Next Article in Journal
Unified Method for Target and Non-Target Monitoring of Pesticide Residues in Fruits and Fruit Juices by Gas Chromatography-High Resolution Mass Spectrometry
Next Article in Special Issue
Enhancing the Techno-Functionality of Pea Flour by Air Injection-Assisted Extrusion at Different Temperatures and Flour Particle Sizes
Previous Article in Journal
Hot Air Drying of Sipunculus nudus: Effect of Microwave-Assisted Drying on Quality and Aroma
Previous Article in Special Issue
Sustainable Environmental Assessment of Waste-to-Energy Practices: Co-Pyrolysis of Food Waste and Discarded Meal Boxes
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

A Systematical Rheological Study of Maize Kernel

1
School of Public Health, Anhui Medical University, Hefei 230032, China
2
Institute of Food Science and Technology, Chinese Academy of Agricultural Sciences, Key Laboratory of Agro-Products Processing, Ministry of Agriculture and Rural Affairs, Beijing 100193, China
3
State Key Laboratory of Food Science and Technology, School of Food Science and Technology, Jiangnan University, Wuxi 214122, China
*
Author to whom correspondence should be addressed.
Foods 2023, 12(4), 738; https://doi.org/10.3390/foods12040738
Submission received: 15 December 2022 / Revised: 31 January 2023 / Accepted: 2 February 2023 / Published: 8 February 2023

Abstract

:
In this study, the rheological behavior of maize kernel was systematically investigated using a dynamic mechanical analyzer. The loss in toughness caused by drying resulted in a downward shift in the relaxation curve and an upward shift in the creep curve. The long relaxation behavior became obvious when the temperature was above 45 °C, resulting from the weakening of hydrogen bonds with temperature. The maize kernel relaxed more rapidly at high temperatures, caused by a reduction in the cell wall viscosity and polysaccharide tangles. The Deborah numbers were all much smaller than one, suggesting that the Maxwell elements showed viscous behavior. Maize kernel, as a viscoelastic material, showed a dominant viscous property at high temperatures. The decline in β with increasing drying temperature indicated an increase in the width of the relaxation spectrum. A Hookean spring elastic portion made up the majority of the maize kernel creep strain. The order–disorder transformation zone of maize kernel was about 50–60 °C. Due to the complexity of maize kernel, the William–Landel–Ferry constants differed from the universal values; these constants should be ascertained through experiments. Time-temperature superposition was successfully used to describe the rheological behavior. The results show that maize kernel is a thermorheologically simple material. The data acquired in this study can be used for maize processing and storage.

1. Introduction

As an important staple crop, maize is widely grown throughout the world [1]. The water content of maize at harvest ranges from 20% to 28% [2]. For safe storage, maize should be dried to a 12–13% water content [3].
Maize can be dried by many methods, such as hot air drying, vacuum drying, solar drying, and microwave drying. Hot air drying is a procedure in which a product is dried with heated air. This drying method is simple and thus commonly used [4]. However, its efficiency is low [5]. In vacuum drying, the product is dried at sub-atmospheric pressure. Products dried using this method are generally of a better quality [6]. The synergistic combination of vacuum and hot air drying results in high-quality products with a short processing time and low power consumption. Maize can be effectively dried using a combination of the vacuum and hot air processes [7].
Foods are viscoelastic materials [8]. The data on viscoelastic behavior of foods can be used to describe processing operations [9]. The drying conditions, maize varieties, and ripening stage can affect the viscoelastic properties of maize kernel. Stress relaxation is an essential characteristic of a viscoelastic material [10]. In a stress relaxation experiment, a constant deformation is applied to the product and the decline in the stress over time to keep this deformation is measured [11]. Li et al. [12] found that the five-element Maxwell equation could predict the stress relaxation property of sweet potato better than the three-element Maxwell equation. Wang et al. [11] studied the stress relaxation property of rice. They measured the relaxation modulus. Abedi and Takhar [13] investigated the stress relaxation behavior of banana. The ripening stage could affect the rheological behavior pattern of banana in the drying process.
Creep is another essential characteristic of a viscoelastic material [10]. In a creep experiment, a fixed force is applied to the product and the displacement is measured. Li et al. [12] performed the creep-recovery cycle test on sweet potato. They repeated the cycle three times. After the first cycle, no significant differences in the creep elements were observed. Ditudompo et al. [14] found that the Burgers’ equation could describe the creep behavior of an extruded cornstarch well (R2 > 0.92). Ozturk and Takhar [15] studied the viscoelastic behavior of carrot. The creep property of carrot was highly correlated with the moisture content. The storage of maize can be simulated by creep experiments [16]. However, it is difficult to conduct stress relaxation and creep experiments over a wide time range.
Prolonging the time and raising the temperature have an equivalent influence on the viscoelastic behavior of materials. This is known as the theory of time–temperature superposition (TTS) [17]. The applicable standards for TTS are as follows [18]: (a) neighboring curves can be joined to form a smooth curve; (b) all the viscoelastic functions must be superimposed with an identical αT; and (c) the relationship between αT and temperature must satisfy the empirical equations. Polymers that follow TTS are called thermorheologically simple polymers; those that do not follow TTS are called thermorheologically complex polymers [19]. TTS has previously been tested in various polymers [11,20,21,22,23,24,25]. However, the application of TTS to maize kernel has not been reported.
The purposes of the current study were (i) to describe the viscoelastic behavior of maize kernel dried using a combination of vacuum and hot air processes; (ii) to assess the applicability of TTS to maize kernel; and (iii) to characterize the viscoelastic behavior of maize kernel over a wide timescale using TTS.

2. Materials and Methods

2.1. Materials

The variety of maize employed in this study was Nongda 86. Its seed kernels were obtained from freshly harvested maize (Beijing Sinong Seed Co., Ltd., Beijing, China) from Zhangye city, Gansu province, China. The kernels’ initial water content was 27%. Prior to further testing, the kernels were kept in self-sealing bags and refrigerated at 4 °C.

2.2. Drying of Maize Kernels

The maize kernels were dried using the two methods detailed below. The samples’ water content was determined using a GTR800E single-kernel moisture tester (Shizuoka Seiki Co., Ltd., Shizuoka, Japan).

2.2.1. Natural Drying (ND)

The maize kernels were placed on six glass plates (diameter: 20 cm) (Lanyi reagent company, Beijing, China) in one or two (thin) layers and then dried indoors to a water content of 13%. Finally, the samples were kept in self-sealing bags and refrigerated at 4 °C.

2.2.2. Hot Air/Vacuum Drying (HVD)

In a 50 °C oven (Shanghai Yiheng Scientific Instrument Co., Ltd., Shanghai, China), the maize kernels were firstly dried to a water content of 18%. Then, in a −0.1 MPa DZ-3 vacuum drier (Taisite Instrument Co. Ltd., Tianjin, China), the samples were separately dried at 75 °C, 65 °C, 55 °C, 45 °C, and 35 °C to a water content of 13%. Finally, the samples were kept in self-sealing bags and refrigerated at 4 °C.

2.3. Sample Preparation

To ensure the maize kernels had a flat and smooth surface, every sample was carefully polished by hand to prevent any physical failure caused by using abrasive papers. A Fowler PRO-MAX digital vernier caliper (Newton, MA, USA) with a precision of 0.01 mm was employed to determine the width and length of every sample. The dimensions of the samples were about 9 mm × 4 mm × 4 mm (length × width × height). The sample preparation process was based on our previous research [26].

