Next Article in Journal
Design of Position Control Method for Pump-Controlled Hydraulic Presses via Adaptive Integral Robust Control
Next Article in Special Issue
Cryoconcentration by Centrifugation–Filtration: A Simultaneous, Efficient and Innovative Method to Increase Thermosensitive Bioactive Compounds of Aqueous Maqui (Aristotelia chilensis (Mol.) Stuntz) Extract
Previous Article in Journal
Biogenic Amine Content in Retailed Cheese Varieties Produced with Commercial Bacterial or Mold Cultures
Previous Article in Special Issue
Evaluation of Wound Healing Potential of Novel Hydrogel Based on Ocimum basilicum and Trifolium pratense Extracts
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Effect of Drying Techniques on the Physical, Functional, and Rheological Attributes of Isolated Sunflower Protein and Its Hydrolysate

1
School of Food and Biological Engineering, Jiangsu University, Zhenjiang 212013, China
2
Department of Agricultural and Biosystems Engineering, Faculty of Agriculture, Benha University, Moshtohor, Qaluobia 13736, Egypt
3
Department of Food Science and Nutrition, College of Sciences, Taif University, P.O. Box 11099, Taif 21944, Saudi Arabia
4
CSIR—Food Research Institute, Accra P.O. Box M20, Ghana
5
Faculty of Technology and Technical Sciences, St. Kliment Ohridski University-Bitola, Dimitar Vlahov, 1400 Veles, North Macedonia
*
Authors to whom correspondence should be addressed.
Submission received: 29 November 2021 / Revised: 18 December 2021 / Accepted: 20 December 2021 / Published: 22 December 2021

Abstract

:
The influence of freeze and convection (at 40 and 50 °C) drying on the physical, functional, and rheological attributes of sunflower protein (SP) and its hydrolysate (SPH) was investigated. Compared with convectively-dried samples, the lightness, turbidity, bulk density, and particle size values of the freeze-dried SP and SPH were substantially higher, but the browning index was lower (p < 0.05). Additionally, freeze-dried samples exhibited good solubility and foaming characteristics, whereas lower emulsion properties with the most pH values were observed. Furthermore, SPHs possessed higher solubility as well as foamability over SPs under varying pH values (2.0–10.0), whilst reduction in the emulsion activity index was clearly observed (p < 0.05). Convectively-dried powders exhibited greater viscosity and consistency coefficient; and significantly lower flow behavior index of dispersions, relative to the respective freeze-dried preparations, indicating that dehydration methods influenced the flow behavior of the investigated samples. From a molecular weight analysis, convectively-dried samples at various temperatures were characterized with high proportion of small-sized particles at ≤1 kDa fractions over the respective powders obtained by freeze drying. The observations made, thus, would benefit food processors and manufacturers in electing better dehydration technique based on the desired traits of SP and SPH powders for successful application in food product formulations.

1. Introduction

The recent increase in awareness of the nutritional and functional benefits of animal and/or vegetable protein, by the global populace, has resulted in a commensurate increase in the use of the macromolecule (in human diet). Production of vegetable proteins, in contrast with animal proteins, requires less land, fossil energy and water resources; and thus, make it more sustainable [1]. Soy protein is currently the most available plant proteins worldwide. To meet future demands of a more affordable, diverse, and high-quality plant protein, other protein resources must (as matter of importance) be explored. Sunflower, an important oilseed crop in the world, is largely cultivated on five continents (Asia, Africa, Europe, North and South America) with production of 56.07 million tons in 2019 [2]. Two main products are obtained from sunflower seeds which are, first, oil (mostly utilized for human nutrition) and, second, meal (primarily underexploited/considered as waste-product). Sunflower meal contains a high content of protein (300 to 500 g/kg), which indicates good potential for use as cheap and sustainable protein source [3]. Sunflower protein (SP) is low in anti-nutritional substances (e.g., cyanogens, protease inhibitors, lectins, and goitrogenic factors), and is devoid of toxic compounds [4], and its amino acids composition (except lysine) complies mainly with the FAO recommendation [5], making it an alternative source of protein. Literature also indicates that sunflower protein/hydrolysate has various bioactivate functions including ACE-inhibition [6], antioxidant [7], and antimicrobial [8] activities. Additionally, the functionalities of SP are close to soy protein, making it a substantial and sustainable natural protein-resource for new food formulations.
The functionalities of proteins are important indicators for food preparation, processing, and consumption. Native SP (from the meal), however, does not have desirable functional traits (such as solubility), observably under its isoelectric point due to the heating steps used throughout oil extraction from the seed [9]. Efforts to enhance the techno-functionalities of SP applying physical [9], chemical [10] and enzymatic approaches [11] are reported. Application of proteolysis (enzymolysis) is essentially considered a suitable and valuable approach for enhancing the functionality of native proteins and keeping their nutritive attributes by preparing peptides with high antioxidative action [12,13].
Furthermore, successful application of protein and/or hydrolysate in food processing basically relies on their functionality, which are mainly dependent on the dehydration/drying techniques. Dehydration techniques (including freeze and convection oven drying) are vital in the preparation of protein and hydrolysate powders aimed at prolonging their shelf-life, and reducing transportation and handling/storage costs. Freeze drying (FD), named as cryodesiccation or lyophilization, is a sophisticated drying method, which includes three (3) stages: freezing, sublimation, and thereafter desorption. FD minimizes the microbiological reactions and ceases most of the deterioration [14], as well as prevents denaturation and the Maillard reaction of protein/hydrolysate [15]. This technique, therefore, is generally used for dehydrating heat-sensitive materials (e.g., protein/hydrolysate) to analyze the physicochemical attributes and functionality/bioactivity of such materials. Convection drying is a comparatively low-cost technique and the dehydration can be performed below the denaturation temperature of proteins, but the residence dehydration time is much longer [16]. However, these drying techniques may, characteristically, induce the limited denaturation of proteins by altering their structures differently and, likewise, impact the functionality and bioactivity of the proteins and/or hydrolysates. It has, previously, been indicated that dehydration techniques have remarkable effects on the rheological, functional, and structural traits of lentil [17], peanut [18], and mung bean [16] protein, as well as egg white hydrolysate [19]. Understanding the alterations in physical, and rheological attributes of protein/hydrolysate and connecting them with the changes in functionality provide strong insights on their behavior during food formulation. Nevertheless, to date, no comparative investigation with in-depth information/analysis on the characteristics of SP and its hydrolysate as affected by various drying procedures has been reported. Motivated by this background, this study compares the impact of two dehydration techniques (freeze and convection oven drying) on the physical, functional, and rheological traits of SP and its hydrolysate.

2. Materials and Methods

2.1. Materials

Alkaline protease 2.4 LFG (150,000 U/mL activity), which was used to prepare the SP hydrolysates (SPHs) was acquired from Novozymes Bio-Technology Company Ltd. (Tianjing, China). Sunflower meal/by-product was generously obtained from Xinjiang Jinhai Oil Company Ltd. (Xinjiang, China). The by-product was powdered (to ≤60-mesh size) using a DFT-100A portable mill, then kept in zip bags (4 °C) for other processing. The content of SP in the meal was 29.31% (Kjeldahl method).

