sous-vide at different time, temperature, and sodium tripolyphosphate (STPP) concentration
Time (h) Temperature (°C) STPP concentration (%) Shear force (kg) Total collagen (%) Myofibrillar solubility (%) Sarcoplasmic solubility (%)
3 45+60 0 11.42±0.94b 4.55±0.47b 4.51±0.23d 2.93±0.33def
0.2 10.34±1.32b 3.82±0.72bc 5.25±0.19c 3.41±0.08bc
0.4 9.92±1.40b 3.54±0.04bc 6.21±0.32ab 4.09±0.08a
45+70 0 18.24±1.71a 7.53±1.50a 4.18±0.08d 2.50±0.01gh
0.2 13.14±1.68ab 3.79±0.02bc 4.51±0.08d 2.83±0.71efg
0.4 9.52±1.19b 2.98±0.16c 5.85±0.19a 3.18±0.14cde
6 45+60 0 15.56±0.29ab 4.44±0.03b 4.17±0.09d 2.75±0.04fg
0.2 13.45±1.64ab 3.98±0.30bc 4.70±0.04d 3.22±0.08cd
0.4 9.96±1.27b 3.38±0.19bc 5.33±0.50bc 3.75±0.18ab
45+70 0 14.15±1.70ab 4.03±0.45bc 3.45±0.32e 1.60±0.28i
0.2 12.33±1.71ab 3.65±0.42bc 4.20±0.16d 2.17±0.19h
0.4 12.05±1.09b 3.62±0.18bc 4.61±0.22d 2.74±0.13fg
P time 0.438 0.039 <0.001 <0.001
P temperature 0.812 0.187 <0.001 <0.001
P concentration 0.013 <0.001 <0.001 <0.001
P time×P temp 0.143 0.053 0.382 <0.007
P time×P temp×P conc. 0.507 0.020 0.189 0.487

Mean±SD.

a–i Means value within different letters in same column referring to significant different (p<0.05).

Download Excel Table

Sous-vide cooked for a shorter period showed a lower cooking loss than prolonged cooking time. Meanwhile, a higher temperature of two-stage sous-vide (45+70°C) contributed to a higher cooking loss as compared to a mild temperature (45+60°C), this was consistent with our previous finding on the Korean beef semitendinosus (Hanwoo steers) and Korean native black goat biceps femoris and gluteus medius (Capra hircus coreanae), regardless of phosphate addition (Ismail et al., 2019b; Ismail et al., 2019c).

Colour analysis

Table 4 lists the mean values of the meat’s CIE L*, CIE a*, and CIE b* for various temperatures, cooking periods, and STPP concentrations. The CIE L*, CIE a*, and CIE b* of sous-vide meat were significantly affected by the time (p<0.001), temperature (p<0.001) and STPP concentration (p=0.007) as well as their interaction (p<0.001). The CIE L* in the present study were in line with Roldán et al. (2014). As shown in Table 4, the sous-vide treatment at 45+60°C resulted in lower CIE L* after brining with STPP. Nevertheless, sous-vide beef at 45+70°C for both cooking durations (3 and 6 h) demonstrated a small increase in CIE L* with STPP concentration. We empirically observed that the effect of pH in the present study is not the primary concern to relate with the lower CIE L*, because as mentioned above phosphate addition has no significant effect on pH. Contrary to the findings of Ayub and Ahmad (2019) and Roldán et al. (2014), they found that lower CIE L* were due to the phosphates that increase the pH thereby lowering the CIE L* of meat. According to Roldán et al. (2014), the addition of phosphate alters the pH and ionic strength and causes the myofibrillar proteins to swell. Lower CIE L* result from the enlarged proteins’ deeper light penetration into the tissue. Nevertheless, this argumentation is much related to the actomyosin complex dissociation in which phosphate promotes the depolymerization of myosin and actin filaments into separate fibres (Glorieux et al., 2017; Tan et al., 2018). This is demonstrated by Table 3’s data, which shows that myofibrillar solubility is higher at mild temperatures (45+60°C) than at high temperatures (45+70°C). Therefore, this finding strongly supports the lower CIE L* in sous-vide cooked beef due to the higher protein solubility. Nevertheless, larger loss caused by increased protein denaturation and shortened sarcomere at a temperature above 60°C was the cause of the higher CIE L* for treatment at 45+70°C (Ismail et al., 2019c). According to Christensen et al. (2011), this resulted in an increase in light scattering and higher CIE L*.

