The material selected for this study was fresh apples of the “Grey Reinette” variety purchased from the local market. Skin and inedible parts were removed from the raw material. Cubes with an edge length of 15 mm were cut out of the flesh. Immediately after cutting, the raw material was frozen by hydrofluidization, immersion, and, for comparative purposes, chamber freezing.
Hydrofluidization and immersion freezing of the prepared samples were conducted on the test rig used previously for fluid flow (Palacz et al., 2019) and thermal experiments (Palacz et al., 2021) on the hydrofluidization method. The multistage pump (EBARA EVMSL 5) with inverter (TECO L510s) allowed for maintaining smooth control of the flow conditions for immersion freezing, where only fluid circulation within the heat exchanger was needed, as well as for hydrofluidization freezing, characterized by the intensive flow. The tank in which the freezing was conducted had a volume of approximately 19 l. During freezing, the samples were stored in a box made of a plastic net to ensure that all samples were removed at the same time. Figure 1 presents the hydrofluidization process schematically where the fluid flow from the orifices allows the immersed products to move and the heat and mass transfer between the fluid and frozen samples occur.
The mass flow rate of both solutions used as freezing medium and pumped into the tank was controlled based on the Coriolis-type mass flow meter (Endress + Hauser Promass E300) installed next to the pump as described in detail by Palacz et al. (2019). Such equipment guaranteed an accuracy of ±2% for the mass flow rate measurement in the conditions used for both hydrofluidization and immersion freezing compared in this study. The product temperature during the freezing process was monitored with T-type thermocouples. That sensor was installed in the center of the sample to assess the freezing time. It should be noted that due to the sample movement during the hydrofluidization process, the position of the thermocouple could change slightly, therefore the issue of thermocouple location was the main challenge for thermal measurements. To minimize the effect of the thermocouple's imperfect position or possible movement, thermocouples having a sheath diameter of 0.5 mm were used. It allowed them to be seamlessly inserted into analyzed samples. Moreover, the thermocouple sheath was marked with a permanent marker to keep the same depth while inserting the thermocouple into samples during each run. To follow good practices for preparing experiments, the same researcher installed thermocouples during each run, and every test was repeated at least five times to minimize effects related to possible variation in temperature position and other unavoidable factors.
Nevertheless, the mentioned temperature readings were used to estimate the freezing time for the experiments carried out. Two different water-based solutions were used for both freezing methods, i.e., the ethanol solution having a mass concentration of 50% and the glycerol solution having a concentration of 40%. Ethanol-based solutions were previously used in numerical and experimental studies related to hydrofluidization freezing (Fikiin, 2008; Verboven et al., 2003; Palacz et al., 2021; Stebel et al., 2021) with a good result; however, a more viscous glycerol-based solution was analyzed only in numerical studies of the hydrofluidization method (Stebel et al., 2022a; Stebel et al., 2022b). In the case of a glycerol solution, during freezing experiments, the flow rate was set at 2.2 l/min and 45.2 l/min for the immersion and hydrofluidization freezing processes, respectively. In the case of the ethanol solution, the flow rates were 1.4 l/min and 28.8 l/min for immersion and hydrofluidization methods. The flow rates were different for the two solutions to keep a similar Reynolds number for both fluids. The temperature of the freezing medium was maintained at the level of -16°C. Approximately 160 g of apple samples were frozen for each experimental run. After the freezing process, food samples were immediately transported from one laboratory where the freezing was carried out to another laboratory where quality tests were performed by using a portable medical freezer (DOMETIC CDF-11) and stored until the next day. The temperature of the samples was maintained at -16°C during the transportation and storage phases. The total mass of the apple cubes used for the quality analysis was equal to 500 g for both, hydrofluidization and immersion freezing methods. In addition, a similar number of samples was frozen using chamber freezing. The chamber freezing temperature was carried out at the level of -16°C under conditions of natural convection. Despite efforts to minimize the sample temperature variation at each step of the conducted research (freezing, transportation, and storage), some temperature variation was expected. In order to assess the accuracy of the conducted temperature measurements the uncertainty analysis was performed. The Type-B (Eq. (1)) uncertainty for the T-type class I thermocouples and National Instruments acquisition system was calculated.
$$\:{u}_{{B}_{i}}=\frac{{I}_{i}}{\sqrt{3}}$$
1
where \(\:{u}_{{B}_{i}}\)states for Type-B uncertainty and \(\:{I}_{i}\) is the accuracy of measurement equipment.
