Next Article in Journal
Urban Regeneration Assessment in Castilla and León (Spain): A Comparative Study of Proposals in Homogeneous Residential Neighborhoods Built Up during Developmentalism
Previous Article in Journal
Comparative Life Cycle Assessment of Four Municipal Water Disinfection Methods
Previous Article in Special Issue
Life Cycle CO2 Emissions Analysis of a High-Tech Greenhouse Horticulture Utilizing Wood Chips for Heating in Japan
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Research on Thermal Environment of Container Farms: Key Factor Identification and Priority Analysis

1
School of Mechanical Engineering, Tongji University, Shanghai 201804, China
2
Institute for Carbon Neutrality, Tongji University, Shanghai 200092, China
3
Internet-Based Collaborative Research Center on Protected Agriculture, Tongji University, Shanghai 200092, China
4
National Engineering Technology Research Center for Prefabrication Construction in Civil Engineering, Tongji University, Shanghai 200092, China
5
College of Architecture and Energy Engineering, Wenzhou University of Technology, Wenzhou 325027, China
*
Author to whom correspondence should be addressed.
Sustainability 2024, 16(14), 6105; https://doi.org/10.3390/su16146105
Submission received: 25 June 2024 / Revised: 13 July 2024 / Accepted: 15 July 2024 / Published: 17 July 2024
(This article belongs to the Special Issue Controlled Environment Agriculture for Sustainable Farming)

Abstract

:
Container farms (CFs), a controlled environment agricultural technology designed to solve food insecurity, are receiving increasing attention from researchers. However, the complex geometric structures and artificial lighting used in CFs present challenges in effectively controlling the thermal environment. This study aims to identify the primary factors that impact the thermal environment of CFs while conducting factor ranking and significance analysis, providing a theoretical basis for future thermal environment optimization. The research method of theoretical analysis, CFD simulation, and an orthogonal experimental design were adopted to achieve the above objectives. Theoretical analysis revealed that factors influencing the thermal environment are the HVAC system’s supply air temperature, humidity, flow rate, and the light source used. Four evaluation indices, including the mean value and range between layers of temperature and moisture content, were used. The results revealed that supply air temperature and light source are significant for mean temperature, while supply air temperature and humidity are significant for mean moisture content. In the case of range between layers, supply air flow rate and light source display a significant correlation. These findings suggest that future optimization should prioritize the regulation of the HVAC system’s supply air and light source.

1. Introduction

1.1. Background

Global warming is a pervasive environmental issue that significantly threatens food security [1]. Affected by environmental problems, such as land desertification and water shortages [2], per capita cultivated land is projected to be less than one-third of the 1970 levels by 2050 [3]. Controlled environment agriculture (CEA) has emerged as a potential solution [4,5]. It comprises two main types: greenhouses and vertical farming [6]. Vertical farming, specifically plant factories, represents a viable approach [7] as it minimizes land requirements [8,9] and reduces water waste [10] while enabling year-round production [6]. Consequently, plant factories have gained significant popularity on the market.
Container farms (CFs) are a type of plant factory constructed by modifying standard-sized metal containers, such as 20 or 40-foot ones. Their unique origin and structure afford CFs advantages such as ease of transportation, rapid construction, and adaptability to various scenarios, setting them apart from traditional vertical farming methods [11]. In areas where conventional planting methods are not feasible, such as deserts, islands, and highlands, CFs can offer stable crop planting environments to farmers. In terms of commercial applications, CFs offer the prospect of high annual productivity and efficient space utilization [12,13].

1.2. Overviews and Limitations

Several studies have emphasized the significant influence of the thermal environment on the growth of various plant species, such as tomatoes [14,15], cucumbers [16,17], and lettuce [18,19]. As a result, controlling the temperature and humidity, which are key elements of the thermal environment, has become a primary concern in CFs. CFs are a specialized form of CEA, sharing many of the same challenges in terms of maintaining optimal conditions for plant growth. However, the uniqueness of CFs lies in their planting systems, including complex geometric structures of planting racks and artificial lighting required for photosynthesis, which poses challenges in achieving precise temperature and humidity control, as well as maintaining energy efficiency within the enclosed systems. Ventilation has emerged as a prominent method for controlling the thermal environment in current CEA practices [20]. Consequently, researchers have focused on optimizing internal airflow to improve the thermal environment [9,21]. For instance, researchers [21] have employed CFD simulations to analyze the impact of various ventilation duct configurations on airflow uniformity and the thermal environment within indoor vertical farms. Nevertheless, there exist notable differences between this expansive indoor vertical farm setting and CFs. Specifically, in a CFD simulation study focused on CFs, the researchers primarily scrutinized airflow uniformity, overlooking the consideration of the thermal environment [9]. Although some studies have explored the impact of ventilation on the thermal environment, comprehensive investigations on this specific issue are limited.
In terms of control facilities, some researchers examined a mechanical ventilation system to enhance temperature control in CEA during the summer season [22]. A new temperature controller and a novel multivariable control strategy were also proposed [23,24]. Although these researchers improved temperature regulation in CFs, they may not have thoroughly addressed the factors influencing the thermal environment. Regarding specific factors affecting the thermal environment, some researchers found a negative correlation between crop height and indoor temperature differences, observing that taller crops could lead to uneven relative humidity [25]. Vent locations were demonstrated to impact temperature distribution significantly [26]. While these studies have identified some location factors influencing the thermal environment, there is still a need to uncover many more, such as the parameters of air conditioning and light sources. Moreover, a more quantitative evaluation of these specific factors is necessary to understand their precise effects on temperature and humidity within CFs.

1.3. Aims and Methods

This study aims to comprehensively investigate and quantitatively evaluate the factors that impact the thermal environment of CFs. The objective is to establish a theoretical foundation for future research endeavors on thermal environment optimization and energy efficiency in CFs. Through theoretical analysis of the energy and mass conservation equations within CFs, specific factors that impact the thermal environment are extracted. Furthermore, quantitative prioritization and significance analysis of these specific factors will be performed using orthogonal experimental simulation. By identifying and analyzing these factors, researchers and CF users can gain a deeper understanding of the thermal environment, which, in turn, will facilitate the application and optimization of CFs.
In terms of research methodology, computational fluid dynamics (CFD) simulation has been widely employed as a powerful tool for evaluating the thermal environment in CEA studies [9,21,27,28,29,30,31,32]. This is supported by numerous studies focusing on crop heat and mass transfer modeling [2,33,34]. Consequently, this research adopts a CFD numerical simulation approach. A numerical model of the plant canopy within the CF is established (Section 2.1) and incorporated as source terms in the CFD simulation. Simulations are then conducted based on different cases (Section 2.2). Experimental data are used for model validation to ensure the accuracy of the simulations (Section 3.1).