2.2. Preparation of SP and SPH

SP was isoelectrically (pH 4.5) precipitated and extracted as detailed (earlier) in our research [20] with slight changes. Following extraction, SP suspension was first divided into two (2) equal aliquots. The first aliquot was brought to pH 7.0 and then subjected to various dehydration techniques (Section 2.3) to produce dried SPs. The second aliquot was hydrolyzed (50 °C, pH 9.0, alkaline protease 0.32 mL/LSP-suspension, impeller agitator 100 rpm and 90 min) as outlined previously in our study [21]. The pH was set by 1.0 M sodium hydroxide. After proteolysis of SP, the enzyme action was terminated (15 min, 90 °C). Hydrolysates (at 19.92% degree of hydrolysis) were neutralized (at 25 °C), centrifuged (5810 R, Eppendorf AG Barkhausenweg 122339, Hamburg, Germany—5000 rpm, 15 min), and thereafter the supernatants (pH 7.0) were dried (Section 2.3).

2.3. Drying Process of SP and SPH

Two drying techniques (freeze drying and convection oven drying) were applied to dehydrate the SP and SPH solutions. To prepare the freeze-dried powder (with ≤5% moisture content), SP and SPH suspensions were first frozen (24 h) on petri dishes at −18 °C, and afterward lyophilized (for ~48 h) under 0.096 mBar vacuum at −81 °C using a Lyo-Quest-85-Plus freeze dryer (Telstar Lyo-Quest, Spain).
Convection oven drying was conducted at two different temperatures using a ventilated air-drying oven (DHG-9123A, Blue Pard, Shanghai Yiheng Scientific company Ltd., Shanghai, China). The temperature was set at 40 and 50 °C (for ~60 and 48 h, respectively to realize a moisture content ≤5% in the dried samples). Air velocity/circulation was 1.0 m/s. The basis for selecting these drying temperatures was due to the fact that the lower temperature (<40 °C) was noted to extend the dehydration time resulting in microbial spoilage, and the higher temperature (>50 °C) could negatively affect the functionality and bioactivity of samples. Resulting SP and SPH powders were pulverized (to ≤40-mesh size) using a DFT-100A portable mill, then kept at −20 °C for characterization.

2.4. Analysis of Color and Browning Index (Bindex)

Color traits (including L*, a* and b*) of all dried samples were examined with CR-400 colorimeter (Hangzhou Ke Sheng Instrument Company Ltd., Osaka, Japan). The L* is the luminosity/lightness component, varying from 0 (black) to 100 (white), a* and b* values range from −100 to + 100 green (−)/red (+) and blue (−)/yellow (+), respectively.
Bindex was estimated as outlined by the method of Ding and Ling [22] applying Equation (1):
B index = N 0.31 0.172 × 100
where
N = a * + 1.75 L * 5.65 L * + a * 0.301 b *

2.5. Particle Sizing (Psizing) and Bulk Density (Bdensity)

The Psizing of SPs and hydrolysates (SPHs) was quantified (at 23 ± 1 °C) using an Anton Litesizer-500 (Austria). Freeze and oven-dried SP and SPH were dissolved in distilled H2O (1 mg/mL, pH 7.0) before determining Psizing.
Bdensity was determined by transferring the dehydrated samples (5 g) into graduated cylinders (25 mL) and tapped gently until the volume of SP and SPH remained constant. The volume was estimated, and the Bdensity was thereafter recorded as g/mL.

2.6. Turbidity

The turbidity (Ttur) of SP and SPH solutions (1.4 mg/mL in 0.05 M (pH 7.0) phosphate buffer) was assayed by determining the absorbance at an ambient temperature using a spectrophotometer at 600 nm. The absorbance was then utilized as a Ttur index [13].

2.7. Protein Solubility

Solubility (Psolubility) of SP and SPH suspensions (3.0%, w/v) was assessed under varying pH from 2.0 to 12.0. The respective dispersions (solubilized) were subjected to centrifugation (15 min, 4000 rpm), and the amount of protein was assayed in supernatants through the protocol described by Lowry and colleagues [23]. Bovine serum albumin was employed as the standard.

2.8. Foaming Attributes

Foaming capacity (Fcapacity) and foam stability (Fstability) of SP and SPH solutions (4 g in 100 mL distilled H2O) were examined under varied pH (2.0–12.0). Respective suspensions were blended (at 10,000 rpm) using FSH-2A-homogenizer (Fang Ke Instrument (Changzhou) Company Ltd., Jiangsu, China) for 5 min, and transferred immediately into calibrated cylinders. The total volume of the resultants (after 30 s and 20 min) was recorded to quantify Fcapacity and Fstability, respectively. Computation was performed by the following Equations (3) and (4):
F capacity   % = Volume   after   whipping   30   s Initial   volume Initial   volume × 100
F stability   % = Volume   after   whipping   20   min Initial   volume Initial   volume × 100

2.9. Emulsion Properties

Emulsion activity index (EAindex) and emulsion stability index (ESindex) of samples (SPs and SPHs) were measured with the method of Liu et al. [24] with slight alterations. Five milliliters of sunflower oil and 15 mL of (1%) sample dispersions under varied pH (2.0–12.0) were homogenized (FSH-2A, Fang Ke Instrument (Changzhou) Company Ltd., Jiangsu, China) for 1 min (at 10,000 rpm). The aliquot of the emulsion (100 μL) was directly picked from the bottom of the tube and mixed with SDS (10 mL, 0.1%) at zero and 10 min after homogenization. Absorbance of the resultant (at 500 nm) was analyzed at zero min (S0) and 10 min (S10) subsequent to emulsion formation using spectrophotometer. EAindex and ESindex were computed as:
EA index m 2 / g = 2 × 2.303 × 100 × S 0 10000 × 0.25 × P c
ES index min = S 0 × 10 S 0 S 10
where Pc (g/mL) depicts SP and SPH concentration.

2.10. Apparent Viscosity (Aviscosity) and Shear Stress (Sstress)

The Aviscosity and Sstress of SP and SPH suspensions (1 mg/mL in deionized H2O) were examined using a DHR-1-rheometer (WatersTM, TA Instruments Company, New Castle, DE, USA). The DHR-1-rheometer parameters were set at duration 120.0 s, 25 °C, linear mode, shear rate from 0.01 to 100.0 s−1 and test interval of 1000 μm. The correlation among the shear rate and shear stress was modeled by applying the power law equation [25] as follows:
τ = K γ n
where τ (Pa) represents shear stress; γ (1/s) represents shear rate; K (Pa.s) represents consistency coefficient; and n depicts flow behavior/attribute index.

2.11. Molecular Weight (MWeight) Distribution

Following the procedure of Wang et al. [26] with minor modifications, HPLC system (1525-Waters, Thermo. Scientific Inc., Waltham, MA, USA; TSKgel SWXL-2000, 30 × 0.78 cm; Tosoh, Japan) was utilized to characterize the Mweight distribution of freeze- and oven-dried SP and SPH. Respective suspensions were eluted (flow rate—1.0 mL/min) at 220 nm. Standards with known Mweight were applied for the calibration curve: bovine serum albumin (67,000 Da), peroxidase (40,200 Da), ribonuclease A (13,700 Da), glycine tetramer (246 Da) and p-aminobenzoic acid (137.14 Da). The SP and SPH were separated into six (6) fractions: Mweight > 10,000, 10,000–5000, 5000–3000, 3000–1000, 1000–500, and <500 Da, and the content of the fractions expressed in percent (%).

