Next Article in Journal
From Waste Biomass to Cellulosic Ethanol by Separate Hydrolysis and Fermentation (SHF) with Trichoderma viride
Previous Article in Journal
The Effect of Soil Surface Mounds and Depressions on Runoff
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Experimental and Simulation Research on the Energy-Saving Potential of a Sunspace—Taking an Apartment in Qingdao as an Example

1
iSMART, Qingdao University of Technology, Qingdao 266033, China
2
College of Architecture and Urban Planning, Qingdao University of Technology, Qingdao 266033, China
3
Department of Architecture, The University of Kitakyushu, Kitakyushu 808-0135, Japan
4
School of Environmental and Municipal Engineering, Jilin Jianzhu University, Changchun 130118, China
*
Authors to whom correspondence should be addressed.
Sustainability 2023, 15(1), 176; https://doi.org/10.3390/su15010176
Submission received: 23 November 2022 / Revised: 18 December 2022 / Accepted: 19 December 2022 / Published: 22 December 2022

Abstract

:
This paper investigates the thermal performance and energy-saving potential of the sunspace of an old apartment building in Qingdao through experiments and software simulations. This study found that different modes of user behavior can have a substantial impact on the thermal performance of the studied sunspace in winter. The bedroom with a non-ventilated sunspace showed a higher average temperature than the bedroom without the sunspace. However, the bedroom with the sunspace had more heat loss than the bedroom without the sunspace when the sunspace was naturally ventilated, especially at night. In Delta temperature (DT)-controlled ventilation mode, the heating load of a bedroom can be reduced by 2.94 kWh/m2 compared with the non-ventilated mode. Simply optimizing the roof configuration of the sunspace can significantly improve the heat gain of the non-ventilated sunspace and reduce the energy consumption for heating by 26 kWh/m2. Compared with the unoptimized sunspace, the optimized ventilated sunspace can reduce the heating load of the bedroom by 38.82 kWh/m2. In addition, the overheating of the room in summer can be solved by opening the exterior windows of the sunspace for ventilation during the day and leaving the door of the sunspace open at night.

1. Introduction

Since the 1970s, the energy crisis and environmental pollution have attracted significant attention worldwide. Energy consumption in buildings is an important part of the final energy consumption, accounting for 20% to 40% of direct energy consumption [1]. According to the Ministry of Energy, the problem of inefficient energy use in China is more severe compared to many developed countries such as America, Japan, and many countries in Europe. The 2018 China Building Energy Consumption Report states that the amount of energy consumed during building construction and operation in a year can occupy about 20.62% of the total annual energy consumption. The construction and operation of urban residential buildings consume significantly more energy than that of rural residential buildings. In 2020, China proposed a dual carbon target, which aims to hit peak carbon emissions by 2030 and achieve carbon neutrality by 2060.
The use of passive solar technologies in buildings can help reduce the energy consumption for heating in buildings. Common passive solar technologies include sunspaces, Trombe walls, and double-skin façades. Sunspaces consist of three main components: the exterior glazing, the intermediate cavity, and the interior wall or glazing. Among them, the cavity thickness of a sunspace, which is equivalent to the depth of the sunspace, is the largest [2]. Functionally, a sunspace can increase the floor area and be used as a place that allows people to breathe fresh air, hang clothes for drying, and grow plants. Therefore, sunspaces are often used in urban residential buildings. A sunspace can also separate the indoor and outdoor environments and act as a buffer to reduce heat loss from buildings [3]. The glazing performance of a sunspace plays an important role in reducing indoor energy consumption. It can allow useful visible light and short-wave radiation to penetrate into a room and block long-wave radiation emission. This can further increase the indoor air and wall-surface temperatures. In this way, the internal cavity of the sunspace can be transformed into an effective heat collector, reducing the thermal difference between the indoor and outdoor environments. The stored heat can be further transferred into the neighboring rooms through air-forced convection or by conduction through the interior walls [4].
Some studies have indicated that a sunspace can reduce 8% to 58% of a building’s energy consumption [5,6,7,8]. Hilliaho et al. [9] found that the maximum temperature difference between the outdoor air and the sunspace can reach 14 °C. Moreover, many parameters can influence the thermal performance of a sunspace. These parameters include glazing area [10,11], location [12,13,14,15], glazing type [16,17,18], form [19,20,21], shading devices [22], ventilation [23,24], thermal mass [25,26,27,28,29,30], and doors and windows [31,32]. Liu et al. [33] evaluated the sunspace temperature when the sunspace was set with three depths (150 cm, 120 cm, and 90 cm). They found that the highest indoor temperature was gained when the depth of the sunspace was 150 cm. Ulpiani et al. [34] evaluated the thermal performance of a sunspace when it was set to four depths (250 cm, 220 cm, 200 cm, and 150 cm). They also concluded that the thermal performance of the sunspace increased with decreasing depth. Babaee et al. [35] explored the thermal load of a room with different glazing types (single, double glazing, and double low-E glazing) and found that the sunspace with double low-E glazing had the best energy performance when the optimal window-to-ground ratio was 2.21. This is because low-E glass has a selective surface that permits short-wave radiation transmission to penetrate into the interior [36] and reduces heat loss transmission to the external environment [37]. Oliveti et al. [38] reached a similar conclusion. It was found that a ventilated sunspace has better energy-saving performance. Ulpiani et al. [34] found that the energy efficiency of a studied sunspace can be increased by more than 27% by using mechanically controlled ventilation. Zhu et al. [39] found that convective heat transfer from a sunspace to its adjacent rooms can reduce the energy consumption of the adjacent rooms by 28%. Ma et al. [40] used EnergyPlus to simulate the energy-saving performance of a ventilated sunspace. The results showed that compared with the bedroom without a sunspace, a bedroom with a ventilated sunspace reduced the energy consumption for heating by 14.7%. At the same time, compared with the bedroom with a non-ventilated sunspace, the bedroom with a ventilated sunspace can also reduce the energy consumption for heating by 4.4%. Similarly, Aelenei et al. [31] also used EnergyPlus software to simulate the energy performance of a sunspace in Portugal. The results showed that using a sunspace reduced the energy consumption for heating by 48% in winter, but it also increased the cooling energy consumption by 10% in summer. However, the increase in cooling energy consumption can be avoided by installing an adjustable shading system or a ventilation system. According to Chiesa et al. [41], shading and ventilation systems are two important techniques to reduce the cooling demand, and shading is more effective than ventilation. When these two techniques are combined, the overheating problem caused by a sunspace can be effectively solved. Batinah and Fayez [42] explored the optimal design of a sunspace in different climate zones in Spain in summer. They also concluded that an optimized sunspace should have an external shading system and proper ventilation.
In general, current studies indicate that there are numerous studies that focus on the thermal performance of non-ventilated sunspaces. In comparison, studies about the optimization of a ventilated sunspace are rather limited. Most studies investigated the energy-saving potential of sunspaces in detached houses, while very few studies analyzed the energy-saving potential of sunspaces in multi-story residential apartments. In this study, an old six-story apartment in Qingdao, which is located in the cold climate zone of China, was selected for a case study. The novelty of this paper is to propose an approach for evaluating the sunspace performance that combines parametric design and performance evaluation. A simplified interzone airflow model ‘ZoneCrossMixing’ was used to simulate the air exchange between the bedroom and the sunspace. The impact of different modes of user behavior and physical configuration on the thermal performance of a sunspace was analyzed and investigated through actual measurements and software simulations. The purpose of this study is not only to propose an optimal sunspace solution but also to provide various feasible optimization options through a large number of simulated cases. It is expected to make some suggestions for residents to use the sunspace rationally in different seasons and to provide various possible optimization solutions for designers to improve residential thermal comfort and energy-saving efficiency. Figure 1 shows the flow chart of the paper structure.