2. Materials and Methods

2.1. Canopy Numerical Model

In order to simplify the calculations, accurate crop models were not developed in this study. Instead, some CFD simulation studies have utilized porous media to represent the plant canopy [21,30,31]. Within CFs, the energy exchange between plants and the environment occurs through sensible heat, latent heat, and radiation [2]. This can be described using Equation (1):
R LED = Q sen + Q lat
where RLED can be calculated using Equation (2) [21]:
R LED = α LED P LED η LED C A C
where αLED is 0.93 [33], PLED is 12 W, ηLED is 52%, and CAC is 0.9 in this study [34].
For the canopy model, the difficulty lies in calculating sensible heat and latent heat (water evaporation). The sensible heat is driven by the temperature difference, as follows [21]:
Q sen = L A I ρ a c p a ( T leaf T a ) r a 1
According to research [2], LAI is taken as 2.96 m2/m2 in this study. The aerodynamic resistance (ra) is a parameter related to l and u, which is given by Equation (4) [34]:
r a = 350 l / u L A I 1
where l is taken as 0.1 m [21].
The water vapor pressure difference between the canopy and the surrounding air drives the latent heat exchange of the canopy. According to the research in vertical farms [34], the latent heat can be given by
Q lat = λ w E T = λ w L A I ω leaf ω a r a + r s
where ωleaf and ωa are given by Equations (6) and (7), respectively, and rs is given by Equation (8).
ω leaf = P w - leaf M w R mol T leaf
ω a = R H a P w - a M w R mol T a
r s = 60 1500 + P P F D 200 + P P F D
Pw-leaf and Pw-a can be calculated from the Clausius–Clapeyron equation, see Equation (9); the PPFD of LED in this study is 250 μmol/m2 s.
P w = 611 exp [ λ w M w R mol ( 1 273.15 1 T ) ]
Equations (1)–(9) together form the canopy heat and mass transfer model, in which the unknown parameters include Tleaf, Ta, u, and RHa. Ta, u, and RHa can be obtained through simulation, and the only unknown parameter in the model is Tleaf. Then, the leaf temperature can be calculated from Equation (1). Write Equations (1)–(9) as user-defined functions (UDF) and load them into CFD.

2.2. Case Design

2.2.1. Variable Extraction

This study aims to examine the influencing variables of the thermal environment inside the CF and their degree of influence; therefore, the variables need to be extracted first. To facilitate the theoretical analysis, the following assumptions have been applied:
(1)
The airtightness of the CF under study is excellent, meaning that the penetration of external wind is negligible.
(2)
The auxiliary facilities in the CF, such as water pumps and fans, maintain stability and a constant calorific value.
(3)
The research focuses on the summer season, and the impact of the external environment changes at different moments is not considered.
The energy balance equation of the CF is as follows [2]:
Q canopy + Q LED + Q equipment + Q envelope + Q HVAC + Q infiltration = 0
According to the assumptions, Qinfiltration can be considered 0 and Qequipment constant. Qcanopy, QLED, Qenvelope, and QHVAC are given by Equations (11)–(14), respectively:
Q canopy = Q sen + Q lat
Q LED = N η LED P LED
Q envelope = K A envelope ( T o T a )
Q HVAC = ρ a c p a q s ( T s T a )
According to Equation (10), the unknown variables that affect Ta are PLED, qs, Ts, Tleaf, u, and RHa. Tleaf is a variable related to Ta (Section 2.1), and u is a variable related to qs (the supply air mode is fixed), so there is the following equation:
T a = f ( P LED , q s , T s , R H a )
Similar to the energy equation, the humidity in the CF is analyzed through the following mass balance:
m canopy + m HVAC + m media = 0
m canopy = V canopy E T L A D
m HVAC = q s ( ω s ω a )
In Equation (16), mmedia is 0 because the media where lettuce grows is covered [35]. Similarly, the function of humidity can be obtained:
R H a = f ( T a , q s , T s , R H s )
Equations (15) and (19) show that the CF’s thermal environment (Ta, RHa) is influenced by PLED, qs, Ts, and RHs. In other words, the light source and HVAC system supply air parameters affect the thermal environment inside the CF.

2.2.2. Factor Level

The orthogonal experimental design, particularly suitable for investigating multifactor and multilevel problems, offers a cost-effective alternative to traditional enumeration methods by leveraging orthogonality to select representative factors and levels from a comprehensive test [36]. It is an efficient and scientifically rigorous approach that helps elucidate the significance of various factors influencing test indicators. Many researchers have utilized the orthogonal experimental method to effectively reduce the number of simulations [36,37,38]. In line with this, we conducted simulations using the orthogonal experimental method. Four factors were extracted in Section 2.2.1 and considered in the orthogonal experimental design. Table 1 shows each influencing factor and level value, and Table 2 is the corresponding orthogonal design table. The four levels in Table 1 were selected based on the air conditioning supply parameters adopted during the experiment presented in Section 2.5.
In Table 1, the LED duty cycle is a current light source modification solution. Several scholars have discovered that altering the duty cycle of the LED to create pulsed light rather than continuous light (duty cycle of 1) [39,40,41] does not adversely affect crop growth. By appropriately setting a reasonable duty cycle and frequency, the photosynthetic efficiency can approach that of continuous light. The duty cycle determines the time the LED remains powered on within a given unit time. When applied to CFD simulation, this can be regarded as a multiplication of PLED by the duty cycle, thus modifying the parameters of the PLED.

2.3. Physical Models and Grids

The physical model employed in this study is based on an actual CF (located in Shanghai, China). As illustrated in Figure 1, the CF consists of five planting layers on both the left and right sides, with each layer equipped with 40 LED light strips (SANANBIO, Quanzhou, China). The physical model is created using the DesignModeler component in Workbench 2020 R2. Fluent Meshing (Ansys, Canonsburg, PA, USA) is employed for grid division. Additionally, certain boundary conditions are encrypted, and boundary layers are generated to ensure accurate simulations.
To ensure accuracy, five sets of grids were established with different levels of precision. Figure 2 presents the independence analysis of these grids. The average relative error between the standard (2.14 million) and the precision grid (3.42 million) is 0.46%, while between the fine grid (2.95 million) and the precision grid (3.42 million), it is 0.68%. However, both the coarse grid (1.10 million) and general grid (1.51 million) exhibit average relative errors above 1.5%. The max relative errors are 15.63%, 6.21%, 3.19%, and 4.46% for the coarse, general, standard, and fine grids, respectively. Based on the results of the independence analysis, the standard grid (2.14 million) was selected for simulations.

2.4. Solution Setting

The simulation is conducted with Fluent 2020 R2. The primary investigation is the canopy plane’s thermal environment (temperature, humidity) in the CF. The simulation is conducted under the steady state [30].

2.4.1. Solution Equations

The energy, turbulent viscosity, radiation, and species models were turned on. The equations involved, which include continuity, momentum, energy, and species, are as follows:
x i ( ρ i u i ) = 0
x i ( ρ i u i u j ) = x i P δ ij + x i μ t ( u i x j + u j x i ) + ρ i g i + S m
c p x i ( ρ i T u i ) = x i ( λ f + μ t c p P r t ) T x i + S s
x i ( ρ i C u i ) = φ ij x i + S w
where i and j represent the three directions of the three-dimensional Cartesian coordinate system; ρi adopts the Boussinesq assumption in the simulation; φij results from temperature and concentration gradients. There are three source terms in the equation. Sm represents the momentum source term generated due to porous media, Ss is the heat source term generated by sensible heat exchange of lettuce, and Sw is the species source term of lettuce water vapor, as follows:
S m = ( μ α u i + 0.5 C 2 ρ | u | u i )
S s = Q sen L A D
S w = E T L A D
In Equations (25) and (26), LAD is considered 2.467 [2].