2.4. Stress Relaxation Tests

A TA Instruments Q800 dynamic mechanical analyzer (DMA, New Castle, DE, USA) was employed to measure the maize kernels’ relaxation moduli. The compression mode was employed to carry out the tests. A 0.05 N preload was applied to ensure complete contact between the surface of the sample and the compression plate. The samples were equilibrated at four test temperatures (85 °C, 65 °C, 45 °C, and 25 °C) for 1 min. Then, the tests were conducted at the 0.8% strain for 10 min.
The generalized Maxwell equation (Equation (1)) was employed to describe the maize kernel relaxation modulus:
E ( t ) = m = 1 n E m exp ( t / τ m ) + E 0
where E(t) (MPa) and E0 (MPa) denote the stress relaxation modulus and equilibrium modulus, respectively; τm (s) is the relaxation time; Em (MPa) is the corresponding relaxation modulus; and t (s) is the time.
The Kohlrausch–Williams–Watts (KWW) equation (Equation (2)) was also employed to predict the modulus:
E ( t ) = E 0 exp [ ( t / τ ) β ]
where E0 (MPa) denotes the initial modulus, τ (s) denotes the relaxation time, and β denotes the exponent.

2.5. Creep Tests

The DMA was also employed to determine the kernels’ strains. The compression mode was employed to carry out the tests. A 0.05 N preload was supplied to each kernel. The samples were equilibrated at four test temperatures (60 °C, 50 °C, 40 °C, and 30 °C) for 1 min. Then, the tests were performed at 0.08 MPa stress for 10 min.
Burgers’ equation (Equation (3)) was employed to predict the creep strain.
ε = σ E 1 + σ E 2 [ 1 exp ( t E 2 η 2 ) ] + σ η 1 t
In Equation (3), t (s) represents time; ε (dimensionless) denotes the kernel’s strain; σ (MPa) denotes the applied stress; η1 (MPa s) and η2 (MPa s) denote the Maxwell dashpot viscosity and Kelvin dashpot viscosity, respectively; and E1 (MPa) and E2 (MPa) denote the Maxwell spring modulus and Kelvin spring modulus, respectively. Generally, η2/E2 = τ, where τ denotes the Kelvin element retardation time.
From the differentiation of Equation (3), the creep rate ε’ can be acquired:
ε = d ε d t = σ η 1 + σ η 2 exp ( t E 2 η 2 )
When the time tends to infinity, ε’ will gradually reach a certain value:
ε ( ) = d ε d t | t = = σ η 1
Another equation, called the Findley power law, can also describe the strain of polymers [27]. It has the following form:
ε = a t b + ε 0
Here, ε0 denotes the initial deformation, while a and b denote the material parameters.
A simpler two-parameter power law equation has also been employed to describe the strain of the polymers [28,29,30]:
ε = a t b

2.6. Time–Temperature Superposition

The strain and relaxation modulus of the maize kernels over a wide time range were described using TTS. The master curves were produced by transferring the short-period curves horizontally [31]. The William–Landel–Ferry (WLF) [19] (Equation (8)) or Arrhenius equation [32] (Equation (9)) can be employed to describe the shift factor.
log α T = C 1 T 0 C 1 T T T 0 + C 2
In Equation (8), αT, T0 (K), and T (K) represent the shift factor, reference temperature, and test temperature, respectively, while C1 and C2 (K) are the WLF parameters.
log α T = E a R ( 1 T 1 T 0 )
In Equation (9), R (J/K·mol) and Ea (kJ/mol) denote the gas law constant and activation energy, respectively, while T0 (K) and T (K) denote the reference and experimental temperatures, respectively.
As suggested by Ferry [18], the Arrhenius equation and WLF equation were both employed to describe αT. C1, C2, Ea, and the coefficient of determination (R2) were determined from the experimental data using regression analysis. Temperatures of 25 °C and 30 °C were chosen as the reference temperatures for stress relaxation and creep TTS, respectively.

2.7. Statistical Analysis

Three replicates were conducted for every sample. Universal Analysis 2000 software (version 4.3A, TA Instruments, New Castle, DE, USA) was employed to acquire the DMA data. SPSS 17.0 (SPSS Inc., Chicago, IL, USA) was employed to conduct the statistical analysis.

3. Results and Discussion

3.1. Analysis of Stress Relaxation Behavior

The stress relaxation data for maize kernels at different temperatures are plotted in Figure 1. Three stages make up the curve. The first stage shows an immediate elastic response. The timescale of this region is nearly zero. The second stage demonstrates a marked stress relaxation process. The timescale is from about 0 to 20 s. In the third stage, the modulus approaches a constant. The maize kernel is viscoelastic [8]. When strain is applied to the kernel, the stress increases immediately due to the elastic portion of the kernel and then decreases gradually due to the viscous portion of the kernel. This trend is identical to the findings reported in a previous study [33]. It can be seen from Figure 1 that the kernels dried by ND had the highest initial relaxation modulus. This is because the kernels dried by ND had the highest toughness. For the same sample, when the temperature was increased, the initial modulus declined markedly. This was caused by an increase in the motion of the polymer chains, resulting in a lowering of the threshold for specific strain conditions [34].
Table 1 presents the elements of the generalized Maxwell equation. As listed in Table 1, the R2 was around 0.94–0.98, indicating that the stress relaxation modulus was described well by the generalized Maxwell equation. A similar finding was reported in a study on oat grains [35]. Table 1 also shows that E1 decreased as the temperature increased, suggesting that higher temperatures reduced the maize kernel’s deformation resistance [36]. Hence, the stress required to achieve the given strain decreased. At the same temperature, τ fluctuated with drying conditions and did not display a decreasing or increasing tendency, suggesting that the relaxation time was not related to the drying method. In general, τ decreased as the temperature increased. That is, kernels relaxed more rapidly at higher temperatures. The reason for this is as follows: when a maize kernel becomes soft due to high temperatures, the cell wall viscosity and polysaccharide tangles will be reduced, leading to faster stress dissipation [36]. By using the expression below, the Deborah number (De) can be obtained:
D e = τ / t
Here, t denotes the observation time. De can provide useful information on materials’ viscoelastic properties: De ≈ 1 denotes a viscoelastic behavior; De << 1 denotes a viscous fluid; and De >> 1 denotes an elastic solid [37]. The De values (0.094–0.385) calculated in this study were much smaller than one; this suggests that the Maxwell element showed viscous behavior.
Table 1 also shows that when the temperature increased, E0 decreased. This suggests that maize kernel is mainly a viscous material at high temperatures. The maize kernel’s viscous portion begins to decrease as the temperature decreases. According to the theory of free volume, molecular dynamics can account for the influence of the viscous portion. When the temperature is low, there is not enough free volume to allow for the motion of molecular chains; thus, the molecules’ movement is fixed. This makes the kernel glassy. When the temperature is high, the free volume becomes larger and the molecular chains can move. This makes the kernel rubbery [11]. In addition, the E0 for ND was higher than the E0 for HVD (at the same temperature), except in the case of HVD 55 °C at 45 °C and HVD 75 °C at 65 °C. This suggests that the kernel’s stiffness changed. This is owing to the changes in the kernel’s cellular structure caused by hot air/vacuum drying, giving rise to a decline in the stiffness. The change in E0 may have resulted from the breakdown of cell membranes caused by HVD [38], leading to the collapse of the cells [33,39].