2.12. Statistical Analysis

Experimental results (in triplicates) were subjected to ANOVA by Minitab (Version 18.0, Minitab Inc., Pennsylvania, PA, USA) to assay the differences (at p < 0.05) in averages following various dehydration methods. Pearson’s correlation (by XLSTAT-V1.0-2016 software—XLSTAT Inc., New Yourk, NY, USA) was performed to quantify the interrelation between the physical, functional, and rheological parameters.

3. Results and Discussion

3.1. Physical Attributes

Color property plays a substantial sensory role in the acceptability of plant proteins (e.g., SP and its hydrolysate) for food preparations. Color attributes of SP and SPH prepared by freeze and convection (40 and 50 °C) drying were presented in Table 1. Remarkable changes were recorded (p < 0.05) in the color parameters (noticeably L*) of dried SPs and their hydrolysates (SPHs), suggesting that the color of dried samples was influenced by dehydration techniques and enzymolysis. Among the dried samples, freeze-dried SP (FDSP) and SPH (FDSPH) were noted to be lighter than the other respective powders obtained by convective drying under varied temperatures. The convectively-dried (CD) SP and SPH (especially at 50 °C) showed maximal a* and b*, and consequently the least L* values, reference to the freeze-dried (FD) samples; illustrating the CD powders were darker. These observations are possibly linked to the oxidation of the convectively-dried SP and SPH due to the direct contact between samples and oxygen (O2) under high temperature throughout the drying time. This also may be ascribed to the low temperature, pressure and O2 in freeze drying, contributing to high luminosity. Such findings were also indicated for chickpea protein in earlier research [27]. Furthermore, the L* value of dehydrated powders was notably reduced (p < 0.05) from 50.82 ± 0.89 to 38.27 ± 0.12 (for freeze-dried SP), 42.36 ± 0.43 to 34.12 (for CDSP40) and 37.25 ± 0.45 to 30.48 ± 0.31 (for CDSP50) following proteolysis, indicating loss of luminosity. The reduction in luminosity is possibly correlated with unfavorable chemical modifications during enzymatic hydrolysis (such as Maillard reaction—RMaillard) [28,29]. This can clarify the noticeable decrease in luminosity of SPHs (particularly convectively-dried SPH at 50 °C—CDSPH50). This outcome is also consistent with the higher Bindex (Table 1) for the freeze- and convectively-dried SPH compared with the dehydrated SPs. Finally, such results indicated that freeze drying was the most efficient technique in controlling the Maillard reaction and the luminosity of the samples.
Table 1 shows the browning index of SP and SPH obtained by different dehydration techniques. Results indicated that FDSP isolate displayed a lower (p < 0.05) Bindex than the remaining samples. Contrarily, CDSPH50 had the highest browning intensity implying it was the darkest among the other preparations. Moreover, the Bindex of freeze and convectively (at 40 and 50 °C) dried SPH was respectively 59.60, 53.45, and 54.07% greater than FDSP, CDSP40, and CDSP50. The high Bindex of the SPH (notably CDSPH50), relative to the remaining powders, can be attributed to the fact that the enzymolysis and convective drying accelerated the interaction of aldehydes and free amino clusters via RMaillard, inducing a browner product than the other samples. Similar observation regarding the effect of convection oven drying on Bindex of mung bean protein has been reported [16]. The intense RMaillard in the convectively-dried hydrolysates was further confirmed by its high Bindex (Table 1).
For the Psizing of the SP and SPH powders (Table 1), all the samples had Psizing lower than 550 nm. Freeze-dried SP and SPH showed higher Psizing (548.61 ± 21.93 and 388.55 ± 15.69 nm, respectively) than convectively-dried samples (p < 0.05). This phenomenon can be linked to the formation of ice crystal(s), inducing the aggregates of particles and recombining into bigger Psizing during the lyophilization [30], thus FDSP and FDSPH showed the highest Psizing over the respective CDSP and CDSPH under varied temperatures. Further, lower Psizing values (p < 0.05) were observed for dehydrated SPHs (especially CDSPH50—280.71 ± 6.47 nm) with reference to the SP powders. This implied that enzymolysis caused a breakdown of SP molecules (as supported by the observations of Mweight analysis (Section 3.6), resulting in the release and/or formation of small-sized peptide(s). Similar outcomes due to freeze and convection oven drying [16,27], as well as enzymolysis [13] on Psizing were recorded (by other researchers).
The Bdensity of SPs and their hydrolysates was substantially influenced (p < 0.05) by the drying and proteolysis techniques. The highest Bdensity (395.27 ± 2.73 kg/m3) was noticed for FDSP followed by CDSP40, CDSP50, FDSPH, CDSPH40, and CDSPH50 (Table 1). The observed variations can be linked to the dehydration method/temperature, enzymolysis, particle porosity and Psizing distribution. The outcomes of Bdensity of freeze- and convectively-dried SP and SPH were consistent with the observations of Psizing of the corresponding samples. Rudra et al. [31] and Joshi et al. [17] observed comparable Bdensity for freeze-dried cowpea protein (428.5 kg/m3) and lentil protein (276 kg/m3), respectively.
The effect of dehydration approaches on the turbidimetric value of SP and SPH is presented in Table 1. In general, turbidities of freeze-dried SP and SPH were markedly higher than the respective convectively-dried powders under varying temperatures (p < 0.05). This finding may be credited to the higher Psizing of lyophilized samples. On the other hand, freeze- and convectively (at 40 and 50 °C)-dried SPH had low turbidity values (by 51.72, 51.47 and 56.07%) compared with the respective SP isolates (p < 0.05), consistent with Psizing results. This reduction is probably linked to the disordering/unfolding of protein-protein interactions following enzymolysis, which induced formation of peptides with small Psizing/aggregates; and this may further elucidate the recorded lower turbidities of the dehydrated SPHs. Low turbidimetric values with lower Psizing of protein are reported [32]. These outcomes were supported by the discussed observations on Psizing and Mweight in this study.

3.2. Solubility

Psolubility is essential in food formulations as it can affect the nutritional value and/or functionality (emulsion, flavor creation, foam, gel, and texture) of protein [33]. To comprehend the differences in the quality of dehydrated SPs and hydrolysates, their solubility under varying pH (2.0–12.0) was assessed (Figure 1). All preparations exhibited a U-shaped Psolubility trend (especially the dried SP isolates) with the lowest (Psolubility) value at pH 4.0, consistent with data from other studies [1,12]. Additionally, all samples displayed maximal Psolubility under extreme alkaline (pH 10.0, 12.0) conditions. This can be ascribed to the increased net charges of samples at pH ≥ 10.0, contributing to the dissociation of protein/hydrolysate aggregates which enhanced the protein-H2O associations and thus improved Psolubility. Freeze-dried SP and SPH had notably higher Psolubility than their respective oven-dried powders (at 40 and 50 °C) under varied pH (6.0–12.0). There was, however, no clear Psolubility trend amongst the convectively-dried powders at various temperatures for both CDSPs and CDSPHs. The low Psolubility of the oven-dried samples is probably linked to the outer skin (highly H2O resistant film) formed at the interface of particle-air due to the unfolding/denaturation of protein through drying [27]. The higher Psolubility of FD powders is credited to lower denaturation of protein during dehydration, which boosted H2O-soluble aggregates. Moreover, Psolubility of the SPHs was a noticeably improved reference to the respective dehydrated SPs under varying pH (p < 0.05). This increase is perhaps, first, associated with their smaller Psizing and reduction in Mweight (as exhibited in Section 3.1 and Section 3.6) as well as the freshly formed amine and carboxyl clusters during enzymolysis, contributing to the formation of hydrogen bridges with water, which improved Psolubility in aqueous suspensions [34]. Secondly, enzymolysis induced the released soluble peptides from in-soluble aggregates. Such observations are comparable to what is found in literature regarding the influence of dehydration techniques [35] and proteolysis [28] on solubility. Based on Psolubility profiles, freeze-dried powders (particularly FDSPH) would be useful for several functional utilizations in food and/or pharmaceutical industry.