2.4.2. Model Selection and Solver Settings

The Realizable k-ε model has been utilized in the research of turbulence flow in greenhouse and vertical farms [27,28,29,32]. Compared to the standard k-ε model, it has demonstrated improved accuracy in predicting jet diffusion [32], especially in the context of the high-speed fan utilized in our simulations. This improved accuracy is particularly advantageous for capturing the dynamics of the jet generated by the high-speed fan, ensuring more reliable predictions and insights into the flow behavior. For the radiation model, the discrete ordinates (DO) model was chosen due to its suitability for complex structural scenes. DO models have been successfully employed in various agricultural environment simulations [21,30].
The Pressure-Based algorithm was used in the simulation, with the SIMPLE algorithm employed to solve the pressure–velocity coupling due to the presence of porous media and porous–jump boundary conditions. Except for the pressure term, which utilized PRESTO! [21], second-order upwind discretization schemes were applied for the remaining terms. To save computational costs, adjustments were made to the residuals used as convergence criteria. Turbulence residuals were set to less than 1 × 10−3, while continuity and momentum residuals were set to less than 1 × 10−4. The DO radiation residual was set to less than 1 × 10−5, and both energy and species residuals were set to less than 1 × 10−6.

2.4.3. Boundary Conditions Settings

The settings of all boundary conditions (BCs) of the model are shown in Table 3.

2.5. Experiment Setup

In December 2022, a planting trial was conducted to validate the accuracy of the model. In all layers except the fifth planting layer, 100 heads of Italian lettuce (Lactuca sativa L. var. ramosa Hort) were planted on both frames, resulting in a total of 800 heads. Figure 3 illustrates the positioning of nineteen measurement locations within each layer, which were used to collect temperature and humidity data near the crop canopy (0.05 m away from the plants). Within the planting area on the first to fourth planting layers, four measurement points were placed, with a spacing of 0.5 m from the front and rear of the planting frame. On the fifth planting layer, two measuring points were situated at the center of the planting frame. The last one was at the center of the outlet. The Temperature and Humidity Data Logger (Testo 175H1, Testo, Titisee-Neustadt, Germany) was used at these 19 measuring points. To collect input parameters for the simulation, a hot wire anemometer (VelociCalc 9565, TSI, Shoreview, MN, USA) was employed to measure the supply air velocity at the inlet. The Temperature and Humidity Data Logger (COS03, Renke, Jinan, China) recorded the temperature and humidity of the supply and outdoor air. A detailed description of all instruments can be found in Appendix A. The Temperature and Humidity Data Logger was placed at each measuring point for seven days, while only data from the light period were selected. Simultaneously, the hot wire anemometer was conducted during the same time frame. Since the supply air flow rate remained stable throughout the experiment, the average value was used.

2.6. Evaluation Index

The air around the canopy of each layer is divided to present the results and discuss them more intuitively. Taking the first planting layer (L1) as an example, as shown in Figure 4, the plane is divided into 200 areas. Each area’s height is 0.05 m above the crop.
To assess the relative influence of different factors on the thermal environment around the canopy, the arithmetic mean of temperature and moisture content (MT, MMC) for each layer was used as an indicator for data analysis. Furthermore, to evaluate the variations between layers, the range between layers (RBL) was calculated as the range of mean differences. These two indicators, MT and MMC, provide insights into the overall temperature and moisture levels within each layer, respectively, while the RBL takes into consideration the variability between layers.

3. Results and Discussion

3.1. Validation

The comparison between the test and simulated data for temperature and humidity is presented in Figure 5. The relative error between simulation and experimental data is commonly used as a metric to evaluate the acceptability of simulation results and is typically set at around 5%, taking into account factors such as instrument inaccuracies, measurement uncertainties, model simplifications, and parameter variations. In Figure 5a, most of the simulated values fall within the box region of the test values, indicating a good agreement. The majority of the relative errors between the simulation and the test are less than 5%. The average error for temperature is 0.22 °C, and the average relative error is 2.59%, both falling within the measurement uncertainty of the instrument.
In Figure 5b, humidity is expressed as moisture content. Only a few individual values deviate slightly from the uncertainty range, and all the differences in humidity do not exceed 1 g/kg. The average error for humidity is 0.39 g/kg, which, when converted to relative humidity, is equivalent to 2.88%, close to the instrument’s uncertainty. The average relative error is 5.02%. It is important to note that considering all lettuces in the simulation are at the same growth stage (with the same LAI) compared to the actual scenario (with different LAI), these deviations are acceptable. The verification work demonstrates that the established model effectively represents the temperature and humidity distribution within the CF.
However, it is necessary to pay attention to the excessive deviations observed at specific measuring points. Measurement location 11 exhibits the largest temperature discrepancy, with a relative error of 6.43%. Measuring points 6 and 16 display significant humidity gaps (0.989 g/kg and 0.842 g/kg, respectively). For temperature, small deviations in the positioning of measuring points, especially those influenced by high-speed fans (as in the case of measuring point 6), can result in considerable temperature errors [21]. Neglecting geometric details, such as hydroponic system accessories near measuring locations 6 and 16, may contribute to humidity errors.

3.2. Thermal Environment of Canopy Plane

3.2.1. Temperature Distribution

In Section 2.6, the canopy plane is divided into 200 areas, and the mean temperature (MT) and standard deviation (SD) for each layer obtained from the simulation in case 4 are calculated and presented in Figure 6. The results for all cases can be found in Figures S1–S16 in the Supplementary Material.
Figure 6 reveals that the majority of high SD areas are located in Layer 1 and Layer 2 predominantly concentrated on the left side of the canopy plane, indicating a significant temperature gradient in a specific portion of the left canopy area. Across all layers, the SD in the edge area exhibits a substantial increase, which can be attributed to the difference between the temperatures in the non-planted area and the canopy itself. When comparing different layers, the SD gradually decreases with the increasing height of the layer. This can be attributed to the high-speed airflow generated by the fan and the air supply duct, which ensures effective air circulation in the upper layers, enabling the timely discharge of hot air and promoting a uniform temperature distribution.
In most areas, MT on the left side is higher than on the right due to airflow tendencies. Nevertheless, there are instances when the MT is lower on the left side. Figure 6 indicates that in Layer 3, the MT is generally higher on the right side than on the left. The velocity vector diagram of the L3 canopy plane depicted in Figure 7 clarifies this observation, revealing the presence of four vortex areas. Apart from the two prominent vortex areas, a smaller one can be observed on the right side of the canopy (top of Figure 7), along with a tiny vortex on the left side. The air surrounding these vortex areas exhibits low velocity and sparse vector distribution. The existence of these two vortexes corresponds with the high MT area identified in Figure 6.