3.2. Analysis of Creep Behavior

Figure 2 displays the creep curves for maize kernels at various temperatures. The curves demonstrate classic creep features and exhibit a similar tendency over time. When stress is applied to the maize kernels, the strain increases instantly because of elasticity. In this portion, the changes in the kernel’s structure are reversible [40]. Then, the strain continues to increase at a decreasing rate due to viscosity. This portion of the curve contributes to the maize kernel’s viscoelastic property [15]. The last portion is a straight line, representing the kernel’s viscous flow deformation. In this compression stage, the kernel’s tissues are permanently destroyed, and the changes can only be partially reversed [41]. This tendency is consistent with the creep strain of barley kernels reported in a previous study [42]. In this study, the creep strain exhibited a strong temperature dependence. When the temperature increased, the strain also increased. A similar finding was reported in a previous study [43]. Therefore, the solid-like characteristics of maize kernel increase as the temperature decreases [14]. It can also be seen from Figure 2 that the ND kernels had the lowest creep strain. This is because the kernels dried using the ND method were more compact and robust.
Table 2 displays the values of the four elements (E1, E2, η1, and η2) mentioned in Section 2.5. The data in this table suggest that Burgers’ equation can predict the creep property well (R2 > 0.98). E1, E2, η1, and η2 all decreased as the temperature increased, suggesting that increasing temperature increases the deformation of maize kernel. E1 is correlated with the elasticity of a semi-crystalline polymer’s crystalline regions. In contrast to the amorphous zones, the crystalline regions are exposed to instant stress because of their relatively high hardness. When the stress is eliminated, E1 can be resumed instantly [44]. E2 is correlated with the rigidity of a material [30]. When the temperature increases, the instant and viscoelastic strain increase. As a result, E1 and E2 decrease. η1 is correlated with the damage from the crystalline regions and the irrevocable strain from the amorphous zones. η2 corresponds to the viscosity of the semi-crystalline polymer’s amorphous zones [45]. When the temperature increases, the values of η1 and η2 decline. This indicates that the molecular chains become more active [29]. The ND elements exhibited the greatest declines (particularly in η1, which depends heavily on temperature and is correlated with the rate of long-period deformation), suggesting that the viscoelastic property of kernel dried by ND was more sensitive to temperature.
It can also be seen from Table 2 that E2 was larger than E1. In comparison to the viscous and retarded elastic strain, the instant strain was larger. Thus, the Hookean spring elastic portion mainly makes up maize kernel’s creep strain [14]. Table 2 also presents ε’ (∞) and τ for the maize kernels. When the temperature was increased, ε’ (∞) increased monotonically. At 60 °C, the kernels generally showed a longer retardation time (τ), indicating that the samples’ viscoelastic property remained for longer [41].
Overall, the kernel dried by ND had a higher stiffness. The viscoelastic property of the kernel dried by ND was more sensitive to temperature. The relaxation time of the kernel was not related to the drying method.

3.3. Applicability of Time-Temperature Superposition

Figure 1 and Figure 2 were replotted in logarithmic form (left-hand graphs in Figure 3 and Figure 4) to employ TTS. It can be seen that for stress relaxation, the impact of temperature was obvious at 45 °C and became more significant at 65 °C; for creep, the impact was obvious at 50 °C and became significant at 60 °C. The temperature range of 45–65 °C and 50–60 °C denoted the order–disorder transformation zones of maize kernel [46]. These are close to the gelatinization temperature of maize starch (64–72 °C) [47]. Furthermore, from the graphs on the left in Figure 3, it can be seen that the long relaxation behavior became obvious when the temperature was above 45 °C. Specific molecular interactions in maize kernel and temperature have an impact on when the long relaxation process occurs [21]. In general, in this study, long relaxation occurred when the time was approximately 200 s at 65 °C, and when it was 160 s at 85 °C. This is because the hydrogen bonds were weakened when the temperature was raised [48].
By shifting the stress relaxation (left-hand graphs in Figure 3) and creep (left-hand graphs in Figure 4) curves horizontally, the master curves were acquired (right-hand graphs in Figure 3 and Figure 4). As noted earlier, the applicable standards for TTS are as follows: (a) neighboring curves can be joined to form a smooth curve; (b) all the viscoelastic functions must be superimposed with an identical αT; and (c) the relationship between αT and the temperature must satisfy the empirical equations. It can be seen from the graphs in Figure 3 and Figure 4 (right) that the relaxation and creep curves exhibited superposition for all the samples, although the master curves showed some tails and irregularities (Figure 3(A2–E2) and Figure 4(B2,C2)). The absence of superposition is correlated with the relaxation mechanism, which depends on temperature and takes place inside the maize kernel [49]. This is parallel to the findings reported in a previous study [50]. In that study, the curves obtained at higher temperatures could not be overlapped smoothly with only horizontal shifting. Nevertheless, by shifting the obtained curves both vertically and horizontally, smooth master curves were acquired. In general, the relaxation modulus and creep strain master curves were considered to be continuous and smooth in this study. Some curves (Figure 3(F2) and Figure 4(A2,D2–F2)) demonstrated an excellent fit for the TTS. These findings suggest that the maize kernel is a thermorheologically simple material [51]. The applicability of TTS has also been evaluated for other food materials [46,49,52,53,54], and soybean [55] was reported as a thermorheologically simple material.
It can be seen from Figure 5 and Figure 6 that αT reduced monotonically with the temperature. This is similar to the finding of Meza et al. [56]. Furthermore, αT versus the temperature data could be fitted well by both the Arrhenius equation and the WLF equation (R2 > 0.857). However, the correlation between αT and the temperature was not very high for some samples. This is because the temperatures were too close, and no remarkable viscoelastic variations were detected, similar to the finding reported by Ahmed [49]. In his study, no noticeable rheological variations were found in guar gum dispersions due to the closeness of the temperatures.
Table 3 and Table 4 display the Arrhenius activation energy (Ea) and WLF constants. The WLF constants, C1 and C2, are given by [57]:
C 1 = B 2.303 f
C 2 = f α
Here, B can be taken equal to unity, f denotes the fractional free volume at the reference temperature, and α denotes the thermal expansion coefficient. Although the WLF constants fluctuated with the drying conditions, the WLF constants for ND and HVD 35 °C were the highest, suggesting that the fractional free volume and thermal expansion coefficient were the smallest.
It can be seen that the R2 of the WLF equation was higher than that of the Arrhenius equation. A similar finding was reported in a previous study [58]. Despite this, it should be noted that there are two coefficients in the WLF equation, while there is only one coefficient in the Arrhenius equation. Hence, the WLF equation is more adaptable. We can also observe from Table 3 and Table 4 that the C1 and C2 values ranged from 2.226 to 32.020 and 8.346 to 257.500 K, respectively. The values of C1 were slightly different from the universal value (C1 = 17.44), while the values of C2 were considerably different (C2 = 51.6 K) [57]. This is due to the complexity of food polymers (including different charges on the surfaces of polymers, various polysaccharides, 20 amino acids of protein, large polydispersities) [59] and is similar to the finding of Altay and Gunasekaran [60]. In their research, the values of C1 and C2 for a gelatin–xanthan gum system ranged from 4.52 to 22.68 and 58.97 to 204.97 K, respectively. Peleg [61] stated that employing universal values in food polymers ought to be approached cautiously. The values of C1 and C2 for food materials should be ascertained through experiments due to their substance-specific characteristics [59].
In this study, the Ea values for stress relaxation ranged from 90.86 to 134.8 kJ/mol. These values are very close to the Ea value (94.3 kJ/mol) of rice kernels (13.8% moisture content) for stress relaxation [11]. This is likely because maize kernels and rice kernels are both mainly composed of starch. The values for creep ranged from 111.035 to 292.645 kJ/mol, which are higher than the values (62.960–166.539 kJ/mol) for starch films [62,63]. Furthermore, the values for creep were significantly higher than those for stress relaxation. The Ea for relaxation and creep represents the energy needed for the initiation of the molecular movements, leading to relaxation and creep [64]. From Table 3 and Table 4, the values for the kernels dried by HVD at 75 °C and HVD at 55 °C were the highest, suggesting that more energy was required, while the values for the kernels dried by HVD at 45 °C were the lowest.
Overall, TTS was successfully applied to the maize kernel. TTS extended the timescale of the relaxation modulus and creep strain from 2 to 5–6 log periods (Figure 3) and 2 to 4–7 log periods (Figure 4), respectively. Performing a DMA experiment for such a long time is extremely challenging. The application of TTS to the rice kernel was reported in a previous study [11]. In that study, the timescale of the relaxation modulus for the rice kernel was extended from 2 to 6 log periods.