Then, the combined standard uncertainty was calculated (Eq. (2)).
\(\:U\left(T\right)=\sqrt{\sum\:_{i=1}^{n}{{u}_{B}}_{i}^{2}}\) where U(T) is the combined standard uncertainty that includes all the equipment.(2)
The accuracy of the applied instrumentation as well as evaluated uncertainty are listed in Table 1.
Table 1
The accuracy of the temperature measurements.
Instrumentation | Accuracy | \(\:{\varvec{u}}_{\varvec{B}}\) |
T-type thermocouple class I | ± 0.5°C | ± 0.29°C |
National Instruments 9214 module | ± 0.4°C | ± 0.23°C |
\(\:U\left(T\right)\) | ± 0.37°C |
As it can be seen in that table, the combined standard uncertainty was below ±0.5°C. On the other hand, the errors related to the misplacement of the thermocouple were not included in the calculations above. Therefore, the expanded uncertainty (coverage factor equal of 3) equal to ±1.1°C was used to evaluate the accuracy of the temperature measurements. On the other hand, that factor is related just to the temperature reading not the temperature variation during the transportation. In that case, it was assumed that the temperature variation does not exceed ±0.5°C due to their heat capacity.
The difference between immersion and hydrofluidization methods is the convection regime. In immersion freezing, the liquid medium cools down the food product by natural convection, while the hydrofluidization is based on forced convection. The food samples analyzed in this study were frozen in the glycerol solution under conditions with a heat transfer coefficient of 1055 W/(m2K) and 190 W/(m2K) for the hydrofluidization and immersion freezing, respectively. In the case of the ethanol solution, the heat transfer coefficients for these two methods were 835 W/(m2K) and 154 W/(m2K). For the latter solution, the values were lower because the viscosity of this medium is lower, which determines the Prandtl number. To determine the heat transfer coefficients for hydrofluidization freezing, the correlation proposed by Whitaker (1972) was used as in previous studies of this method (Orona et al., 2017; Stebel et al., 2022a). The bulk velocity of the fluid flowing towards the food products was determined based on the measured volumetric flow rate of a fluid and a known cross-section area of the tank where the fluid flow had a vertical direction (100x60 mm) according to Eq. (3):
where \(\:v\) is the bulk velocity of the fluid, \(\:\dot{Q}\) is the volumetric flow rate of the fluid based on measurements, and \(\:A\) is the cross-section area of the tank region where the flow was observed.
The characteristic length for Reynolds and Nusselt number determination was a cube edge length (15 mm) which reports for the food product according to Whitaker’s correlation and the studies referred to above. Reynolds number given in Eq. (4) describes whether the flow is turbulent, and Prandtl number given in Eq. (5) is related to the fluid properties.
$$\:\text{R}\text{e}=\frac{\rho\:\:v\:L}{\mu\:}$$
4
where Re is the Reynolds number, \(\:\rho\:\) is the fluid density, \(\:L\) is the characteristic length of the frozen sample, and \(\:\mu\:\) is the dynamic viscosity of the fluid.
$$\:\text{P}\text{r}=\frac{{c}_{p}\:\mu\:\:}{k}$$
5
where Pr is the Prandtl number, \(\:{c}_{p}\) is the specific heat capacity, and \(\:k\) is the thermal conductivity of the fluid, respectively.