3.2.2. Moisture Distribution

The MMC and SD of each area in case 4 are calculated and shown in Figure 8. The results of all cases can be viewed in Figures S17–S32 in the Supplementary Material.
As seen in Figure 8, the areas of high SD of humidity are larger than the temperature, revealing that the vortex zone has a more pronounced impact on moisture. No direct relationship between high MMC and high SD is observed. Examining both Figure 6 and Figure 8, it can be noted that the MMC follows a similar trend as the MT.

3.3. Factor Analysis

Figure 9 displays each case’s average value and RBL using the evaluation index in Section 2.6. It is evident that across all 16 cases, there are minimal differences in temperature and humidity from the first to the fourth planting layers, while a significant decrease is observed in the fifth planting layer. The layer at which the maximum value occurs varies across cases, whereas the minimum value consistently emerges in the fifth planting layer. Case 15 exhibits the highest MT in Layer 4, with a value of 28.70 °C, while case 13 has the highest MMC in Layer 1, with a value of 14.388 g/kg. The smallest MT and MMC values are observed in case 1, with 18.03 °C and 6.658 g/kg, respectively, in Layer 5. Regarding RBL, case 9 demonstrates the maximum TRBL at 2.38 °C, and it also has the highest MCRBL at 0.712 g/kg. On the other hand, case 12 exhibits the minimum TRBL at 0.47 °C, while case 8 has the minimum MCRBL at 0.140 g/kg.

3.3.1. Range Analysis

Orthogonal experiments offer the advantage of providing a comprehensive and comparable understanding of the influence of changing factors within an experiment. This understanding is typically obtained by analyzing the range (R) of the average values for each level of the factors [36,37,42]. The size of R is positively associated with the degree of influence exerted by the factors. Table 4 lists the range analysis table of 12 indicators. The specific calculation of each indicator can be found in Table S1 in the Supplementary Material.
According to Table 4, the factors influencing MT and MMC in order of significance are as follows: for MT, it is A > D > B > C, while for MMC, it is A > C > B > D. Hence, changes in supply air temperature have the most substantial impact on both MT and MMC. For MT, the LED duty cycle is second, while the flow rate and humidity of the supply air have little impact. In the case of MMC, the humidity is second, whereas the others have minimal effects.
The order of factors influencing TRBL is D > B > C > A, whereas for MCRBL, it is B > D > C > A. For TRBL, the LED duty cycle has the highest impact, followed by the supply air flow rate. Conversely, for MCRBL, the flow rate is the highest. Both temperature and humidity have little effect on the two RBLs.
Figure 10 provides a visual representation depicting how each factor influences MT, MMC, TRBL, and MCRBL. The consistency of the range rules of MT and MMC across each layer allows the use of L1 data as an illustrative example.
When factor A increases, both MT and MMC exhibit significant increases, TRBL shows a relatively mild change, and MCRBL experiences a slight increase. When factor B increases, all indices decrease, but the RBLs decrease significantly. When factor C increases, MMC exhibits a significant increase, while MCRBL shows a downward trend. When factor D increases, apart from MMC, the other three indicators exhibit slight changes initially, followed by a rapid increase.

3.3.2. Analysis of Variance

An analysis of variance (ANOVA) can be used to determine whether each factor has a significant impact on the results [36,37,42]. The significance level is typically selected at 0.01 and 0.05, and it is determined by the p-value derived from the F-distribution. When p < 0.01, it indicates a highly significant impact of the factor on the results, denoted by “**”. When 0.01 ≤ p < 0.05, it indicates a significant effect, represented by “*”. When p ≥ 0.05, the factor is considered to have no significant impact. The ANOVA results of the orthogonal test are presented in Table 5, and the corresponding data can be found in Table S1 in the Supplementary Material.
As shown in Table 5, considering L1 data as an example (since the results of MT and MMC in each layer are consistent), factors A and D have a highly significant impact on MT, while factors A and C have a significant effect on MMC. For both the two RBLs, factors B and D are significant, whereas factors A and C are not. Moreover, the ranking of the F-values for each indicator is consistent with the R-value in Table 4.

3.4. Limitations

In our CFD simulation, we utilized a steady-state approach. However, it is important to acknowledge that there might exist discrepancies between the simulation and the actual situation due to the dynamic changes occurring within the CF. However, considering the applicability of steady-state simulation for this study, this gap may not be significant. Nonetheless, in future research, transient-state simulations can be attempted for validation. Another aspect to consider is that the canopy model employed in our study focused on specific crop species. It is conceivable that various crop species may exhibit distinctions in the computation of the canopy model. Hence, further discussion, such as controlled variable experiments in future research, is needed to determine whether the model applies to other crops.
Lastly, it is worth noting that the canopy model used in our research is static. However, plant growth is a dynamic process that is influenced by date [2]. The dynamic changes in plant growth involve LAI and LAD values in the canopy model, which can affect the characterization of heat transfer in the canopy model and further affect the calculation conclusions of Tleaf. However, the experimental results in reference [2] indicate that there are no significant differences between different stages. Nonetheless, in order to achieve a higher level of accuracy, future studies should consider proposing a more precise model that accounts for this dynamic nature, such as changing the constant LAI and LAD values to variable ones. Meanwhile, the interaction between factors can also serve as the next research object.

4. Conclusions

This research primarily focused on the thermal environment within CFs. Through theoretical analysis, we identified the key factors that affect canopy temperature and humidity. Subsequently, a physical model and grid for the CF were established and validated using experimental data. Our study employed CFD numerical simulation and conducted an orthogonal experiment to investigate the order and significance of these factors. The main conclusions are as follows:
(1)
The distribution patterns of temperature and humidity within the CF are generally consistent. Heat and moisture tend to accumulate in the vicinity of the outlet due to airflow tendencies. The average values in the first to fourth planting layer exhibit similarity, with a noticeable decrease in the fifth planting layer. The vortex, which has a more significant role in humidity, influences the distribution of the thermal environment.
(2)
Range analysis indicates that the order of factors influencing MT is A > D > B > C, and for MMC, it is A > C > B > D. Regarding TRBL, the order is D > B > C > A, while for MCRBL, it is B > D > C > A. Increasing the supply air temperature leads to a rapid rise in MT and MMC, with slower changes in RBLs. Higher supply air flow rates cause a decrease in all four indicators, with RBLs showing a more pronounced effect. An increase in supply air humidity contributes to a significant rise in MMC, but both RBLs exhibit a downward trend. The LED duty cycle initially yields a gentle effect on the indicators, followed by a rapid increase. However, MCRBL experiences a slight decrease initially.
(3)
ANOVA reveals that the significant factors influencing MT are supply air temperature and LED duty cycle, while supply air temperature and humidity are critical for MMC. This suggests that future efforts in optimizing the thermal environment around CFs should prioritize adjusting the supply air temperature, humidity level, and light source. For both RBLs, supply air flow rate and LED duty cycle demonstrate a significant correlation, emphasizing the importance of these two factors in achieving a more uniform thermal environment.