3.4. Long-Period Stress Relaxation and Creep Response

Figure 7 shows the TTS predicted relaxation moduli for maize kernels dried under different conditions at 25 °C. As shown in this figure, the resulting predicted curves provide an accelerated stress relaxation characterization up to 1600 h. Constant strain-rate tests performed for yellow-dent maize kernels were reported in a previous study [9]. In that study, the timescale of the relaxation modulus was extended to 106 s; however, the modulus was higher than that in the current study. This may have resulted from differences in the maize varieties and experimental conditions.
Figure 8 shows the TTS predicted creep strains for maize kernels dried under different conditions at 30 °C. The timescale of the creep strain was extended to 5000 h, and the trends of the curves were similar in the longer time frame. The timescale of the creep strain for the starch film was extended to 107 s in another study [63].
Table 5 lists the equation parameters for long-period stress relaxation. We can observe from this table that both the Maxwell equation and KWW equation fitted the data well. The R2 of the Maxwell equation was slightly higher than that of the KWW equation. Similar studies have reported fitting the generalized Maxwell equation to the relaxation modulus data of corn kernels [9] and rice kernels [11]. In this study, the values of E1, E2, and E3 were close to each other. This suggests that the maize kernel’s inner structure changed slightly during the long-period relaxation [36]. Furthermore, for the entire relaxation time, the Deborah numbers (De1 = 0.00034–0.14918; De2 = 0.00004–0.21343; De3 = 0.00011–0.06139) were much smaller than one. This suggests that the first, second, and third Maxwell elements all show a viscous response.
The value of the exponent β declined from 0.194 to 0.145 when the drying temperature was increased to 75 °C, except in the case of HVD at 45 °C. In accordance with Ngai et al.’s coupling theory [65], the medium and the relaxing species become more tightly coupled when β declines. This is correlated with overall declines in the molecular motion and increases in the width of the relaxation spectrum. The decline in β results from the increase in the density of the polymer with increasing drying temperature. When the space among the molecules is reduced, the motion of molecular chains is more limited, increasing the width of the relaxation spectrum [66].
Table 6 lists the equation parameters for long-period creep. We can observe from this table that the creep data were described well by Burgers’ equation for all the maize kernels. Nonetheless, long-term experiments have considerably different elements than short-term experiments, caused by fitting a higher rate of creep to the strain in the short-term experiments [29].
The Findley power law equation can better represent the TTS predicted strain than Burgers’ equation. The two-parameter power law equation also describes the creep data well. When the parameter ε0 was removed, a and b became more consistent. This is similar to the finding from a previous study [19].

4. Conclusions

With regard to stress relaxation, creep, and TTS, the rheological properties of maize kernels dried using a combination of vacuum and hot air processes were investigated. The kernel dried by ND had a higher stiffness. The viscoelastic property of kernel dried by ND was more sensitive to temperature. The relaxation time of the kernel was not related to the drying method. The long relaxation behavior became obvious when the temperature was above 45 °C, as a result of the weakening of hydrogen bonds with temperature. The Des were all much smaller than one, indicating that the Maxwell elements showed viscous behavior. The maize kernels exhibited dominant viscous behavior at high temperatures. The decline in β with increasing drying temperature indicates an increase in the width of the relaxation spectrum. The order–disorder transformation zone of maize kernel was about 50–60 °C. Finally, TTS was successfully applied to predict the long-period relaxation modulus (up to 1600 h) and creep strain (up to 5000 h). These results show that maize kernel is a thermorheologically simple material. Further structural analyses (such as atomic force microscope and nuclear magnetic resonance analyses) are needed to clarify the molecular mechanism of the HVD kernels. The data acquired in this study can be used for maize processing and storage.

Author Contributions

Conceptualization, S.S. and A.S.; methodology, S.S. and J.X.; software, S.S.; validation, S.S.; formal analysis, S.S.; writing—original draft preparation, S.S.; writing—review and editing, A.S. and J.X.; visualization, A.S. and J.X.; project administration, S.S.; funding acquisition, S.S. All authors have read and agreed to the published version of the manuscript.

Funding

This study was supported by Grants for Scientific Research of BSKY (XJ201623) from Anhui Medical University.