3.3. Foaming Attributes

The foaming capacity (Fcapacity) and stability (Fstability) of protein depend mainly on their interaction at interfacial surfaces, which is influenced by molecular flexibility, net charge, conformation and hydrophobicity. Foaming attributes of SPs and SPHs from different dehydration techniques at pH 2.0–12.0 are displayed in Figure 2A,B. Fcapacity of all dehydrated samples (Figure 2A) was noticeably affected by the drying methods, and pH. All preparations showed low Fcapacity at pH 6.0, and thereafter substantially improved (p < 0.05) on both sides of this pH value. Additionally, freeze-dried powders exhibited higher Fcapacity values compared with oven-dried samples (remarkably at pH 2.0–6.0 and 10.0), suggesting that FDSP and FDSPH possessed rapid conformational changes at the interface of gas (air) and solvent with a decrease in surface tension. This increase may be associated with the improved Psolubility of FD powders (as displayed in Figure 1) due to increased positive/negative net charges of these samples in solutions [1]. This may as well be ascribed to higher hydrophobicity and molecular flexibility (lower disulfide bridges) of FDSP and FDSPH, stimulating an increase in Fcapacity. Furthermore, freeze- and convectively-dried SPH had higher Fcapacity than the respective SP isolates under varying pH (p < 0.05). Thus, the enzymolysis altered surface-stabilizing subunits with enhanced interaction at the interface of gas-solvent. In addition, enzyme action potentially reduced molecular weight and increased the flexibility/diffusion speed of SPHs, contributing to an increased absorption rate and stabilization of freshly formed foams at the interface, which (then) improved foam expansion [36].
The Fstability of dehydrated samples were also influenced by drying approaches and pH. Lower Fstability values of convectively-dried samples at various temperatures were observed with reference to the freeze-dried powders. This may be credited to limited alterations in the protein/hydrolysate molecules, resulting in reduced mechanical strength and/or film thickness of gas-solvent interface, which decreased Fstability. Further, freeze- and convectively-dried SPH foam were less stable (p < 0.05) relative to the respective SP powders under acidic (pH 2.0–6.0) conditions. Under alkali pH (8.0–12.0); however, SPHs presented significantly high Fstability compared with SP isolates. These changes in Fstablity under varied pH suggested that the flexibility of the molecules to migrate/orient at interface of gas-solvent was affected by enzymolysis. There are also (in agreement) reports high Fcapacity and Fstability of freeze-dried protein [27,37] and enzyme-hydrolyzed protein [36,38]. Nonetheless, Zeng et al. [35] observed that oven-dried collagen peptide possessed greater foamability and Fstability over a freeze-dried sample. This discrepancy can possibly be attributed to the variations in protein fractions and oven drying conditions.

3.4. Emulsion Traits

The EAindex and ESindex were examined and compared for dehydrated SPs (FDSP, CDSP40, CDSP50), and SPHs (FDSPH, CDSPH40, CDSPH50) under varied pH (Figure 3A,B). Emulsion profiles of dried SPs and SPHs were in agreement with Psolubility pattern. All powders (SPs, SPHs) exhibited lower EAindex and ESindex at isoelectric pH (4.0), whilst the higher values were recorded under alkaline pH. The reason may be that protein/hydrolysate could not migrate rapidly to the oil-H2O interface, and/or the isoelectric pH caused the precipitation of macro-molecules of SPs and SPHs and reduction in net charges of macro-molecules, and consequently impairing the emulsification attributes. Such observations were also noted, with regards to the impact of pH on emulsion attributes of egg white hydrolysate [19]. Furthermore, in comparison with the convectively-dried preparations at various temperatures, notably lower EAindex and ESindex of FD samples (FDSP, FDSPH), at alkaline pH, p < 0.05, were observed. However, no difference (at most pH values— p > 0.05) was observed amongst the oven-dried samples (at 40 °C) and the respective dehydrated powders (at 50 °C). The low emulsifying activity/stability values of lyophilized samples are possibly linked to the reduction in hydrophobic/hydrophilic clusters and the charges of FD samples, impairing oil-in-H2O emulsion. The higher EAindex and ESindex values of CD powders could as well be correlated with the partial denaturation of protein/hydrolysate by heating throughout oven drying (at 40 and 50 °C), inducing the exposure of hydrophilic/hydrophobic amino acids to the peripheral environment which promoted the adsorption capacity at oil-H2O interface.
Results also showed that SPHs possessed low EAindex, but high ESindex values compared with the respective SP isolates (p < 0.05). The low EAindex of SPHs can be associated with the formation of small-sized hydrolysates (peptides) during the enzymolysis. These hydrolysates were diffused and adsorbed rapidly at oil/H2O interface, but thereafter exhibited reduced efficiency in decreasing interfacial tension because they were not unfolded and/or rearranged at the interface, which reduced the EAindex [39]. Moreover, a higher ESindex value of SPHs suggested that these hydrolysates were more stable at oil-H2O interface and were sufficiently flexible to generate good interfacial films, contributing to enhancing emulsion stability [35]. In literature, Elavarasan and Shamasundar [40], and Ghribi et al. [27] indicated that oven drying resulted in greater EAindex and ESindex of protein/hydrolysate than the respective freeze-dried samples, which was in agreement with the noted outcomes in this study. Contrarily, Brishti et al. [16], and Musa et al. [41] found that FD samples had higher emulsion traits (EAindex, ESindex) over convectively-dried powders. This difference can be associated with the variations in protein properties and processing conditions (e.g., drying time and temperature).