Supplementary Materials

The following supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/su16146105/s1, Figure S1: The MT and SD in each layer of case 1; Figure S2: The MT and SD in each layer of case 2; Figure S3: The MT and SD in each layer of case 3; Figure S4: The MT and SD in each layer of case 4; Figure S5: The MT and SD in each layer of case 5; Figure S6: The MT and SD in each layer of case 6; Figure S7: The MT and SD in each layer of case 7; Figure S8: The MT and SD in each layer of case 8; Figure S9: The MT and SD in each layer of case 9; Figure S10: The MT and SD in each layer of case 10; Figure S11: The MT and SD in each layer of case 11; Figure S12: The MT and SD in each layer of case 12; Figure S13: The MT and SD in each layer of case 13; Figure S14: The MT and SD in each layer of case 14; Figure S15: The MT and SD in each layer of case 15; Figure S16: The MT and SD in each layer of case 16; Figure S17: The MMC and SD in each layer of case 1; Figure S18: The MMC and SD in each layer of case 2; Figure S19: The MMC and SD in each layer of case 3; Figure S20: The MMC and SD in each layer of case 4; Figure S21: The MMC and SD in each layer of case 5; Figure S22: The MMC and SD in each layer of case 6; Figure S23: The MMC and SD in each layer of case 7; Figure S24: The MMC and SD in each layer of case 8; Figure S25: The MMC and SD in each layer of case 9; Figure S26: The MMC and SD in each layer of case 10; Figure S27: The MMC and SD in each layer of case 11; Figure S28: The MMC and SD in each layer of case 12; Figure S29: The MMC and SD in each layer of case 13; Figure S30: The MMC and SD in each layer of case 14; Figure S31: The MMC and SD in each layer of case 15; Figure S32: The MMC and SD in each layer of case 16; Table S1: Range analysis and ANOVA.

Author Contributions

Conceptualization, Z.N. and D.L.; Methodology, Z.N. and R.S.; Software, Z.N.; Validation, Z.N.; Formal Analysis, Z.N.; Investigation, Z.N.; Resources, D.L.; Data Curation, Z.N.; Writing—Original Draft Preparation, Z.N.; Writing—Review and Editing, Z.N., D.L., C.M., and R.S.; Visualization, Z.N. and C.M.; Supervision, D.L.; Project Administration, D.L.; Funding Acquisition, D.L. All authors have read and agreed to the published version of the manuscript.

Funding

The research has been funded by the Shanghai Agricultural Science and Technology Innovation Technology Research Project grant No. 2023-02-08-00-12-F04603 and the Science and Technology Committee of Shanghai Housing and Urban Rural Development Management Commission grant No. 2023-005-006.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Data supporting reported results will be made available upon request.

Acknowledgments

Many thanks to our colleagues Mao Mao, Yuduo Guo, Niexing Zhou, Pinzhen Chen and Junwei Zhang for helping during the experiment.

Conflicts of Interest

The authors declare no conflicts of interest.

Abbreviation

Aenvelopebuilding envelope area (m2)
Apenetration rate (m2)
ANOVAanalysis of variance
αLEDabsorption coefficient for lettuce
BCboundary condition
Cconcentration of water vapor
C2inertia resistance constant
CACcultivation area coverage
CEAcontrolled environment agriculture
CFcontainer farm
CFDcomputational fluid dynamics
cpaspecific heat of air (J kg−1 K−1)
DOdiscrete ordinates
δijkronecker delta
ETcrop evaporation rate (g m−2 s−1)
ηLEDLED lighting efficiency (%)
gigravitational acceleration in i dimension (m/s2)
HVACheating, ventilation, and air conditioning
Kheat transfer coefficient (W m−2 K−1)
lmean leaf diameter (m)
LADleaf array density (m−1)
LAIleaf area index (m2/m2)
λfthermal conductivity (W m−1 K−1)
λwlatent heat of water vaporization (J/g)
mcanopymass transfer by canopy (g/s)
mHVACmass gain from HVAC system (g/s)
mmediamass gain from lettuce growing media (g/s)
Mwmolar mass of water (kg/mol)
MCRBLmoisture content range between layers (g/kg)
MTmean temperature (℃)
MMCmean moisture content (g/kg)
μtturbulent viscosity (m2/s)
NNumber of LED light strip
ωawater vapor concentration in ambient air (g/m3)
ωleafsaturated water vapor concentration at leaf temperature (g/m3)
Ppressure (Pa)
PLEDLED electricity power (W)
Pw-asaturated water vapor partial pressure at air temperature (Pa)
Pw-leafsaturated water vapor partial pressure at leaf temperature (Pa)
Prtturbulent Prandtl number
φijvector variable denoting the turbulent mass diffusion flux (kg m−2 s−1)
PPFDphotosynthetic photon flux density (μmol m−2 s−1)
Qcanopyenergy transfer by canopy (W)
Qenvelopeenergy transfer by building envelope (W)
Qequipmentenergy gain from equipment in CF (W)
QHVACenergy gain from HVAC system (W)
Qinfiltrationenergy transfer by infiltration (W)
Qlatlatent heat of canopy (W)
QLEDenergy gain from LED (W)
qsflow rate of supply air (m3/s)
Qsensensible heat of canopy (W)
raaerodynamic boundary layer resistance (s/m)
RLEDradiation absorbed by the canopy (W)
Rmolgas molar constant (J mol−1 K−1)
rsstomatal resistance (s/m)
RBLrange between layers,
the difference of the maximum and minimum values between layers
RHarelative humidity of air (%)
ρaair density (kg/m3)
SDstandard deviation
Tatemperature of air (K)
Tleaftemperature of leaf (K)
Totemperature of outside air (K)
Tstemperature of supply air (K)
TRBLtemperature range between layers (℃)
uair speed (m/s)
uiair speed in i dimension (m/s)

Appendix A

Table A1. Characteristics of all instruments.
Table A1. Characteristics of all instruments.
DeviceTest ProjectRangeAccuracyResolution
Testo 175H1Temperature−20 to +55 °C±0.4 °C
(−20 to +55 °C)
0.1 °C
Humidity0 to 100% RH±2% RH
(2 to 98% RH at 25 °C)
0.1% RH
VelociCalc 9565Velocity0 to 50 m/s±0.015 m/s
(5 to 65 °C)
0.01 m/s
COS03Temperature−40 to +80 °C±0.3 °C (25 °C)0.1 °C
Humidity0 to 100% RH±2% RH
(5 to 95% RH at 25 °C)
0.1% RH