Data Availability Statement

Data are contained within the article.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Jiang, J.; Xing, F.; Wang, C.; Zeng, X.; Zou, Q. Investigation and development of maize fused network analysis with multi-omics. Plant Physiol. Biochem. 2019, 141, 380–387. [Google Scholar] [CrossRef] [PubMed]
  2. Li, H.; Wang, Y.; Xue, J.; Xie, R.; Wang, K.; Zhao, R.; Liu, W.; Ming, B.; Hou, P.; Zhang, Z.; et al. Allocation of maize varieties according to temperature for use in mechanical kernel harvesting in Ningxia, China. Int. J. Agric. Biol. Eng. 2021, 14, 20–28. [Google Scholar] [CrossRef]
  3. Bradford, K.J.; Dahal, P.; Van Asbrouck, J.; Kunusoth, K.; Bello, P.; Thompson, J.; Wu, F. The dry chain: Reducing postharvest losses and improving food safety in humid climates. Trends Food Sci. Technol. 2018, 71, 84–93. [Google Scholar] [CrossRef]
  4. Bazyma, L.A.; Guskov, V.P.; Basteev, A.V.; Lyashenko, A.M.; Lyakhno, V.; Kutovoy, V.A. The investigation of low temperature vacuum drying processes of agricultural materials. J. Food Eng. 2006, 74, 410–415. [Google Scholar] [CrossRef]
  5. Maskan, M. Microwave/air and microwave finish drying of banana. J. Food Eng. 2000, 44, 71–78. [Google Scholar] [CrossRef]
  6. Arévalo-Pinedo, A.; Murr, F.E.X. Kinetics of vacuum drying of pumpkin (Cucurbita maxima): Modeling with shrinkage. J. Food Eng. 2006, 76, 562–567. [Google Scholar] [CrossRef]
  7. Onwude, D.I.; Hashim, N.; Chen, G. Recent advances of novel thermal combined hot air drying of agricultural crops. Trends Food Sci. Technol. 2016, 57, 132–145. [Google Scholar] [CrossRef]
  8. Essex, E. Objective measurements for texture in foods. J. Texture Stud. 1969, 1, 19–37. [Google Scholar] [CrossRef]
  9. Waananen, K.M.; Okos, M.R. Stress-relaxation properties of yellow-dent corn kernels under uniaxial loading. Trans. ASAE 1992, 35, 1249–1258. [Google Scholar] [CrossRef]
  10. Ravanbod, R.; Torkaman, G.; Esteki, A. Biotribological and biomechanical changes after experimental haemarthrosis in the rabbit knee. Haemophilia 2011, 17, 124–133. [Google Scholar] [CrossRef] [PubMed]
  11. Wang, P.; Wang, L.; Li, D.; Huang, Z.; Adhikari, B.; Chen, X.D. The stress-relaxation behavior of rice as a function of time, moisture and temperature. Int. J. Food Eng. 2017, 13, 20160162. [Google Scholar] [CrossRef]
  12. Li, Q.; Li, D.; Wang, L.; Özkan, N.; Mao, Z. Dynamic viscoelastic properties of sweet potato studied by dynamic mechanical analyzer. Carbohyd. Polym. 2010, 79, 520–525. [Google Scholar] [CrossRef]
  13. Abedi, F.M.; Takhar, P.S. Stress relaxation properties of bananas during drying. J. Texture Stud. 2022, 53, 146–156. [Google Scholar] [CrossRef] [PubMed]
  14. Ditudompo, S.; Takhar, P.S.; Ganjyal, G.M.; Hanna, M.A. The effect of temperature and moisture on the mechanical properties of extruded cornstarch. J. Texture Stud. 2013, 44, 225–237. [Google Scholar] [CrossRef]
  15. Ozturk, O.K.; Takhar, P.S. Physical and viscoelastic properties of carrots during drying. J. Texture Stud. 2020, 51, 532–541. [Google Scholar] [CrossRef]
  16. Shahgholi, G.; Latifi, M.; Jahanbakhshi, A. Potato creep analysis during storage using experimental measurement and finite element method (FEM). J. Food Process Eng. 2020, 43, e13522. [Google Scholar] [CrossRef]
  17. Lin, C.; He, G.; Liu, J.; Pan, L.; Liu, S.; Li, J.; Guo, S. Construction and non-linear viscoelastic properties of nano-structure polymer bonded explosives filled with graphene. Compos. Sci. Technol. 2018, 160, 152–160. [Google Scholar] [CrossRef]
  18. Ferry, J.D. Viscoelastic Properties of Polymers, 3rd ed.; John Wiley & Sons: New York, NY, USA, 1980; p. 273. [Google Scholar]
  19. Xu, Y. Creep behavior of natural fiber reinforced polymer composites. Ph.D. Thesis, Louisiana State University, Baton Rouge, LA, USA, 2009. [Google Scholar]
  20. Vaidyanathan, T.K.; Vaidyanathan, J. Validity of predictive models of stress relaxation in selected dental polymers. Dent. Mater. 2015, 31, 799–806. [Google Scholar] [CrossRef]
  21. Kontogiorgos, V. Linear viscoelasticity of gluten: Decoupling of relaxation mechanisms. J. Cereal Sci. 2017, 75, 286–295. [Google Scholar] [CrossRef]
  22. Koutsomichalis, A.; Kalampoukas, T.; Mouzakis, D.E. Mechanical testing and modeling of the time–temperature superposition response in hybrid fiber reinforced composites. Polymers 2021, 13, 1178. [Google Scholar] [CrossRef]
  23. Amjadi, M.; Fatemi, A. Creep behavior and modeling of high-density polyethylene (HDPE). Polym. Test. 2021, 94, 107031. [Google Scholar] [CrossRef]
  24. Ghosh, S.K.; Rajesh, P.; Srikavya, B.; Rathore, D.K.; Prusty, R.K.; Ray, B.C. Creep behaviour prediction of multi-layer graphene embedded glass fiber/epoxy composites using time-temperature superposition principle. Compos. Part A Appl. Sci. Manuf. 2018, 107, 507–518. [Google Scholar] [CrossRef]
  25. Xue, C.; Gao, H.; Hu, Y.; Hu, G. Experimental test and curve fitting of creep recovery characteristics of modified graphene oxide natural rubber and its relationship with temperature. Polym. Test. 2020, 87, 106509. [Google Scholar] [CrossRef]
  26. Sheng, S.; Wang, L.; Li, D.; Mao, Z.; Adhikari, B. Viscoelastic behavior of maize kernel studied by dynamic mechanical analyzer. Carbohyd. Polym. 2014, 112, 350–358. [Google Scholar] [CrossRef]
  27. Wang, Z.D.; Zhao, X.X. Modeling and characterization of viscoelasticity of PI/SiO2 nanocomposite films under constant and fatigue loading. Mater. Sci. Eng. A 2008, 486, 517–527. [Google Scholar] [CrossRef]
  28. Tajvidi, M.; Falk, R.H.; Hermanson, J.C. Time–temperature superposition principle applied to a kenaf-fiber/high-density polyethylene composite. J. Appl. Polym. Sci. 2005, 97, 1995–2004. [Google Scholar] [CrossRef]
  29. Xu, Y.; Wu, Q.; Lei, Y.; Yao, F. Creep behavior of bagasse fiber reinforced polymer composites. Bioresour. Technol. 2010, 101, 3280–3286. [Google Scholar] [CrossRef]