3.5. Rheological Analysis

Rheological traits (mainly viscosity) play a crucial role in the texture or mouth-feel of food beverages, as well as in processes such as extrusion and pumping. Lower viscosity for hydrolysate/protein solutions is desirable for piping and pumping, whereas higher viscosity can aid the preparation of sausage/meat analogs and/or used as thickener for soups [14]. For better comprehension of the variations in the behavior of dehydrated SPs and hydrolysates in food system, their rheological attributes following various dehydration techniques were assayed, and the results are presented in Figure 4 and Figure 5 and in Table 2. Viscosity of SPs and SPHs (Figure 4) reduced with increased shear rate (representing shear thinning behavior). The reason can be that, the deformation and disruption of aggregated network of SPs and SPHs suspensions with shear rate (from 2 to 100 s−1), which reduced the viscosities [42]. This is, also, mostly correlated with the orientation of particles and molecules due to increased shear rate to be substantially more than random motion impact generated by the Brown effect.
Further, convection drying of SP and SPH (especially at 50 °C) considerably resulted in a higher viscosity than the respective freeze-dried preparations (at shear rate ranged between 2 and 10 s−1), which was also consistent with the trend of EAindex and ESindex (Figure 3A,B). The high viscosity realized, is mostly due to alterations in protein/hydrolysate structure upon convection drying. Such alterations expose the buried hydrophobic/hydrophilic clusters and sites to surrounding water, and then improves the binding (to H2O molecules) efficiency. Additionally, with increased shear rate (10–100 s−1) all dried samples had very close viscosity values. Results also indicated that SPHs had lower viscosities than the respective SPs preparations at a shear rate of 2–10 s−1. This phenomenon may be credited to the reduction in the hydrodynamic radius of SPHs, which impaired their interaction with H2O [43], thus lowering the viscosity of freeze- and convectively-dried SPH. These data were consistent with the findings of Psizing and Mweight in the current study. Such an outcome agreed with the observation of Lamsal et al. [25] who observed that soy protein possessed higher apparent viscosity relative to its hydrolysates obtained by localized enzymolysis.
Regarding power law constants (consistency coefficient -K, flow behavior index- n), the difference in the mentioned constants was assessed, which was influenced by the dehydration techniques and enzymolysis, applying Equation (7). The power-law equation presented a strong fit for the rheological curves (shear stress against shear rate) of freeze- and convectively-dried SP and SPH (Figure 5). This excellent fit was supported by a high value of R2 (≥0.966) and adj.R2 (≥0.965) and a low value of SSE (≤0.0024) and RMSE (≤0.0099) (Table 2), demonstrating that the existing model accurately described the rheological attributes of all preparations. The maximal K value was recorded for CDSP50 dispersion, followed by CDSP40, FDSP, CDSPH40, CDSPH50, and FDSPH, confirming the order of viscosity (Figure 4) for the respective preparations. Furthermore, SPHs exhibited lower K values (p < 0.05) reference to SP isolates, leading to thinner dispersions. This loss in consistency coefficient upon enzymolysis is mostly attributed to the improved Psolubility (as noted in Figure 1) [25], and the reduction in Psizing (as observed in Table 1) [42] of the respective samples. Furthermore, dehydrated SPs and SPHs dispersions showed a non-Newtonian pseudoplastic behavior (n < 1) with shear thinning. The n values of SP and SPH preparations were in order CDSP50 < CDSP40 < FDSP < CDSPH40 < CDSPH50 < FDSPH (p < 0.05), confirming that dehydration methods and enzymolysis remarkably influenced the flowing behavior of investigated samples. These alterations can possibly be linked to the modification and/or denaturation of SPs/SPHs conformation following the drying process. Finally, freeze-dried samples had lower viscosity and consistency coefficient, but enhanced flow behavior index (of dispersions), indicating that FD samples (observably FDSPH) can find applications in existing food formulations.

3.6. Mweight Analysis

The Mweight distribution of freeze- and convectively-dried SP and SPH (Table 3) showed alterations in SP and SPH compositions with respect to the dehydration techniques used. Relative to oven-dried preparations, the proportion of higher Mweight (≥5000 Da) of the freeze-dried samples, FDSP and FDSPH, was considerably higher. However, small proteins/peptides in FDSP (~9%) and FDSPH (~73%) were noticeably low in ≤3000 Da fractions compared with the respective oven-dried preparations (p < 0.05). These modifications indicated that oven drying resulted in a lower Mweight of SP and SPH and altered their tertiary structure. This observation can particularly be associated with the generation of ice crystal(s) during freeze drying, reflecting the aggregates of particles and recombining into bigger Mweight [30]. Further, freeze and convectively (40 and 50 °C) dried SPH were significantly rich in ≤3000 Da fractions than the SP powders, which improved by 87.57, 87.42, and 80.52%, respectively, demonstrating that lower Mweight of SPs was formed during proteolysis. This may clearly be due to the breakdown of the molecular conformation of SP, leading to the release and/or formation of small-sized hydrolysates/peptides during enzymolysis [13]. A similar effect on Mweight due to proteolysis was also observed [44]. The investigations of Mweight were also in agreement with the data of Psizing (Table 1).

3.7. Correlation Analysis

In this investigation, the effect of dehydration techniques on the physical, functional, and rheological attributes of SP and SPH was examined. To further elucidate the relationship (intrinsically) between these attributes of dehydrated SPs and SPHs, correlational analysis was performed (Table 4). Results of Bindex exhibited negative correlation with Psizing (r = −0.949), Bdensity (r = −0.919), and turbidity (r = −0.945). Inferring from this (high correlation values) is that, Bindex affected the above parameters and vice versa. This demonstrated that the increase in browning index corresponded with reduction in Psizing, bulk density and turbidity. Similarly, a noticed reduction in Psizing was accountable for nearly 98 and 95% of the decrease in Bdensity and turbidity respectively, whereas with a corresponding increase (91%) in ESindex. Such outcomes are typically ascribed to the effect of the convection oven drying and proteolysis, which brought about a decrease in Psizing [13,27], and afterward unfolding of SPs and SPHs structure to display the observed relationship. Moreover, the increase in solubility remarkably contributed to enhancing Fcapacity, Fstability and the flow behavior index (n) of SPs and SPHs, which led to a considerable reduction in the EAindex and consistency coefficient (K). This can be linked to the increased solubility of SP and its hydrolysates as a result of increases in their positive/negative net charges in solutions [1], which enhanced foam formation (foamability) and interfacial films [38] with sufficient strength to stabilize the foams [45]. Reductions in EAindex and K, however, can be explained by the partial alterations and/or denaturation of the protein/hydrolysate conformation, resulting in a decrease in oil/H2O interfacial tension [39] and Psizing of dehydrated SPs and SPHs [42]. Finally, correlational analysis evidenced that dehydration methods altered the physical and rheological traits as well as functionality of SPs and SPHs.

4. Conclusions

In this research, SPs and SPHs were prepared using two drying approaches, i.e., freeze drying and convection oven drying (at 40 and 50 °C). The luminosity, Psizing, Bdensity, and turbidity of freeze-dried SP and SPH were substantially higher, reference to convectively-dried powders; while a lower Bindex was observed (p < 0.05). High solubility, and foaming activity/stability of SP and its hydrolysate, obtained following freeze drying, makes them suitable for food preparations at varied pH values. The high emulsion traits of convectively-dried SPs and SPHs, highpoints the potential for using such in acidic/non-acidic food emulsions. Furthermore, freeze drying resulted in lower viscosity and consistency coefficient, compared with convection drying; whilst greater flow behavior index of dispersions was noticed. There were changes in the rheological and functional characteristics of SPs and SPHs due to the dehydration methods as the correlational analysis confirmed that solubility was negatively interrelated with EAindex and consistency coefficient (K); whereas it showed a positive interrelation with Fcapacity, Fstability, and flow behavior index (n).