References

  1. Richards, C.E.; Gauch, H.L.; Allwood, J.M. International Risk of Food Insecurity and Mass Mortality in a Runaway Global Warming Scenario. Futures 2023, 150, 103173. [Google Scholar] [CrossRef]
  2. Song, R.; Liu, D.; Pan, Y.; Cheng, Y.; Meng, C. Container Farms: Energy Modeling Considering Crop Growth and Energy-Saving Potential in Different Climates. J. Clean. Prod. 2023, 420, 138353. [Google Scholar] [CrossRef]
  3. Engler, N.; Krarti, M. Optimal Designs for Net Zero Energy Controlled Environment Agriculture Facilities. Energy Build. 2022, 272, 112364. [Google Scholar] [CrossRef]
  4. Panwar, N.L.; Kaushik, S.C.; Kothari, S. Solar Greenhouse an Option for Renewable and Sustainable Farming. Renew. Sustain. Energy Rev. 2011, 15, 3934–3945. [Google Scholar] [CrossRef]
  5. Ahamed, M.S.; Sultan, M.; Monfet, D.; Rahman, M.S.; Zhang, Y.; Zahid, A.; Bilal, M.; Ahsan, T.M.A.; Achour, Y. A Critical Review on Efficient Thermal Environment Controls in Indoor Vertical Farming. J. Clean. Prod. 2023, 425, 138923. [Google Scholar] [CrossRef]
  6. Bu, K.; Yu, Z.; Lai, D.; Bao, H. Energy-Saving Effect Assessment of Various Factors in Container Plant Factories: A Data-Driven Random Forest Approach. Clean. Energy Syst. 2024, 8, 100122. [Google Scholar] [CrossRef]
  7. Xu, D.; Ahmed, H.A.; Tong, Y.; Yang, Q.; van Willigenburg, L.G. Optimal Control as a Tool to Investigate the Profitability of a Chinese Plant Factory—Lettuce Production System. Biosyst. Eng. 2021, 208, 319–332. [Google Scholar] [CrossRef]
  8. Kikuchi, Y.; Kanematsu, Y.; Yoshikawa, N.; Okubo, T.; Takagaki, M. Environmental and Resource Use Analysis of Plant Factories with Energy Technology Options: A Case Study in Japan. J. Clean. Prod. 2018, 186, 703–717. [Google Scholar] [CrossRef]
  9. Sohn, J.; Liulys, M.; Avgoustaki, D.D.; Xydis, G. CFD Analysis of Airflow Uniformity in a Shipping-Container Vertical Farm. Comput. Electron. Agric. 2023, 215, 108363. [Google Scholar] [CrossRef]
  10. Schmidt Rivera, X.; Rodgers, B.; Odanye, T.; Jalil-Vega, F.; Farmer, J. The Role of Aeroponic Container Farms in Sustainable Food Systems—The Environmental Credentials. Sci. Total Environ. 2023, 860, 160420. [Google Scholar] [CrossRef]
  11. Liebman-Pelaez, M.; Kongoletos, J.; Norford, L.K.; Reinhart, C. Validation of a Building Energy Model of a Hydroponic Container Farm and Its Application in Urban Design. Energy Build. 2021, 250, 111192. [Google Scholar] [CrossRef]
  12. Kozai, T. Benefits, Problems and Challenges of Plant Factories with Artificial Lighting (PFALs): A Short Review. Acta Hortic. 2018, 1227, 25–30. [Google Scholar] [CrossRef]
  13. He, L.; Fu, L.; Fang, W.; Sun, X.; Suo, R.; Li, G.; Zhao, G.; Yang, R.; Li, R. IoT-Based Urban Agriculture Container Farm Design and Implementation for Localized Produce Supply. Comput. Electron. Agric. 2022, 203, 107445. [Google Scholar] [CrossRef]
  14. Jo, W.J.; Shin, J.H. Development of a Transpiration Model for Precise Tomato (Solanum Lycopersicum L.) Irrigation Control under Various Environmental Conditions in Greenhouse. Plant Physiol. Biochem. 2021, 162, 388–394. [Google Scholar] [CrossRef] [PubMed]
  15. Yamaura, H.; Kanno, K.; Takano, N.; Isozaki, M.; Iwasaki, Y. Supra-Optimal Daily Mean Temperature Stimulates Plant Growth and Carbohydrate Use in Tomato. Sci. Hortic. 2021, 276, 109780. [Google Scholar] [CrossRef]
  16. Medrano, E.; Lorenzo, P.; Sánchez-Guerrero, M.C.; Montero, J.I. Evaluation and Modelling of Greenhouse Cucumber-Crop Transpiration under High and Low Radiation Conditions. Sci. Hortic. 2005, 105, 163–175. [Google Scholar] [CrossRef]
  17. Li, D.; Dong, J.; Gruda, N.S.; Li, X.; Duan, Z. Elevated Root-Zone Temperature Promotes the Growth and Alleviates the Photosynthetic Acclimation of Cucumber Plants Exposed to Elevated [CO2]. Environ. Exp. Bot. 2022, 194, 104694. [Google Scholar] [CrossRef]
  18. Boo, H.-O.; Heo, B.-G.; Gorinstein, S.; Chon, S.-U. Positive Effects of Temperature and Growth Conditions on Enzymatic and Antioxidant Status in Lettuce Plants. Plant Sci. 2011, 181, 479–484. [Google Scholar] [CrossRef]
  19. Jeong, S.J.; Niu, G.; Zhen, S. Far-Red Light and Temperature Interactively Regulate Plant Growth and Morphology of Lettuce and Basil. Environ. Exp. Bot. 2024, 218, 105589. [Google Scholar] [CrossRef]
  20. Hou, Y.; Li, A.; Li, Y.; Jin, D.; Tian, Y.; Zhang, D.; Wu, D.; Zhang, L.; Lei, W. Analysis of Microclimate Characteristics in Solar Greenhouses under Natural Ventilation. Build. Simul. 2021, 14, 1811–1821. [Google Scholar] [CrossRef]
  21. Kang, L.; van Hooff, T. Numerical Evaluation and Optimization of Air Distribution System in a Small Vertical Farm with Lateral Air Supply. Dev. Built Environ. 2024, 17, 100304. [Google Scholar] [CrossRef]
  22. Perone, C.; Orsino, M.; La Fianza, G.; Giametta, F.; Catalano, P. Study of a Mechanical Ventilation System with Heat Recovery to Control Temperature in a Monitored Agricultural Environment under Summer Conditions. J. Build. Eng. 2021, 43, 102745. [Google Scholar] [CrossRef]
  23. Wang, B.; Li, X.; Xu, M.; Wang, L. Research on Improved Partial Format MFAC Greenhouse Temperature Control Method Based on Low Energy Consumption Optimization. Comput. Electron. Agric. 2024, 220, 108845. [Google Scholar] [CrossRef]
  24. García-Mañas, F.; Hägglund, T.; Guzmán, J.L.; Rodríguez, F.; Berenguel, M. A Practical Solution for Multivariable Control of Temperature and Humidity in Greenhouses. Eur. J. Control 2024, 77, 100967. [Google Scholar] [CrossRef]
  25. Mao, Q.; Li, H.; Ji, C.; Peng, Y.; Li, T. Experimental Study of Ambient Temperature and Humidity Distribution in Large Multi-Span Greenhouse Based on Different Crop Heights and Ventilation Conditions. Appl. Therm. Eng. 2024, 248, 123176. [Google Scholar] [CrossRef]
  26. Lim, T.-G.; Kim, Y.H. Analysis of Airflow Pattern in Plant Factory with Different Inlet and Outlet Locations Using Computational Fluid Dynamics. J. Biosyst. Eng. 2014, 39, 310–317. [Google Scholar] [CrossRef]
  27. Rohdin, P.; Moshfegh, B. Numerical Predictions of Indoor Climate in Large Industrial Premises. A Comparison between Different k–ε Models Supported by Field Measurements. Build. Environ. 2007, 42, 3872–3882. [Google Scholar] [CrossRef]