Table 4. Means of colour properties (CIE L*, CIE a*, CIE b*) of cooked sous-vide at different temperature, time, and sodium tripolyphosphate (STPP) concentration
Time (h) Temperature (°C) STPP concentration (%) CIE L* CIE a* CIE b*
3 45+60 0 41.70±0.40d 19.35±0.19a 13.46±0.02g
0.2 38.56±0.01g 19.41±0.33a 13.64±0.05f
0.4 33.92±0.07h 16.51±0.07b 11.84±0.03i
45+70 0 38.67±0.02g 13.71±0.14f 14.12±0.03e
0.2 32.81±0.40i 15.49±0.06cd 14.06±0.11e
0.4 45.15±0.16b 11.90±0.03g 14.71±0.01b
6 45+60 0 50.08±0.33a 14.07±0.65ef 14.32±0.16d
0.2 43.39±0.01c 16.28±1.01bc 14.52±0.10c
0.4 38.62±0.25g 14.51±0.33ef 13.75±0.08f
45+70 0 40.54±0.01e 14.84±0.10de 14.58±0.01b
0.2 39.41±0.30f 15.89±0.01bc 15.24±0.01a
0.4 43.25±0.05c 16.11±0.06bc 13.25±0.08h
P time <0.001 <0.001 <0.001
P temperature <0.001 <0.001 <0.001
P concentration <0.001 0.007 0.007
P time×P temp <0.001 <0.001 <0.001
P time×P temp×P conc. <0.001 <0.001 <0.001

Mean±SD.

a–i Means value within different letters in same column referring to significant different (p<0.05).

Download Excel Table

The thermal treatment of the cooked beef semitendinosus at various temperatures and times caused a change in the CIE a*, as indicated in Table 4. The sous-vide samples that were prepared at lower temperatures and in less time had the highest CIE a*. However, the effect of phosphate on CIE a* varied and even showed significance (p=0.007). According to Lawrie and Ledward (2006), the myoglobin denaturation and endpoint cooking temperature have a significant impact on the CIE a* characteristics of cooked beef. However, as stated by Hunt et al. (1999), myoglobin denaturation started at 55°C and continued until 80°C, occurring at no precise temperature threshold. Also, it depends on the cooking time and cooking temperature (Roldán et al., 2013), as evidenced by the significant effect of time (p<0.001) and temperature (p<0.001) in Table 4. However, there was inconsistent effect of STPP in sous-vide cooked meat on CIE a*. A similar effect can be seen on the CIE b* as the results show the variable. Although phosphate reduced the amount of oxidation in the restructured beef steaks studied by Lamkey et al. (1986) STPP considerably changed the values of raw beef CIE L*, CIE a*, and CIE b* but had no effect on cooked beef (Long et al., 2011). Higher CIE b* in Table 4 can be linked to a temperature increase and prolonged cooking time as a consequence of metmyoglobin formation. It seems that phosphate alone could not effectively modify the redox state of cooked meat, instead oxidation mechanism to metmyoglobin was more dominant. Similar results were also obtained by Roldán et al. (2014) and Önenç et al. (2004) who considered the effect of phosphate added was negligible to CIE b*.

Shear force, total collagen, myofibrillar and sarcoplasmic solubility

Shear strength, total collagen, myofibrillar, and sarcoplasmic solubility of sous-vide cooked beef samples at different durations, temperatures, and phosphate concentrations are shown in Table 3. Only the different STPP doses had a statistically significant impact on shear force (p=0.013). The values of shear force in two-stage sous-vide without phosphate incorporation were significantly affected by cooking temperature and cooking duration (p<0.001), contrary to what was previously reported in our work (Ismail et al., 2019c). Meaning that, STPP effectively speeds up the dissociation of actomyosin and reduces the toughness of beef semitendinosus. According to Table 3, regardless of cooking temperature or cooking duration, STPP at the highest concentration of 4% considerably reduces the shear force values of all treatments. In contrast, Roldán et al. (2014) found that adding phosphate to sous-vide lamb loins increases the hardness and shear force values. The reasons for these contrasts will be explained based on the total collagen content, myofibrillar, and sarcoplasmic solubility as described in Table 3 and all these parameters were significantly affected by the STPP concentration (p<0.001).