Author Contributions

Experimental design, Methodology, Data analysis, Writing—review and editing, M.D. and A.T.P.; Visualization, Funding acquisition, R.S.; Funding acquisition, E.K.; Visualization, A.T.P.; Conceptualization, review/editing, B.K.M.; Funding acquisition, Supervision, Resources, R.H.; Data analysis, Visualization, H.W.; Conceptualization, Validation, M.F. All authors have read and agreed to the published version of the manuscript.

Funding

This research received no external funding.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Available upon request from the corresponding author.

Acknowledgments

This study was supported by the Primary Research and Development Plan (2016YFD0401401); and National Primary Research and Development Plan of Jiangsu Province (BE2016352, BE2016355). Taif University Researchers Supporting Project Number (TURSP-2020/307), Taif University, Taif, Saudi Arabia.

Conflicts of Interest

The authors declared no conflict of interest.

References

  1. Shen, Y.; Tang, X.; Li, Y. Drying methods affect physicochemical and functional properties of quinoa protein isolate. Food Chem. 2021, 339, 127823. [Google Scholar] [CrossRef] [PubMed]
  2. FAOSTAT. Crop Production Statistics; FAO: Rome, Italy, 2021; Available online: https://faostat.fao.org/site/567/default.aspx#ancor (accessed on 19 December 2021).
  3. Dorrell, D.G.; Vick, B.A. Properties and Processing of Oilseed Sunflower. In Sunflower: Technology and Production; Schneiter, A.A., Ed.; Technology and Production, American Society of Agronomy: Madison, WI, USA, 1997; pp. 709–745. [Google Scholar]
  4. González-Pérez, S.; Vereijken, J.M. Sunflower proteins: Overview of their physicochemical, structural and functional properties. J. Sci. Food Agric. 2007, 87, 2173–2191. [Google Scholar] [CrossRef]
  5. Gassmann, B. Preparation and application of vegetable proteins, especially proteins from sunflower seed, for human consumption. An approach. Nahrung 1983, 27, 351–369. [Google Scholar] [CrossRef] [PubMed]
  6. Megías, C.; Pedroche, J.; del Mar Yust, M.; Alaiz, M.; Girón-Calle, J.; Millán, F.; Vioque, J. Purification of angiotensin converting enzyme inhibitory peptides from sunflower protein hydrolysates by reverse-phase chromatography following affinity purification. LWT Food Sci. Technol. 2009, 42, 228–232. [Google Scholar] [CrossRef]
  7. Ren, J.; Zheng, X.Q.; Liu, X.L.; Liu, H. Purification and characterization of antioxidant peptide from sunflower protein hydrolysate. Food Technol. Biotechnol. 2010, 48, 519–523. [Google Scholar]
  8. Taha, F.S.; Mohamed, S.S.; Wagdy, S.M.; Mohamed, G.F. Antioxidant and antimicrobial activities of enzymatic hydrolysis products from sunflower protein isolate. World Appl. Sci. J. 2013, 21, 651–658. [Google Scholar]
  9. Malik, M.A.; Saini, C.S. Improvement of functional properties of sunflower protein isolates near isoelectric point: Application of heat treatment. LWT Food Sci. Technol. 2018, 98, 411–417. [Google Scholar] [CrossRef]
  10. Provansal, M.M.P.; Cuq, J.A.; Cheftel, J.C. Chemical and nutritional modifications of sunflower proteins due to alkaline processing. formation of amino acid cross-links and isomerization of lysine residues. J. Agric. Food Chem. 1975, 23, 938–943. [Google Scholar] [CrossRef] [PubMed]
  11. Minones-Conde, J.; Yust, M.M.; Pedroche, J.; Millán, F.; Rodríguez-Patino, J.M. Effect of enzymatic treatment of extracted sunflower proteins on solubility, amino acid composition, and curface activity. J. Agric. Food Chem. 2005, 53, 8038–8045. [Google Scholar] [CrossRef] [PubMed]
  12. Mintah, B.; He, R.; Dabbour, M.; Xiang, J.; Akomeah, A.; Ma, H. Techno-functional attribute and antioxidative capacity of edible insect protein preparations and hydrolysates thereof: Effect of multiple mode sonochemical action. Ultrason. Sonochem. 2019, 58, 104676. [Google Scholar] [CrossRef]
  13. Mintah, B.; He, R.; Dabbour, M.; Xiang, J.; Hui, J.; Agyekum, A.; Ma, H. Characterization of edible soldier fly protein and hydrolysate altered by multiple-frequency ultrasound: Structural, physical, and functional attributes. Process Biochem. 2020, 95, 157–165. [Google Scholar] [CrossRef]
  14. Alinejad, M.; Motamedzadegan, A.; Rezaei, M.; Regenstein, J.M. The impact of drying method on the functional and antioxidant properties of whitecheek shark (Carcharhinus dussumieri) protein hydrolysates. J. Food Process. Preserv. 2016, 41, e12972. [Google Scholar] [CrossRef] [Green Version]
  15. Deshwal, G.K.; Singh, A.K.; Kumar, D.; Sharma, H. Effect of spray and freeze drying on physico-chemical, functional, moisture sorption and morphological characteristics of camel milk powder. LWT Food Sci. Technol. 2020, 134, 110117. [Google Scholar] [CrossRef]
  16. Brishti, F.H.; Chay, S.Y.; Muhammad, K.; Ismail-Fitry, M.R.; Zarei, M.; Karthikeyan, S.; Saari, N. Effects of drying techniques on the physicochemical, functional, thermal, structural and rheological properties of mung bean (Vigna radiata) protein isolate powder. Food Res. Int. 2020, 138, 109783. [Google Scholar] [CrossRef] [PubMed]
  17. Joshi, M.; Adhikari, B.; Aldred, P.; Panozzo, J.F.; Kasapis, S. Physicochemical and functional properties of lentil protein isolates prepared by different drying methods. Food Chem. 2011, 129, 1513–1522. [Google Scholar] [CrossRef]
  18. Gong, K.; Shi, A.; Liu, H.; Liu, L.; Hu, H.; Adhikari, B.; Wang, Q. Emulsifying properties and structure changes of spray and freeze-dried peanut protein isolate. J. Food Eng. 2016, 170, 33–40. [Google Scholar] [CrossRef]
  19. Chen, C.; Chi, Y.; Xu, W. Comparisons on the functional properties and antioxidant activity of spray-dried and freeze-dried egg white protein hydrolysate. Food Bioprocess Technol. 2012, 5, 2342–2352. [Google Scholar] [CrossRef]