  28. Kim, K.; Yoon, J.-Y.; Kwon, H.-J.; Han, J.-H.; Eek Son, J.; Nam, S.-W.; Giacomelli, G.A.; Lee, I.-B. 3-D CFD Analysis of Relative Humidity Distribution in Greenhouse with a Fog Cooling System and Refrigerative Dehumidifiers. Biosyst. Eng. 2008, 100, 245–255. [Google Scholar] [CrossRef]
  29. Tamimi, E.; Kacira, M. Analysis of Climate Uniformity in a Naturally Ventilated Greenhouse Equipped with High Pressure Fogging System Using Computational Fluid Dynamics. Acta Hortic. 2013, 1008, 177–183. [Google Scholar] [CrossRef]
  30. Boulard, T.; Roy, J.-C.; Pouillard, J.-B.; Fatnassi, H.; Grisey, A. Modelling of Micrometeorology, Canopy Transpiration and Photosynthesis in a Closed Greenhouse Using Computational Fluid Dynamics. Biosyst. Eng. 2017, 158, 110–133. [Google Scholar] [CrossRef]
  31. Zhang, Y.; Kacira, M. Analysis of Climate Uniformity in Indoor Plant Factory System with Computational Fluid Dynamics (CFD). Biosyst. Eng. 2022, 220, 73–86. [Google Scholar] [CrossRef]
  32. Agati, G.; Franchetti, B.; Rispoli, F.; Venturini, P. Thermo-Fluid Dynamic Analysis of the Air Flow inside an Indoor Vertical Farming System. Appl. Therm. Eng. 2024, 236, 121553. [Google Scholar] [CrossRef]
  33. Frantz, J.M.; Ritchie, G.; Cometti, N.N.; Robinson, J.; Bugbee, B. Exploring the Limits of Crop Productivity: Beyond the Limits of Tipburn in Lettuce. J. Am. Soc. Hortic. Sci. 2004, 129, 331–338. [Google Scholar] [CrossRef]
  34. Graamans, L.; van den Dobbelsteen, A.; Meinen, E.; Stanghellini, C. Plant Factories; Crop Transpiration and Energy Balance. Agric. Syst. 2017, 153, 138–147. [Google Scholar] [CrossRef]
  35. ASABE EP653-2021; Heating, Ventilating, and Air Conditioning (HVAC) for Indoor Plant Environments without Sunlight 2021. ASABE: St. Joseph, MI, USA, 2021.
  36. Yin, Y.; Li, A.; Wu, D.; Li, J.; Guo, J. Low-Resistance Optimization and Secondary Flow Analysis of Elbows via a Combination of Orthogonal Experiment Design and Simple Comparison Design. Build. Environ. 2023, 236, 110263. [Google Scholar] [CrossRef]
  37. Gong, G.; Xu, C.; Jiao, J.; Liu, Y.; Xie, S. Investigation of Moisture Condensation on Papermaking Plant Envelopes in High Humidity Environment by Orthogonal Analysis and CFD Simulation. Build. Environ. 2011, 46, 1639–1648. [Google Scholar] [CrossRef]
  38. Lu, Y.; Wang, T.; Zhao, C.; Zhu, Y.; Jia, X.; Zhang, L.; Shi, F.; Jiang, C. An Efficient Design Method of Indoor Ventilation Parameters for High-Speed Trains Using Improved Proper Orthogonal Decomposition Reconstruction. J. Build. Eng. 2023, 71, 106600. [Google Scholar] [CrossRef]
  39. Yoshioka, M.; Yago, T.; Yoshie-Stark, Y.; Arakawa, H.; Morinaga, T. Effect of High Frequency of Intermittent Light on the Growth and Fatty Acid Profile of Isochrysis galbana. Aquaculture 2012, 338–341, 111–117. [Google Scholar] [CrossRef]
  40. Son, K.-H.; Jeon, Y.-M.; Oh, M.-M. Application of Supplementary White and Pulsed Light-Emitting Diodes to Lettuce Grown in a Plant Factory with Artificial Lighting. Hortic. Environ. Biotechnol. 2016, 57, 560–572. [Google Scholar] [CrossRef]
  41. Olvera-Gonzalez, E.; Escalante-Garcia, N.; Myers, D.; Ampim, P.; Obeng, E.; Alaniz-Lumbreras, D.; Castaño, V. Pulsed LED-Lighting as an Alternative Energy Savings Technique for Vertical Farms and Plant Factories. Energies 2021, 14, 1603. [Google Scholar] [CrossRef]
  42. Li, T.; Lei, J.; Luo, H.; Essah, E.A.; Cheng, Y. Parametric Study and Optimization on Exhaled Particle Dispersion in a Ward Heated by Impinging Jet Ventilation Using Orthogonal-Based Grey Relational Method. J. Build. Eng. 2024, 91, 109619. [Google Scholar] [CrossRef]
Figure 1. Models of the CF. Positioned at the top, bag air ducts (BRISTON, Suzhou, China) supply treated air to the CF, while the air outlet is located at the middle height of the rear wall. To enhance indoor airflow organization, a fan (FUJIWARA, Taizhou, China) is located below the air duct.
Figure 1. Models of the CF. Positioned at the top, bag air ducts (BRISTON, Suzhou, China) supply treated air to the CF, while the air outlet is located at the middle height of the rear wall. To enhance indoor airflow organization, a fan (FUJIWARA, Taizhou, China) is located below the air duct.
Sustainability 16 06105 g001
Figure 2. Independence analysis of grids. Use the temperature distribution on the center line of the canopy plane from left to right as an indicator. (a) The first planting layer. (b) The third planting layer. (c) The fifth planting layer. The grid standards are coarse (1.10 million), general (1.51 million), standard (2.14 million), fine (2.95 million), and precision (3.42 million).
Figure 2. Independence analysis of grids. Use the temperature distribution on the center line of the canopy plane from left to right as an indicator. (a) The first planting layer. (b) The third planting layer. (c) The fifth planting layer. The grid standards are coarse (1.10 million), general (1.51 million), standard (2.14 million), fine (2.95 million), and precision (3.42 million).
Sustainability 16 06105 g002
Figure 3. The distribution of nineteen measuring points in the CF. (a) Viewed from the Z direction. (b) Viewed from the X direction.
Figure 3. The distribution of nineteen measuring points in the CF. (a) Viewed from the Z direction. (b) Viewed from the X direction.
Sustainability 16 06105 g003
Figure 4. The plane is divided into 200 areas, represented by a black-and-white grid.
Figure 4. The plane is divided into 200 areas, represented by a black-and-white grid.
Sustainability 16 06105 g004
Figure 5. Validation. (a) Comparison of temperature at 19 measurement points between simulation and experiment. The box plot represents the experimental data, the red scatter represents the simulated data, and the yellow histogram represents the relative error. (b) Comparison of moisture content at 19 measurement points between simulation and experiment. Histograms represent experimental and simulated data, and red scatters represent differences. The solid black lines represent the range in which instrument error is converted into moisture content (−0.35 g/kg to 0.35 g/kg).