The amount of total collagen is a good indicator of how tender or tough meat will be after being cooked sous vide. However, we discovered that collagen content and collagen solubility play a minor or nonexistent impact in tenderising red meat in our earlier investigation on the cow semitendinosus (Ismail et al., 2019c) and the goat gluteus medius and biceps femoris (Ismail et al., 2019b). Similarly, in the present study, we noticed that cooking temperature and time were less or not correlated to the total collagen (time: r=–0.187; temperature: r=0.125) as well as their interaction (p=0.053) detected no significant effect on collagen content. Interestingly, the addition of phosphate resulted in a significant interaction between the STPP concentration, time, and temperature in total collagen for sous-vide cooked beef (p time×p temperature×p STPP=0.02). Nevertheless, the detected differences in Table 3 for total collagen between treatments were not significant in a Duncan test. The fact that total collagen content decreased with the STPP (p<0.001) and it clearly can be seen that sous-vide treated without phosphate resulted in the highest toughness. According to Chaosap et al. (2021), lower collagen content proportionally lowered the toughness of the meat. However, total collagen was not always used as an indicator to determine the toughness/tenderness, because previous literature has found the effect of myofibrillar components to be more dominant in causing toughness at certain temperatures (Christensen et al., 2000; Ismail et al., 2019c; Ismail et al., 2022; Purslow, 2018).

As mentioned by Christensen et al. (2000), the degree of toughness caused by collagen and myofibrillar proteins occurs through sequential and multi-steps non-proportional to the cooking conditions. Even though gradual cooking between 50°C and 60°C causes collagen denaturation, once the temperature rises over 60°C, the influence of collagen strength is less noticeable and the myofibrillar components start to take over the strength (Purslow, 2018). The phosphate addition could affect both collagen and myofibrillar components (p<0.001). Phosphate can make connective tissues more tender by increasing the solubility of collagen, decreasing the degree to which collagen in connective tissues is cross-linked, and dissociating the actomyosin complex (Shen et al., 2016; Shi et al., 2021). As shown in Table 3, the effect of temperature, time, and STPP concentration is significant on myofibrillar solubility (p<0.001). The myofibrillar solubility decrease with the increased cooking temperature and cooking time and increase with STPP concentration. There are several effects of phosphate on myofibrillar proteins. Phosphates promote the ionic effect, which changes the pH and deviates the pH of proteins from the isoelectric point. Thus, the charges repulse each other and enlarge the space of myofibrils owing to entrapping more water (Shi et al., 2021). Next, due to the buffering effect of phosphate, the pH elevation may play a role in the increased activation of calpains (Shi et al., 2021). The activity of calpains was believed to dissociate the permanent actomyosin bridge. All these mechanisms are thought to contribute to meat tenderness. According to Maqsood et al. (2018), increasing myofibrillar solubility reflects the higher myofibrillar degradation or myofibrillar proteolysis. Thus, lowering the toughness or shear force values (Table 3).

The sarcoplasmic solubility was affected by the time (p<0.001), temperature (p<0.001), STPP concentration (p<0.001), and interaction between temperature and time (p<0.007). The decreased sarcoplasmic proteins based on cooking temperature and time were approximately 15%–42% and 6%–36%, respectively. Increasing temperature and cooking time decreased the sarcoplasmic solubility, this was likely due to the denaturation of actomyosin that causes structural change and presses out the sarcoplasmic protein fluid from the myofibers (Li et al., 2013). However, the addition of phosphate effectively reduces the loss of sarcoplasmic protein as shown in Table 3. The relationship between lower shear force values and higher sarcoplasmic solubility (r=–0.493) can be explained by protein aggregation and water retention. According to Tornberg (2005), the sarcoplasmic proteins aggregated at a temperature between 40°C to 60°C, which is the ideal temperature along with the myofibrillar proteins to provide consistency in the cooked meat. These proteins form a gel and minimize water loss from the meat proteins (Ismail et al., 2019a; Mudalal et al., 2014). According to Li et al. (2013), the higher water retention was substantially connected with the lower shear force value, and the addition of STPP amplifies this correlation (Table 3).