  30. Jabbar, A.; Militký, J.; Madhukar Kale, B.; Rwawiire, S.; Nawab, Y.; Baheti, V. Modeling and analysis of the creep behavior of jute/green epoxy composites incorporated with chemically treated pulverized nano/micro jute fibers. Ind. Crop. Prod. 2016, 84, 230–240. [Google Scholar] [CrossRef]
  31. Zhang, Y.; Sun, Z.; Li, Y.; Huang, P.; Chen, Q.; Fu, S. Tensile creep behavior of short-carbon-fiber reinforced polyetherimide composites. Compos. Part B Eng. 2021, 212, 108717. [Google Scholar] [CrossRef]
  32. Rajan, S.; Sutton, M.A.; Wehbe, R.; Gurdal, Z.; Kidane, A.; Emri, I. Characterization of viscoelastic bending stiffness of uncured carbon-epoxy prepreg slit tape. Compos. Struct. 2021, 275, 114295. [Google Scholar] [CrossRef]
  33. Liu, Y.; Wang, Y.; Lv, W.; Li, D.; Wang, L. Freeze-thaw and ultrasound pretreatment before microwave combined drying affects drying kinetics, cell structure and quality parameters of Platycodon grandiflorum. Ind. Crop. Prod. 2021, 164, 113391. [Google Scholar] [CrossRef]
  34. Qiao, J.; Wu, Y.; Li, L. Effect of filler content on the stress relaxation behavior of fly ash/polyurea composites. Polym. Test. 2020, 81, 106168. [Google Scholar] [CrossRef]
  35. Zhao, N.; Li, B.; Zhu, Y.; Li, D.; Wang, L. Viscoelastic analysis of oat grain within linear viscoelastic region by using dynamic mechanical analyzer. Int. J. Food Eng. 2020, 16, 20180350. [Google Scholar] [CrossRef]
  36. Ozturk, O.K.; Takhar, P.S. Stress relaxation behavior of oat flakes. J. Cereal Sci. 2017, 77, 84–89. [Google Scholar] [CrossRef]
  37. Sozer, N.; Dalgic, A.C. Modelling of rheological characteristics of various spaghetti types. Eur. Food Res. Technol. 2007, 225, 183–190. [Google Scholar] [CrossRef]
  38. Khan, M.; Nagy, S.A.; Karim, M.A. Transport of cellular water during drying: An understanding of cell rupturing mechanism in apple tissue. Food Res. Int. 2018, 105, 772–781. [Google Scholar] [CrossRef]
  39. Wang, J.; Mujumdar, A.S.; Deng, L.Z.; Gao, Z.J.; Xiao, H.W.; Raghavan, G. High-humidity hot air impingement blanching alters texture, cell-wall polysaccharides, water status and distribution of seedless grape. Carbohydr. Polym 2018, 194, 9–17. [Google Scholar] [CrossRef]
  40. Ozturk, O.K.; Singh Takhar, P. Selected physical and viscoelastic properties of strawberries as a function of heated-air drying conditions. Dry. Technol. 2019, 37, 1833–1843. [Google Scholar] [CrossRef]
  41. Sandhu, J.S.; Takhar, P.S. Effect of frying parameters on mechanical properties and microstructure of potato disks. J. Texture Stud. 2015, 46, 385–397. [Google Scholar] [CrossRef]
  42. Zhu, Y.; Fu, N.; Li, D.; Wang, L.; Chen, X.D. Physical and viscoelastic properties of different moisture content highland barley kernels. Int. J. Food Eng. 2017, 13, 20170186. [Google Scholar] [CrossRef]
  43. Ornaghi, H.L.; Monticeli, F.M.; Neves, R.M.; Zattera, A.J.; Cioffi, M.O.H.; Voorwald, H.J.C. Effect of stacking sequence and porosity on creep behavior of glass/epoxy and carbon/epoxy hybrid laminate composites. Compos. Commun. 2020, 19, 210–219. [Google Scholar] [CrossRef]
  44. Hao, A.; Chen, Y.; Chen, J.Y. Creep and recovery behavior of kenaf/polypropylene nonwoven composites. J. Appl. Polym. Sci. 2014, 131, 8864–8874. [Google Scholar] [CrossRef]
  45. Militký, J.; Jabbar, A. Comparative evaluation of fiber treatments on the creep behavior of jute/green epoxy composites. Compos. Part B Eng. 2015, 80, 361–368. [Google Scholar] [CrossRef]
  46. Ahmed, J. Optimization of high-pressure-assisted xanthan gum dispersions for the maximization of rheological moduli: Application of time-pressure/ temperature superposition principle. Food Hydrocolloid. 2022, 122, 107080. [Google Scholar] [CrossRef]
  47. Mateyawa, S.; Xie, D.F.; Truss, R.W.; Halley, P.J.; Nicholson, T.M.; Shamshina, J.L.; Rogers, R.D.; Boehm, M.W.; Mcnally, T. Effect of the ionic liquid 1-ethyl-3-methylimidazolium acetate on the phase transition of starch: Dissolution or gelatinization? Carbohyd. Polym. 2013, 94, 520–530. [Google Scholar] [CrossRef] [PubMed]
  48. Kontogiorgos, V.; Shah, P.; Bills, P. Influence of supramolecular forces on the linear viscoelasticity of gluten. Rheol. Acta 2016, 55, 187–195. [Google Scholar] [CrossRef]
  49. Ahmed, J. Effect of pressure, concentration and temperature on the oscillatory rheology of guar gum dispersions: Response surface methodology approach. Food Hydrocolloid. 2021, 113, 106554. [Google Scholar] [CrossRef]
  50. Nakada, M.; Miyano, Y.; Cai, H.; Kasamori, M. Prediction of long-term viscoelastic behavior of amorphous resin based on the time-temperature superposition principle. Mech. Time Depend. Mat. 2011, 15, 309–316. [Google Scholar] [CrossRef]
  51. Yusoff, N.I.M.; Jakarni, F.M.; Nguyen, V.H.; Hainin, M.R.; Airey, G.D. Modelling the rheological properties of bituminous binders using mathematical equations. Constr. Build. Mater. 2013, 40, 174–188. [Google Scholar] [CrossRef]
  52. Singh, A.P.; Lakes, R.S.; Gunasekaran, S. Viscoelastic characterization of selected foods over an extended frequency range. Rheol. Acta 2006, 46, 131–142. [Google Scholar] [CrossRef]
  53. Sun, S.; Song, Y.; Zheng, Q. Morphology and mechanical properties of thermo-molded bioplastics based on glycerol-plasticized wheat gliadins. J. Cereal Sci. 2008, 48, 613–618. [Google Scholar] [CrossRef]
  54. Ahmed, J. Applicability of time–temperature superposition principle: Dynamic rheology of mung bean starch blended with sodium chloride and sucrose—Part 2. J. Food Eng. 2012, 109, 329–335. [Google Scholar] [CrossRef]
  55. Herum, F.L.; Mensah, J.K.; Barre, H.J.; Majidzadeh, K. Viscoelastic behavior of soybeans due to temperature and moisture content. Trans. ASAE 1979, 22, 1219–1224. [Google Scholar] [CrossRef]
  56. Meza, B.E.; Verdini, R.A.; Rubiolo, A.C. Temperature dependency of linear viscoelastic properties of a commercial low-fat soft cheese after frozen storage. J. Food Eng. 2012, 109, 475–481. [Google Scholar] [CrossRef]