The mentioned Nusselt number for the hydrofluidization freezing was based on the correlation by Whitaker (1972) given in Eq. (6):
$$\:\text{N}\text{u}=2+(0.4\:{\text{R}\text{e}}^{0.5}+0.06\:{\text{R}\text{e}}^{0.67}){\text{P}\text{r}}^{0.4}{\left(\mu\:/{\mu\:}_{w}\right)}^{0.25}$$
6
where Nu is the Nusselt number and \(\:\left(\mu\:/{\mu\:}_{w}\right)\) is the ratio of the viscosity of the fluid at bulk temperature and the cooled object wall temperature (assumed in this study as 1 due to immediate reduction of the food sample surface temperature).
In the case of immersion freezing, the correlation of Kramers (1946), given in Eq. (7) below was used which is more suitable for a still flow of the fluid:
$$\:\text{N}\text{u}=(0.35+0.56\:{\text{R}\text{e}}^{0.52}){\text{P}\text{r}}^{0.3}$$
7
Eventually, after determining the Nusselt number from one of the correlations assumed for hydrofluidization and immersion freezing, the convective heat transfer coefficient can be evaluated using Eq. (8):
$$\:h=\frac{\text{N}\text{u}\:k\:}{L}$$
8
where \(\:h\) is the heat transfer coefficient for immersion or hydrofluidization freezing.
All the values for thermodynamic parameters used in Eqs. (4)-(8) for both fluids are summarized in Table 2.
Table 2
Thermodynamic properties of fluids used for heat transfer coefficient determination evaluated in -16°C (Melinder, 2010).
Parameter | Unit | Water-glycerol (50%) | Water-ethanol (40%) |
Density | kg/m3 | 1 142 | 959.8 |
Dynamic viscosity | kg/(m·s) | 0.03521 | 0.01852 |
Thermal conductivity | W/(m·K) | 0.3926 | 0.3453 |
Specific heat capacity | J/(kg·K) | 3 024 | 3 745 |
In the case of the reference chamber freezing method, the heat transfer coefficient reaches a significantly lower range of 6–9 W/(m2K) according to Fellows (2000) due to the still air used as a medium.
Analyzes
Analysis of residual ethanol content by the SPME-GC-FID (Solid Phase Microextraction-Gas Chromatography-Flame Ionization Detector) method
The sample was transferred to a 15 ml vial (Supelco). Then 2 ml of distilled water with 40 mg/l of 4-methyl-2-pentanol (Fluka) and 1 g of NaCl were added to the vial. The method of extracting the compound was performed on the SPME fiber (Supelco INC., Bellefonte PA, USA) according to the methodology prepared by Satora et al. (2008).
Analysis of residual glycerol content by the UHPLC-IR (Ultra-High Performance Liquid Chromatography) method
Glycerol content was determined by UHPLC using a Shimadzu NEXERA XR series device with an RF-20A refractometer detector (Kyoto, Japan). Separation was carried out using a Shodex Asahipak NH2P-50 4.6 × 250 mm column (Showa Denko Europe, Germany), thermostated at 30°C. The mobile phase consisted of aqueous acetonitrile (70%). The isocratic elution program (0.8 ml/min) lasted 16 minutes (Satora and Pater, 2023). Quantitative determinations were made using standard curves prepared for glycerol.
Analysis of vitamin C by HPLC
The vitamin C content was determined as the sum of L-ascorbic acid and L-dehydroascorbic acid by HPLC (PN-EN 14130:2003) using a Merck-Hitachi HPLC-System LaChrome with UV/VIS detector (model 7420), degasser (model L-7612), programmable autosampler (model L-7250), pump (model L-7100) and column oven thermostat (model L-7360). For the identification of vitamin C and its quantitative analysis, an external standard was used, which consisted of L-ascorbic acid dissolved in 2% metaphosphoric acid.