Conclusion

Two-stage cooking time and temperature, and the addition of STPP play a significant effect on the physicochemical properties of beef semitendinosus. Prolonged cooking at higher temperature (45+70°C for 6 h) resulted in greater cooking loss and decreased water content and WHC. Regardless of cooking temperatures and times, the effect of STPP at higher concentrations (0.4%) has been demonstrated by a reduction in cooking loss and an increase in WHC and water content. The effect of phosphate on colour properties was not apparent though it was significant. Cooking temperatures and times had little effect on the shear force values, but STPP made them tender. The shear force values were less connected with total collagen and more correlated with the solubility of the proteins (sarcoplasmic and myofibrillar). Nonetheless, the STPP concentration (0.4%) was effective in dissociating the collagen and producing tender sous-vide meat.

Conflicts of Interest

The authors declare no potential conflicts of interest.

Acknowledgements

This work was supported by research grant UniSZA/2021/GOT/01 (Project No. R0343) from the Universiti Sultan Zainal Abidin, Terengganu, Malaysia.

Author Contributions

Conceptualization: Tang JYH, Ishamri I. Data curation: Ruslan NN, Ishamri I. Formal analysis: Ruslan NN. Methodology: Ruslan NN, Ishamri I. Software: Tang JYH, Ishamri I. Validation: Tang JYH, Huda N, Ismail-Fitry MR. Investigation: Ruslan NN, Ishamri I. Writing – original draft: Ruslan NN, Ishamri I. Writing – review & editing: Ruslan NN, Tang JYH, Huda N, Ismail-Fitry MR, Ishamri I.

Ethics Approval

This article does not require IRB/IACUC approval because there are no human and animal participants.

References

1.

AOAC. 2000 Official methods of analysis. Association of Official Analytical Chemists. Washington, DC, USA: .

2.

AOAC. 2002 Official methods of analysis. 16th edAssociation of Official Analytical Chemists. Arlington, TX, USA: .

3.

Ayub H, Ahmad A. 2019; Physiochemical changes in sous-vide and conventionally cooked meat. Int J Gastron Food Sci. 17:100145

4.

Baldwin DE. 2012; Sous vide cooking: A review. Int J Gastron Food Sci. 1:15-30

5.

Chaosap C, Chauychuwong N, Chauychuwong R, Sriprem C, Sivapirunthep P, Sazili AQ. 2021; Carcass composition, meat quality, calpain activity, fatty acid composition and ribonucleotide content in southern Thai native goats and three-way crossbred goats. Foods. 10:1323

6.

Christensen L, Ertbjerg P, Aaslyng MD, Christensen M. 2011; Effect of prolonged heat treatment from 48°C to 63°C on toughness, cooking loss and color of pork. Meat Sci. 88:280-285

7.

Christensen M, Purslow PP, Larsen LM. 2000; The effect of cooking temperature on mechanical properties of whole meat, single muscle fibres and perimysial connective tissue. Meat Sci. 55:301-307

8.

del Pulgar JS, Gázquez A, Ruiz-Carrascal J. 2012; Physico-chemical, textural and structural characteristics of sous-vide cooked pork cheeks as affected by vacuum, cooking temperature, and cooking time. Meat Sci. 90:828-835

9.

Glorieux S, Goemaere O, Steen L, Fraeye I. 2017; Phosphate reduction in emulsified meat products: Impact of phosphate type and dosage on quality characteristics. Food Technol Biotechnol. 55:390-397

10.

Hoffman LC, Muller M, Vermaak A. 2008; Sensory and preference testing of selected beef muscles infused with a phosphate and lactate blend. Meat Sci. 80:1055-1060

11.

Hunt MC, Sørheim O, Slinde E. 1999; Color and heat denaturation of myoglobin forms in ground beef. J Food Sci. 64:847-851

12.

Ismail I, Hwang YH, Bakhsh A, Joo ST. 2019a; The alternative approach of low temperature-long time cooking on bovine semitendinosus meat quality. Asian-Australas J Anim Sci. 32:282-289

13.