  57. Altay, F.; Gunasekaran, S. Rheological evaluation of gelatin–xanthan gum system with high levels of co-solutes in the rubber-to-glass transition region. Food Hydrocolloid. 2012, 28, 141–150. [Google Scholar] [CrossRef]
  58. Contreras Mateus, M.D.; López López, M.T.; Ariza-León, E.; Chaves Guerrero, A. Rheological implications of the inclusion of ferrofluids and the presence of uniform magnetic field on heavy and extra-heavy crude oils. Fuel 2021, 285, 119184. [Google Scholar] [CrossRef]
  59. Yildiz, M.E.; Kokini, J.L. Determination of Williams–Landel–Ferry constants for a food polymer system: Effect of water activity and moisture content. J. Rheol. 2001, 45, 903–912. [Google Scholar] [CrossRef]
  60. Altay, F.; Gunasekaran, S. Mechanical spectra and calorimetric evaluation of gelatin–xanthan gum systems with high levels of co-solutes in the glassy state. Food Hydrocolloid. 2013, 30, 531–540. [Google Scholar] [CrossRef]
  61. Peleg, M. On the use of the WLF model in polymers and foods. Crit. Rev. Food Sci. 1992, 32, 59–66. [Google Scholar] [CrossRef]
  62. Li, M.; Li, D.; Wang, L.; Adhikari, B. Creep behavior of starch-based nanocomposite films with cellulose nanofibrils. Carbohyd. Polym. 2015, 117, 957–963. [Google Scholar] [CrossRef]
  63. Shi, A.; Wang, L.; Li, D.; Adhikari, B. Characterization of starch films containing starch nanoparticles. Part 2: Viscoelasticity and creep properties. Carbohyd. Polym. 2013, 96, 602–610. [Google Scholar] [CrossRef]
  64. Laplante, G.; Lee-Sullivan, P. Moisture effects on FM300 structural film adhesive: Stress relaxation, fracture toughness, and dynamic mechanical analysis. J. Appl. Polym. Sci. 2005, 95, 1285–1294. [Google Scholar] [CrossRef]
  65. Ngai, K.L.; Rajagopal, A.K.; Teitler, S. Slowing down of relaxation in a complex system by constraint dynamics. J. Chem. Phys. 1988, 88, 5086–5094. [Google Scholar] [CrossRef]
  66. Enrione, J.I.; Sáez, C.; López, D.; Skurtys, O.; Acevedo, C.; Osorio, F.; Macnaughtan, W.; Hill, S. Structural relaxation of salmon gelatin films in the glassy state. Food Bioprocess Technol. 2012, 5, 2446–2453. [Google Scholar] [CrossRef]
Figure 1. Relaxation modulus data for maize kernels at different temperatures: (A) natural drying (ND); (B) hot air/vacuum drying (HVD) 35 °C; (C) HVD 45 °C; (D) HVD 55 °C; (E) HVD 65 °C; (F) HVD 75 °C.
Figure 1. Relaxation modulus data for maize kernels at different temperatures: (A) natural drying (ND); (B) hot air/vacuum drying (HVD) 35 °C; (C) HVD 45 °C; (D) HVD 55 °C; (E) HVD 65 °C; (F) HVD 75 °C.
Foods 12 00738 g001
Figure 2. Creep data for maize kernels at different temperatures: (A) ND; (B) HVD 35 °C; (C) HVD 45 °C; (D) HVD 55 °C; (E) HVD 65 °C; (F) HVD 75 °C.
Figure 2. Creep data for maize kernels at different temperatures: (A) ND; (B) HVD 35 °C; (C) HVD 45 °C; (D) HVD 55 °C; (E) HVD 65 °C; (F) HVD 75 °C.
Foods 12 00738 g002aFoods 12 00738 g002b
Figure 3. The left-hand graphs show logarithmic plots of stress relaxation data for maize kernels, while the right-hand graphs are the corresponding master curves for maize kernels (A1,A2)—ND; (B1,B2)—HVD 35 °C; (C1,C2)—HVD 45 °C; (D1,D2)—HVD 55 °C; (E1,E2)—HVD 65 °C; (F1,F2)—HVD 75 °C).
Figure 3. The left-hand graphs show logarithmic plots of stress relaxation data for maize kernels, while the right-hand graphs are the corresponding master curves for maize kernels (A1,A2)—ND; (B1,B2)—HVD 35 °C; (C1,C2)—HVD 45 °C; (D1,D2)—HVD 55 °C; (E1,E2)—HVD 65 °C; (F1,F2)—HVD 75 °C).
Foods 12 00738 g003aFoods 12 00738 g003bFoods 12 00738 g003c
Figure 4. The left-hand graphs show logarithmic plots of creep data for maize kernels, while the right-hand graphs are the corresponding master curves for maize kernels (A1,A2)—ND; (B1,B2)—HVD 35 °C; (C1,C2)—HVD 45 °C; (D1,D2)—HVD 55 °C; (E1,E2)—HVD 65 °C; (F1,F2)—HVD 75 °C.
Figure 4. The left-hand graphs show logarithmic plots of creep data for maize kernels, while the right-hand graphs are the corresponding master curves for maize kernels (A1,A2)—ND; (B1,B2)—HVD 35 °C; (C1,C2)—HVD 45 °C; (D1,D2)—HVD 55 °C; (E1,E2)—HVD 65 °C; (F1,F2)—HVD 75 °C.
Foods 12 00738 g004aFoods 12 00738 g004bFoods 12 00738 g004c
Figure 5. Shift factors fitted to the Arrhenius equation for maize kernels (stress relaxation): (A1) ND; (A2) HVD 35 °C; (A3) HVD 45 °C; (A4) HVD 55 °C; (A5) HVD 65 °C; (A6) HVD 75 °C. Shift factors fitted to the William–Landel–Ferry (WLF) equation for maize kernels (stress relaxation): (B1) ND; (B2) HVD 35 °C; (B3) HVD 45 °C; (B4) HVD 55 °C; (B5) HVD 65 °C; (B6) HVD 75 °C.
Figure 5. Shift factors fitted to the Arrhenius equation for maize kernels (stress relaxation): (A1) ND; (A2) HVD 35 °C; (A3) HVD 45 °C; (A4) HVD 55 °C; (A5) HVD 65 °C; (A6) HVD 75 °C. Shift factors fitted to the William–Landel–Ferry (WLF) equation for maize kernels (stress relaxation): (B1) ND; (B2) HVD 35 °C; (B3) HVD 45 °C; (B4) HVD 55 °C; (B5) HVD 65 °C; (B6) HVD 75 °C.
Foods 12 00738 g005aFoods 12 00738 g005bFoods 12 00738 g005c
Figure 6. Shift factors fitted to the Arrhenius equation for maize kernels (creep): (A1) ND; (A2) HVD 35 °C; (A3) HVD 45 °C; (A4) HVD 55 °C; (A5) HVD 65 °C; (A6) HVD 75 °C. Shift factors fitted to the WLF equation for maize kernels (creep): (B1) ND; (B2) HVD 35 °C; (B3) HVD 45 °C; (B4) HVD 55 °C; (B5) HVD 65 °C; (B6) HVD 75 °C.
Figure 6. Shift factors fitted to the Arrhenius equation for maize kernels (creep): (A1) ND; (A2) HVD 35 °C; (A3) HVD 45 °C; (A4) HVD 55 °C; (A5) HVD 65 °C; (A6) HVD 75 °C. Shift factors fitted to the WLF equation for maize kernels (creep): (B1) ND; (B2) HVD 35 °C; (B3) HVD 45 °C; (B4) HVD 55 °C; (B5) HVD 65 °C; (B6) HVD 75 °C.
Foods 12 00738 g006aFoods 12 00738 g006b
Figure 7. Time-temperature superposition (TTS) predicted stress relaxation for maize kernels dried under different conditions: (A) ND; (B) HVD 35 °C; (C) HVD 45 °C; (D) HVD 55 °C; (E) HVD 65 °C; (F) HVD 75 °C.