  20. Dabbour, M.; Xiang, J.; Mintah, B.; He, R.; Jiang, H.; Ma, H. Localized enzymolysis and sonochemically modified sunflower protein: Physical, functional and structure attributes. Ultrason. Sonochem. 2020, 63, 104957. [Google Scholar] [CrossRef]
  21. Dabbour, M.; Adingba, E.; Mintah, B.; He, R.; Jiang, H.; Ma, H. Proteolysis kinetics and structural characterization of ultrasonic pretreated sunflower protein. Process Biochem. 2020, 94, 198–206. [Google Scholar] [CrossRef]
  22. Ding, P.; Ling, Y.S. Browning assessment methods and polyphenol oxidase in UV-C irradiated Berangan banana fruit. Int. Food Res. J. 2014, 21, 1667–1674. [Google Scholar]
  23. Lowry, O.; Rosebrough, N.; Farr, L.; Randall, R. Protein measurement with the Folin phenol reagent. J. Biol. Chem. 1951, 193, 265–275. [Google Scholar] [CrossRef]
  24. Liu, Y.; Li, X.; Chen, Z.; Yu, J.; Wang, F.; Wang, J. Characterization of structural and functional properties of fish protein hydrolysates from surimi processing by-products. Food Chem. 2014, 151, 459–465. [Google Scholar] [CrossRef] [PubMed]
  25. Lamsal, B.P.; Jung, S.Ã.; Johnson, L.A. Rheological properties of soy protein hydrolysates obtained from limited enzymatic hydrolysis. LWT—Food Sci. Technol. 2007, 40, 1215–1223. [Google Scholar] [CrossRef]
  26. Wang, Y.; Zhang, Z.; He, R.; Kumah, B.; Dabbour, M.; Qu, W.; Liu, D.; Ma, H. Proteolysis efficiency and structural traits of corn gluten meal: Impact of different frequency modes of a low-power density ultrasound. Food Chem. 2021, 344, 128609. [Google Scholar] [CrossRef] [PubMed]
  27. Ghribi, A.M.; Gafsi, I.M.; Blecker, C.; Danthine, S.; Attia, H.; Besbes, S. Effect of drying methods on physico-chemical and functional properties of chickpea protein concentrates. J. Food Eng. 2015, 165, 179–188. [Google Scholar] [CrossRef]
  28. Ghribi, A.M.; Gafsi, I.M.; Sila, A.; Blecker, C.; Danthine, S.; Attia, H.; Bougatef, A.; Besbes, S. Effects of enzymatic hydrolysis on conformational and functional properties of chickpea protein isolate. Food Chem. 2015, 187, 322–330. [Google Scholar] [CrossRef] [PubMed]
  29. Wasswa, J.; Tang, J.; Gu, X.; Yuan, X. Influence of the extent of enzymatic hydrolysis on the functional properties of protein hydrolysate from grass carp (Ctenopharyngodon idella) skin. Food Chem. 2007, 104, 1698–1704. [Google Scholar] [CrossRef]
  30. Yao, Y.; Zhao, G.; Yan, Y.; Chen, C.; Sun, C.; Zou, X.; Jin, Q.; Wang, X. Effects of freeze drying and spray drying on the microstructure and composition of milk fat globules. RSC Adv. 2016, 6, 2520–2529. [Google Scholar] [CrossRef]
  31. Rudra, S.G.; Sethi, S.; Jha, S.K.; Kumar, R. Physico-chemical and functional properties of cowpea protein isolate as affected by the dehydration technique. Legum. Res. 2016, 39, 370–378. [Google Scholar] [CrossRef] [Green Version]
  32. Zisu, B.; Lee, J.; Chandrapala, J.; Bhaskaracharya, R.; Palmer, M.; Kentish, S.; Ashokkumar, M. Effect of ultrasound on the physical and functional properties of reconstituted whey protein powders. J. Dairy Res. 2011, 78, 226–232. [Google Scholar] [CrossRef]
  33. Thiansilakul, Y.; Benjakul, S.; Shahidi, F. Compositions, functional properties and antioxidative activity of protein hydrolysates prepared from round scad (Decapterus maruadsi). Food Chem. 2007, 103, 1385–1394. [Google Scholar] [CrossRef]
  34. Tavano, O.L. Enzymatic protein hydrolysis using protease: An important tool for food biotechnology. J. Mol. Catal. B Enzym. 2013, 90, 1–11. [Google Scholar] [CrossRef]
  35. Zeng, Q.; Zhang, M.; Adhikari, B.P.; Mujumdar, A.S. Effect of drying processes on the functional properties of collagen peptides produced from chicken skin. Dry. Technol. 2013, 31, 1653–1660. [Google Scholar] [CrossRef]
  36. Bao, Z.; Zhao, Y.; Wang, X.; Chi, Y.-J. Effects of degree of hydrolysis (DH) on the functional properties of egg yolk hydrolysate with alcalase. J. Food Sci. Technol. 2017, 54, 669–678. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  37. Lin, N.; Liu, B.; Liu, Z.; Qi, T. Effects of different drying methods on the structures and functional properties of phosphorylated Antarctic krill protein. J. Food Sci. 2020, 85, 1–10. [Google Scholar] [CrossRef]
  38. Chen, C.; Chi, Y.; Zhao, M.; Xu, W. Influence of degree of hydrolysis on functional properties, antioxidant and ACE Inhibitory activities of egg white protein hydrolysate. Food Sci. Biotechnol. 2012, 21, 27–34. [Google Scholar] [CrossRef]
  39. Gbogouri, G.A.; Linder, M.; Fanni, J.; Parmentier, M. Influence of hydrolysis degree on the functional properties of salmon byproducts hydrolysates. J. Food Sci. 2004, 69, 615–622. [Google Scholar] [CrossRef]
  40. Elavarasan, K.; Shamasundar, B.A. Effect of oven drying and freeze drying on the antioxidant and functional properties of protein hydrolysates derived from freshwater fish (Cirrhinus mrigala) using papain enzyme. J. Food Sci. Technol. 2016, 53, 1303–1311. [Google Scholar] [CrossRef] [Green Version]
  41. Musa, K.H.; Abdullah, A.; Mustapha, W.-A.W. Functional properties of surimi related to drying methods. Malays. Appl. Biol. 2005, 34, 83–87. [Google Scholar]
  42. Yanjun, S.; Jianhang, C.; Shuwen, Z.; Hongjuan, L.; Jing, L.; Lu, L.; Uluko, H.; Yanling, S.; Wenming, C.; Wupeng, G. Effect of power ultrasound pre-treatment on the physical and functional properties of reconstituted milk protein concentrate. J. Food Eng. 2014, 124, 11–18. [Google Scholar] [CrossRef]
  43. Krešić, G.; Jambrak, A.R.; Lelas, V.; Herceg, Z. Influence of innovative technologies on rheological and thermophysical properties of whey proteins and guar gum model systems. Mljekarstvo 2011, 61, 64–78. [Google Scholar]
  44. You, L.; Zhao, M.; Cui, C.; Zhao, H.; Yang, B. Effect of degree of hydrolysis on the antioxidant activity of loach (Misgurnus anguillicaudatus) protein hydrolysates. Innov. Food Sci. Emerg. Technol. 2009, 10, 235–240. [Google Scholar] [CrossRef]