Ismail I, Hwang YH, Bakhsh A, Lee SJ, Lee EY, Kim CJ, Joo ST. 2022; Control of sous-vide physicochemical, sensory, and microbial properties through the manipulation of cooking temperatures and times. Meat Sci. 188:108787

14.

Ismail I, Hwang YH, Joo ST. 2019b; Effect of different temperature and time combinations on quality characteristics of sous-vide cooked goat gluteus medius and biceps femoris. Food Bioprocess Technol. 12:1000-1009

15.

Ismail I, Hwang YH, Joo ST. 2019c; Interventions of two-stage thermal sous-vide cooking on the toughness of beef semitendinosus. Meat Sci. 157:107882

16.

Joo ST. 2018; Determination of water-holding capacity of porcine musculature based on released water method using optimal load. Korean J Food Sci Anim. 38:823-828.

17.

Lamkey JW, Mandigo RW, Calkins CR. 1986; Effect of salt and phosphate on the texture and color stability of restructured beef steaks. J Food Sci. 51:873-875

18.

Lawrie RA, Ledward DA. 2006 Lawrie’s meat science. 7th edWoodhead. Sawston, UK:

19.

Li C, Wang D, Xu W, Gao F, Zhou G. 2013; Effect of final cooked temperature on tenderness, protein solubility and microstructure of duck breast muscle. LWT-Food Sci Technol. 51:266-274

20.

Long NHBS, Gál R, Buňka F. 2011; Use of phosphates in meat products. Afr J Biotechnol. 10:19874-19882

21.

Mac Donald GA, Lanier TC. 1994; Actomyosin stabilization to freeze-thaw and heat denaturation by lactate salts. J Food Sci. 59:101-105

22.

Maqsood S, Manheem K, Gani A, Abushelaibi A. 2018; Degradation of myofibrillar, sarcoplasmic and connective tissue proteins by plant proteolytic enzymes and their impact on camel meat tenderness. J Food Sci Technol. 55:3427-3438

23.

Mudalal S, Babini E, Cavani C, Petracci M. 2014; Quantity and functionality of protein fractions in chicken breast fillets affected by white striping. Poult Sci. 93:2108-2116

24.

Önenç A, Serdaroğlu M, Abdraimov K. 2004; Effect of various additives to marinating baths on some properties of cattle meat. Eur Food Res Technol. 218:114-117

25.

Przybylski W, Jaworska D, Kajak-Siemaszko K, Sałek P, Pakuła K. 2021; Effect of heat treatment by the sous-vide method on the quality of poultry meat. Foods. 10:1610

26.

Purslow PP. 2018; Contribution of collagen and connective tissue to cooked meat toughness; some paradigms reviewed. Meat Sci. 144:127-134

27.

Roldán M, Antequera T, Martín A, Mayoral AI, Ruiz J. 2013; Effect of different temperature–time combinations on physicochemical, microbiological, textural and structural features of sous-vide cooked lamb loins. Meat Sci. 93:572-578

28.

Roldán M, Antequera T, Pérez-Palacios T, Ruiz J. 2014; Effect of added phosphate and type of cooking method on physico-chemical and sensory features of cooked lamb loins. Meat Sci. 97:69-75

29.

Shen QW, Swartz DR, Wang Z, Liu Y, Gao Y, Zhang D. 2016; Different actions of salt and pyrophosphate on protein extraction from myofibrils reveal the mechanism controlling myosin dissociation. J Sci Food Agric. 96:2033-2039

30.

Shi H, Shahidi F, Wang J, Huang Y, Zou Y, Xu W, Wang D. 2021; Techniques for postmortem tenderisation in meat processing: Effectiveness, application and possible mechanisms. Food Prod Process Nutr. 3:21

31.

Tan SM, de Kock HL, Dykes GA, Coorey R, Buys EM. 2018; Enhancement of poultry meat: Trends, nutritional profile, legislation and challenges. S Afr J Anim Sci. 48:199-212

32.

Tornberg E. 2005; Effects of heat on meat proteins: Implications on structure and quality of meat products. Meat Sci. 70:493-508

33.

Trbovich VR, Garza H, Wick MP, England EM, Garcia LG. 2018; Investigating the effects of sous vide cooking on tenderness and protein concentration in young fed and cow semitendinosus muscles. Meat Muscle Biol. 1:34