Figure 5. Validation. (a) Comparison of temperature at 19 measurement points between simulation and experiment. The box plot represents the experimental data, the red scatter represents the simulated data, and the yellow histogram represents the relative error. (b) Comparison of moisture content at 19 measurement points between simulation and experiment. Histograms represent experimental and simulated data, and red scatters represent differences. The solid black lines represent the range in which instrument error is converted into moisture content (−0.35 g/kg to 0.35 g/kg).
Sustainability 16 06105 g005
Figure 6. The MT and SD of 200 areas of each layer in case 4. The circle size represents the MT, and the color represents the SD. The horizontal 1–20 is the arrangement in the X direction, and the vertical 1–10 is the Z direction. Vertical 1–5 are the planting area on the left, 6–10 are on the right, and the corridor is omitted.
Figure 6. The MT and SD of 200 areas of each layer in case 4. The circle size represents the MT, and the color represents the SD. The horizontal 1–20 is the arrangement in the X direction, and the vertical 1–10 is the Z direction. Vertical 1–5 are the planting area on the left, 6–10 are on the right, and the corridor is omitted.
Sustainability 16 06105 g006
Figure 7. Velocity vector of L3 in case 4. The four parts circled by red rectangles are the existing vortex areas. The relative size of the rectangle characterizes the strength of the eddy.
Figure 7. Velocity vector of L3 in case 4. The four parts circled by red rectangles are the existing vortex areas. The relative size of the rectangle characterizes the strength of the eddy.
Sustainability 16 06105 g007
Figure 8. The MMC and SD of 200 areas of each layer in case 4. The size of the circle represents the MMC, and the color represents the SD.
Figure 8. The MMC and SD of 200 areas of each layer in case 4. The size of the circle represents the MMC, and the color represents the SD.
Sustainability 16 06105 g008
Figure 9. The MT and MMC of each layer. The number 1–16 is the case number, and the X-axis represents the layers. The left Y-axis is the MT, and the right is the MMC. The direction of the square brackets indicates where the RBL occurs, and the number below is the value of the RBL.
Figure 9. The MT and MMC of each layer. The number 1–16 is the case number, and the X-axis represents the layers. The left Y-axis is the MT, and the right is the MMC. The direction of the square brackets indicates where the RBL occurs, and the number below is the value of the RBL.
Sustainability 16 06105 g009
Figure 10. Trend chart of factors and evaluation index.
Figure 10. Trend chart of factors and evaluation index.
Sustainability 16 06105 g010
Table 1. Factors and levels.
Table 1. Factors and levels.
FactorLevel
1234
ATemperature of supply air, Ts (°C)16182022
BFlow rate of supply air, qs (m3/h)2000225025002750
CRelative humidity of supply air, RHs (%)50607080
DLED duty ratio, (-)0.40.60.81
Table 2. Orthogonal design table.
Table 2. Orthogonal design table.
NoFactor
A (°C)B (m3/h)C (%)D (-)Error
1162000500.41
2162250600.62
3162500700.83
41627508014
5182000600.84
61822505013
7182500800.42
8182750700.61
92020007012
10202250800.81
11202500500.64
12202750600.43
13222000800.63
14222250700.44
152225006011
16222750500.82
Table 3. Settings of boundary conditions.
Table 3. Settings of boundary conditions.
BCsTypeSetting
InletVelocity–inletFrom case
Bag PipePorous–jumpFace Permeability: 2.81 × 10−12
Pressure–Jump Coefficient: 8.58 × 106
FanFanPressure Jump: polynomial
velocity: from 8.27 m/s to 10.63 m/s
FrameWall–Heat FluxNone
LEDWall–Heat FluxFrom case
PlantFluid–Porous Zone and Source TermsUDF
OutletPressure–outlet/
Wall
(except Floor)
Wall–ConvectionHeat Transfer Coefficient: 0.7
Free Stream Temperature: 296.01
(Chinese Standard Weather Data)
Wall Thickness: 0.05
Wall-FloorWall–Heat FluxNone
Table 4. Range analysis table.
Table 4. Range analysis table.
Evaluation IndexRA *RBRCRDOrder of Importance
MT of L15.7491.3190.6714.211A > D > B > C
MT of L25.723 1.315 0.703 4.065 A > D > B > C
MT of L35.889 1.562 0.723 3.993 A > D > B > C
MT of L45.827 1.501 0.714 4.032 A > D > B > C
MT of L55.639 0.561 0.610 3.183 A > D > B > C
TRBL0.119 0.758 0.236 1.019 D > B > C > A
MMC of L13.795 1.015 3.668 0.721 A > C > B > D
MMC of L23.800 0.960 3.663 0.695 A > C > B > D
MMC of L33.808 1.040 3.636 0.739 A > C > B > D
MMC of L43.779 0.984 3.679 0.712 A > C > B > D
MMC of L53.671 0.717 3.789 0.518 A > C > B > D
MCRBL0.119 0.290 0.142 0.289 B > D > C > A
* Ri is the range value of factor i.
Table 5. ANOVA.
Table 5. ANOVA.
IndexFactorsSS *Df *MS *FpSignificant
MT of L1A73.839324.613181.8050.001**
B3.72831.2439.1790.051
C1.19230.3972.9340.200
D44.225314.742108.8900.001**
error0.40630.135
MMC of L1A32.210310.737129.1060.001**
B2.32830.7769.3320.050
C30.105310.035120.6700.001**
D1.43230.4775.7390.093
error0.24930.083
TRBLA0.02830.0090.4930.712
B1.20230.40120.8680.016*
C0.11430.0381.9870.294
D2.66430.88846.2320.005**
error0.05830.019
MCRBLA0.03030.0105.7720.092
B0.17630.05934.3460.008**
C0.04630.0158.9670.052
D0.20530.06839.9590.006**
error0.00530.002
* SS is sum of squares deviations from the mean, Df is degrees of freedom, MS is mean square. “**” indicates a highly significant impact while “*” indicates a significant effect.
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

Nie, Z.; Liu, D.; Meng, C.; Song, R. Research on Thermal Environment of Container Farms: Key Factor Identification and Priority Analysis. Sustainability 2024, 16, 6105. https://doi.org/10.3390/su16146105

AMA Style

Nie Z, Liu D, Meng C, Song R. Research on Thermal Environment of Container Farms: Key Factor Identification and Priority Analysis. Sustainability. 2024; 16(14):6105. https://doi.org/10.3390/su16146105

Chicago/Turabian Style

Nie, Zihao, Dong Liu, Chuang Meng, and Ruizhi Song. 2024. "Research on Thermal Environment of Container Farms: Key Factor Identification and Priority Analysis" Sustainability 16, no. 14: 6105. https://doi.org/10.3390/su16146105

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