Figure 7. Time-temperature superposition (TTS) predicted stress relaxation for maize kernels dried under different conditions: (A) ND; (B) HVD 35 °C; (C) HVD 45 °C; (D) HVD 55 °C; (E) HVD 65 °C; (F) HVD 75 °C.
Foods 12 00738 g007aFoods 12 00738 g007b
Figure 8. TTS predicted creep for maize kernels dried under different conditions: (A) ND; (B) HVD 35 °C; (C) HVD 45 °C; (D) HVD 55 °C; (E) HVD 65 °C; (F) HVD 75 °C.
Figure 8. TTS predicted creep for maize kernels dried under different conditions: (A) ND; (B) HVD 35 °C; (C) HVD 45 °C; (D) HVD 55 °C; (E) HVD 65 °C; (F) HVD 75 °C.
Foods 12 00738 g008aFoods 12 00738 g008b
Table 1. Parameters of the generalized Maxwell equation for maize kernels.
Table 1. Parameters of the generalized Maxwell equation for maize kernels.
Drying ConditionT (°C)E1 (MPa)τ (s)E0 (Mpa)R2
ND253.893155.3246.8940.970
455.934127.6483.0240.973
653.855101.1951.1500.960
852.58070.4490.5020.949
HVD 35 °C253.349231.2583.6090.982
454.180168.6970.6310.989
652.265122.9050.6740.971
852.33787.1830.2030.964
HVD 45 °C252.715198.5212.9150.982
454.308194.8580.7380.990
652.04582.3440.2080.977
852.23756.4940.2250.976
HVD 55 °C254.893211.0595.2520.983
453.368187.6633.3740.980
654.495118.8420.8660.984
854.144119.8100.3590.973
HVD 65 °C253.896206.4804.5160.981
453.848204.3991.5970.989
652.994163.3610.5060.980
853.26986.0220.3640.962
HVD 75 °C252.867154.0095.0580.974
452.260129.4102.6350.974
652.791124.4491.7430.976
851.80387.1040.3670.980
Table 2. Parameters of Burgers’ equation for maize kernels.
Table 2. Parameters of Burgers’ equation for maize kernels.
Drying
Condition
T (°C)E1 (MPa)E2 (MPa)η2 (MPa s)η1 (MPa s)τ (s)ε’ (∞) (×10−6 s−1)R2
ND3012.36693.5132123.7523.343 × 10422.7112.3930.986
409.45944.6211960.3071.500 × 10443.9325.3350.991
508.55029.8931429.9971.362 × 10447.8375.8740.991
608.03715.943884.3836667.17255.47211.9990.995
HVD 35 °C307.87444.3991249.6391.758 × 10428.1464.5510.986
407.51039.7701249.1931.169 × 10431.4106.8460.991
505.54621.809782.9029165.19935.8988.7290.989
605.39212.599869.3415301.57469.00115.0900.995
HVD 45 °C305.86042.303822.9371.662 × 10419.4534.8130.982
405.34621.779685.7197719.94931.48510.3630.990
505.51115.600752.9685336.19648.26714.9920.993
605.8589.698377.3935333.76638.91514.9990.990
HVD 55 °C307.35250.1901277.3061.815 × 10425.4494.4090.985
405.87618.000778.4717273.83943.24810.9980.991
504.25518.021530.0686427.00429.41412.4470.988
603.6719.550512.9694199.83953.71419.0480.992
HVD 65 °C309.74165.2452054.1072.380 × 10431.4833.3610.987
407.92639.0791085.3281.463 × 10427.7735.4670.987
507.54928.8801011.7849554.11935.0348.3730.991
606.95622.999957.6375738.02941.63813.9420.994
HVD 75 °C307.81657.8581335.1591.856 × 10423.0764.3110.986
406.28830.236975.9341.119 × 10432.2777.1490.988
505.26318.040602.6058122.87333.4049.8490.988
604.89620.384621.9517975.71330.51210.0300.988
Table 3. WLF constants and Arrhenius activation energy for maize kernels (stress relaxation).
Table 3. WLF constants and Arrhenius activation energy for maize kernels (stress relaxation).
EquationParametersDrying Condition
NDHVD
35 °C
HVD
45 °C
HVD
55 °C
HVD
65 °C
HVD
75 °C
WLFC114.9905.9958.4067.1747.7585.882
C2 (K)191.70056.655134.40072.00171.98443.807
R20.9970.9990.9530.9920.9860.857
ArrheniusEa126.900124.20090.860117.300126.600134.800
R20.9970.9590.9340.9670.9600.963
Table 4. WLF constants and Arrhenius activation energy for maize kernels (creep).
Table 4. WLF constants and Arrhenius activation energy for maize kernels (creep).
EquationParametersDrying Condition
NDHVD
35 °C
HVD
45 °C
HVD
55 °C
HVD
65 °C
HVD
75 °C
WLFC16.12732.0202.22615.8734.4359.671
C2 (K)17.771257.5008.34673.29713.64740.668
R20.9980.9630.9850.9980.9660.989
ArrheniusEa250.599220.498111.035292.645199.302255.960
R20.9320.9450.8700.9870.9360.955
Table 5. Equation parameters for long-period stress relaxation of maize kernels.
Table 5. Equation parameters for long-period stress relaxation of maize kernels.
EquationParametersDrying Condition
NDHVD 35 °CHVD 45 °CHVD 55 °CHVD 65 °CHVD 75 °C
Generalized Maxwell equationE1 (MPa)4.4492.5431.1182.3262.2971.753
E2 (MPa)2.8021.7961.6743.5552.7633.166
E3 (MPa)2.8582.5732.9234.0503.2532.485
τ1 (s)2177.7954704.46955.1488945.7495866.5358.865 × 105
τ2 (s)123.8261.676 × 1052.760 × 1041.604 × 1052.716 × 105242.832
τ3(s)8.286 × 104155.342821.170143.129146.0802.284 × 104
E0 (MPa)0.6680.0950.0950.3220.1720.365
R20.9940.9800.9900.9920.9870.996
Kohlrausch–Williams–Watts (KWW) equationE0 (MPa)16.50110.6707.84713.55414.00612.593
τ1059.509958.363798.5413888.804791.2911602.291
β0.1940.1910.2710.1870.1490.145
R20.9920.9780.9840.9740.9780.995
Table 6. Equation parameters for long-period creep of maize kernels.
Table 6. Equation parameters for long-period creep of maize kernels.
EquationParametersDrying Condition
NDHVD 35 °CHVD 45 °CHVD 55 °CHVD 65 °CHVD 75 °C
Burgers equationE1 (MPa)10.2446.9685.2516.0818.1616.503
E2 (MPa)17.79513.75411.9297.08118.00411.459
η2 (MPa s)4.317 × 1052.145 × 1043.291 × 1041.142 × 1051.789 × 1051.821 × 105
η1 (MPa s)3.389 × 1074.294 × 1063.126 × 1056.964 × 1078.018 × 1064.386 × 107
R20.9510.9560.9660.9500.9460.922
Two-parameter power law equationa0.0050.0080.0100.0080.0060.009
b0.0880.0830.0970.0880.0800.067
R20.9050.9320.8790.9700.9120.970
Findley power law equationa2.418 × 10−40.0012.829 × 10−40.0034.020 × 10−40.003
b0.2630.2540.3920.1480.2540.113
ε00.0070.0100.0140.0080.0080.007
R20.9780.9770.9740.9820.9740.976
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Sheng, S.; Shi, A.; Xing, J. A Systematical Rheological Study of Maize Kernel. Foods 2023, 12, 738. https://doi.org/10.3390/foods12040738

AMA Style

Sheng S, Shi A, Xing J. A Systematical Rheological Study of Maize Kernel. Foods. 2023; 12(4):738. https://doi.org/10.3390/foods12040738

Chicago/Turabian Style

Sheng, Shaoyang, Aimin Shi, and Junjie Xing. 2023. "A Systematical Rheological Study of Maize Kernel" Foods 12, no. 4: 738. https://doi.org/10.3390/foods12040738

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