  45. Dabbour, M.; He, R.; Mintah, B.; Xiang, J.; Ma, H. Changes in functionalities, conformational characteristics and antioxidative capacities of sunflower protein by controlled enzymolysis and ultrasonication action. Ultrason. Sonochem. 2019, 58, 104625. [Google Scholar] [CrossRef] [PubMed]
Figure 1. Solubility of freeze- and convectively-dried SP and SPH.
Figure 1. Solubility of freeze- and convectively-dried SP and SPH.
Processes 10 00013 g001
Figure 2. Foaming capacity (A) and stability (B) of freeze- and convectively-dried SP and SPH.
Figure 2. Foaming capacity (A) and stability (B) of freeze- and convectively-dried SP and SPH.
Processes 10 00013 g002
Figure 3. EAindex (A) and ESindex (B) of freeze- and convectively-dried SP and SPH.
Figure 3. EAindex (A) and ESindex (B) of freeze- and convectively-dried SP and SPH.
Processes 10 00013 g003
Figure 4. Apparent viscosity of freeze- and convectively-dried SP and SPH.
Figure 4. Apparent viscosity of freeze- and convectively-dried SP and SPH.
Processes 10 00013 g004
Figure 5. Values of shear stress against shear rate for freeze- and convectively-dried SP and SPH.
Figure 5. Values of shear stress against shear rate for freeze- and convectively-dried SP and SPH.
Processes 10 00013 g005
Table 1. Physical traits of freeze and convectively (oven)-dried SP and SPH.
Table 1. Physical traits of freeze and convectively (oven)-dried SP and SPH.
PropertyFDSPCDSP40CDSP50FDSPHCDSPH40CDSPH50
ColorL *50.82 ± 0.89 a42.36 ± 0.43 b37.25 ± 0.45 c38.27 ± 0.12 c34.12 ± 1.08 d30.48 ± 0.31 e
a *0.38 ± 0.04 d0.50 ± 0.01 c0.56 ± 0.05 c0.79 ± 0.02 b0.86 ± 0.03 b0.94 ± 0.02 a
b *3.84 ± 0.08 d4.10 ± 0.09 d4.59 ± 0.23 cd5.56 ± 0.45 bc6.42 ± 0.31 b7.86 ± 0.72 a
Bindex1.10 ± 0.05 f1.61 ± 0.01 e2.09 ± 0.10 d2.72 ± 0.13 c3.46 ± 0.03 b4.55 ± 0.19 a
Psizing548.61 ± 21.93 a456.11 ± 23.80 b397.86 ± 3.70 c388.55 ± 15.69 c304.17 ± 16.99 d280.71 ± 6.47 d
Bdensity395.27 ± 2.73 a336.24 ± 6.24 b312.92 ± 1.92 c290.86 ± 2.85 d275.45 ± 3.55 e258.74 ± 2.16 f
Ttur0.87 ± 0.015 a0.68 ± 0.003 b0.66 ± 0.005 b0.42 ± 0.008 c0.33 ± 0.011 d0.29 ± 0.003 e
Note: Results (average of three (3) times) in one row with various letter-superscript exhibit significantly different (p < 0.05). FDSP: freeze-dried sunflower protein; CDSP40: convectively-dried sunflower protein at 40 °C; CDSP50: convectively-dried sunflower protein at 50 °C; FDSPH: freeze-dried sunflower protein hydrolysate; CDSPH40: convectively-dried sunflower protein hydrolysates at 40 °C; CDSPH50: convectively-dried sunflower protein at 50 °C.
Table 2. Power law constants and goodness of fit for freeze- and convectively-dried SP and SPH.
Table 2. Power law constants and goodness of fit for freeze- and convectively-dried SP and SPH.
SampleConsistency Coefficient -K (Pa.s)Flow Behavior Index- nGoodness of Fit
R2Adj.R2SSERMSE
FDSP0.0219 ± 0.0002 a0.5355 ± 0.014 c0.9880.9870.00080.0057
CDSP400.0238 ± 0.0010 a0.4220 ± 0.009 d0.9720.9710.00060.0049
CDSP500.0247 ± 0.0008 a0.3874 ± 0.021 d0.9660.9650.00050.0044
FDSPH0.0112 ± 0.0020 c0.7464 ± 0.032 a0.9860.9850.00240.0099
CDSPH400.0172 ± 0.0013 b0.5383 ± 0.018 c0.9890.9890.00040.0043
CDSPH500.0131 ± 0.0015 c0.6445 ± 0.008 b0.9880.9880.00100.0063
Note: Results (average of three (3) times) in one column sharing the various letter-superscript exhibit significantly different (p < 0.05). R2: Coefficient of determination; SSE: Sum of squared errors; RMSE: Root mean square error.
Table 3. Mweight (%) of freeze- and convectively-dried SP and SPH.
Table 3. Mweight (%) of freeze- and convectively-dried SP and SPH.
MW (Da)FDSPCDSP40CDSP50FDSPHCDSPH40CDSPH50
>10,00025.8316.1013.548.273.461.24
10,000–500055.9155.9553.237.445.646.07
5000–30009.2016.7915.7911.392.153.17
3000–10007.548.3111.0217.2218.5314.48
1000–5001.281.842.1518.5023.1923.56
<5000.241.014.2737.1847.0351.48
Table 4. Correlation matrix of the physical, functional and rheological parameters of freeze- and convectively-dried SP and SPH.
Table 4. Correlation matrix of the physical, functional and rheological parameters of freeze- and convectively-dried SP and SPH.
VariablesBindexPsizingBdensityTturPSolubilityFcapacityFstabilityEAindexESindexKn
Bindex1
Psizing−0.949 *1
Bdensity−0.919 *0.980 *1
Ttur−0.945 *0.953 *0.963 *1
PSolubility0.512−0.347−0.390−0.5941
Fcapacity0.741−0.703−0.732−0.869 *0.872 *1
Fstability0.757−0.696−0.745−0.871 *0.895 *0.985 *1
EAindex−0.2730.0790.1050.318−0.929 *−0.692−0.7131
ESindex0.805−0.909 *−0.825 *−0.7970.1330.5530.4800.0851
K−0.7210.5860.6470.781−0.933 *−0.904 *−0.959 *0.789−0.2991
n0.502−0.324−0.399−0.5620.961 *0.7970.864 *−0.901 *0.024−0.955 *1
Note: ‘+’ or ‘−‘ sign represent positive and negative correlation, respectively; * sign represents significant at p < 0.05; K—consistency coefficient; n—flow behavior index; data of PSolubility, Fcapacity, Fstability, EAindex and ESindex were selected at pH 8.0 for correlation analysis.
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Dabbour, M.; Sami, R.; Mintah, B.K.; He, R.; Wahia, H.; Khojah, E.; Petkoska, A.T.; Fikry, M. Effect of Drying Techniques on the Physical, Functional, and Rheological Attributes of Isolated Sunflower Protein and Its Hydrolysate. Processes 2022, 10, 13. https://0-doi-org.brum.beds.ac.uk/10.3390/pr10010013

AMA Style

Dabbour M, Sami R, Mintah BK, He R, Wahia H, Khojah E, Petkoska AT, Fikry M. Effect of Drying Techniques on the Physical, Functional, and Rheological Attributes of Isolated Sunflower Protein and Its Hydrolysate. Processes. 2022; 10(1):13. https://0-doi-org.brum.beds.ac.uk/10.3390/pr10010013

Chicago/Turabian Style

Dabbour, Mokhtar, Rokkaya Sami, Benjamin K. Mintah, Ronghai He, Hafida Wahia, Ebtihal Khojah, Anka Trajkovska Petkoska, and Mohammad Fikry. 2022. "Effect of Drying Techniques on the Physical, Functional, and Rheological Attributes of Isolated Sunflower Protein and Its Hydrolysate" Processes 10, no. 1: 13. https://0-doi-org.brum.beds.ac.uk/10.3390/pr10010013

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