Analysis of selected polyphenols by the UHPLC method
A 10 g of food material was weighed and mixed with 20 ml of 70% methanol containing 0.5% HCl. The sample was crushed with a homogenizer for 1 min and then centrifuged at 4,000 rpm (temperature 5°C, 10 min) (Tarko et al. 2017).
A liquid chromatograph of the Shimadzu NEXERA XR series (Kyoto, Japan) with a DAD detector was used for analysis. The separation was carried out on a Synergi Fusion RP-80A 150×4.6 mm (4 µm) Phenomenex column (Torrance, CA, USA), thermostated at 30°C. Acetic acid with a concentration of 2.5% (solution A) and acetonitrile (solution B) was used as the mobile phase. Gallic acid, ellagic acid, (+) catechin and phloridzin were detected at λ = 280 nm, caffeic acid, p-coumaric acid, chlorogenic acid, ferulic acid, vitexin and resveratrol at λ = 325 nm, quercetin, quercetin glucoside, rutin and kaempferol at λ = 360 nm, and hippuric acid, protocatechuic acid, and hyperoside at λ = 250 nm (Tarko et al., 2020). Standard curves prepared for ferulic acid, hippuric acid, caffeic acid, p-coumaric acid, gallic acid, chlorogenic acid, (+) catechin, quercetin (Sigma), protocatechuic acid, ellagic acid, phloridzin, apigenin-8-glucoside (vitexin), resveratrol, rutin, kaempferol, delphinidin-3-O-glucoside (myrtyline), cyanidin-3-O-galactoside (ideaine), cyanidin-3-O-glucoside (curomanine), cyanidin, cyanidin-3-O-rutinoside (keracyanin), pelargonidin-3-O-glucoside (calistefin), peonidine-3-O-glucoside and quercetin-3-O-glucoside (Extrasynthese, France) were used for quantitative determinations.
Polyphenol oxidase (PPO) activity
The polyphenol oxidase activity determination was performed according to the method reported by Cano et al. (1997). The products were grounded (Mikrotron MB 550 Kinematica AG, Switzerland) with pH 7 phosphate buffer extracted for 2 hours at 4°C and centrifuged (4200 g, 10 min, MPW-352RH, MPW Med. Instruments, Poland). To the sample containing active polyphenol oxidase, a solution of catechol in phosphate buffer was added. The determination of polyphenol oxidase activity was based on determining the increase in absorbance at a wavelength of λ = 420 nm (spectrophotometer Shimadzu UV-160 UV-VIS, Tokyo, Japan). Results were expressed in units of polyphenol oxidase activity corresponding to the increase in optical density of the substrate in 1 min per 1 g of raw material.
Color analysis
The color in the CIE L*a*b* system was analyzed using a Konica-Minolta CM-5 spectrophotometer (Japan) with the following measurement parameters: D65, 10º, SCE, calibrated on white and black standards. Based on the obtained results, the color change (ΔE) was calculated (Kriaa and Nassar, 2022) between the samples frozen and thawed at 20ºC for 1 hour. The cubes obtained from fresh apples were stored for 1 hour under the same conditions and ΔE was calculated for them.
Drip loss assessment
Drip loss analysis was performed by placing weighted frozen samples between layers of filter paper in hermetically sealed plastic bags. The thawing and storage of the samples (5 hours) was carried out in a thermostatic chamber at 4°C. Drip loss was expressed as the percentage of moisture contained in the paper with respect to the weight of the sample before thawing (Charoenrein and Owcharoen, 2016).
Texture analysis
Instrumental texture analysis was performed using a TA-XT2 texture meter (Stable Micro Systems, United Kingdom). The force needed for 50% compression of the frozen sample was recorded using the P/45 probe with a movement speed set to 0.5 mm/s. The comparison of the samples was made by analyzing the maximum force and compression work recorded during the 50% sample compression test (Charoenrein and Owcharoen, 2016).
Dry mass analysis
The dry mass was determined using a moisture analyzer (Radwag MA 50.R.NS) at 105°C.