2. Methodology

2.1. Experimental Study

In this paper, an old apartment in Qingdao (elevation: 40 m, longitude: 120.4° E, latitude: 36.1° N), which is located in the cold climate zone of China, was selected for a case study. Qingdao is cold and windy in winter and hot and humid in summer. It is in the transition zone between the humid subtropical climate and the humid continental climate. The appearance of the sunspace is shown in Figure 2. The apartment has two bedrooms, one of which is a bedroom without a sunspace, denoted as bedroom 0. The other one is a bedroom with a sunspace, denoted as Bedroom 1 (Figure 3). The sunspace has four large, south-facing windows and an opaque, flat roof. The apartment is heated by radiators through a district heating system (running 24 h) in winter and cooled by an air conditioner in summer.
The Vantage Pro2 weather station (Davis, CA, USA) was used to measure solar radiation and air temperature. The TC96-V DT controller (Rockson, Shanghai, China) was used to control the air supply between the sunspace and bedroom. The testo-174H thermometers (Testo, Berlin, Germany) were used to measure the temperature of the sunspace and bedroom. The testo-175T3 surface thermometers (Testo, Germany) were used to measure the surface temperature of the radiators. The room plan and the position of the thermometers are shown in Figure 4. Among them, thermometers No. 5 and No. 6 were placed in the position of the air inlet and outlet, respectively. Table 1 shows the uncertainties of these devices in this experiment.
In order to study the thermal performance of a sunspace under different modes of user behavior in winter, the thermal performance of the sunspace in DT-controlled ventilation mode, ‘Door-closed’ mode (non-ventilated sunspace), and ‘Door-open’ mode (naturally ventilated sunspace) were measured in December (heating) and April (HVAC system was shut down). All measurements were obtained over a 24 h period. The Delta-temperature-controlled ventilation system was composed of an intelligent DT controller and two fans. The DT controller was externally connected to two air-temperature sensors, which were used to detect the temperature difference between the sunspace and Bedroom 1. During the actual measurement, the EPS board was used as the door. The air supply fan and the return air fan were installed at a distance of 15 cm from the upper and lower edges of the EPS board. When the temperature difference reached the target temperature, the air in the sunspace was sent to the bedroom by the supply fan, and the air in the bedroom was sent to the sunspace by the return fan. Figure 4 shows the indoor condition of Bedroom 1, which was controlled by the Delta temperature-controlled ventilation system. DT (Delta temperature) is equal to the supply (inlet) air temperature minus the return (outlet) air temperature. The initial temperature difference was set to 3 °C, the final temperature difference was set to 1 °C, and the ventilation volume of the fan was set to 78 m³/h. When the DT was greater than 3 °C, the ventilation equipment started working. When DT was less than or equal to 1 °C, the ventilation equipment stopped working. When the ‘Door-closed’ mode was enabled, the door between the sunspace and Bedroom 1 was closed. When the ‘Door-open’ mode was enabled, the door between the sunspace and Bedroom 1 was kept open allowing the air to ventilate between the sunspace and Bedroom 1. In addition, actual measurements were conducted to address the potential overheating of the sunspace in summer (the HVAC system was shut down). Three modes of user behavior were recorded: ‘Door-closed-window-closed (DCWC)’, ‘Door-open-window-closed (DOWC)’, and ‘Door-closed-window-open (DCWO)’. DCWC means that the door between the sunspace and bedroom is closed, and the window of the sunspace is closed. DOWC means that the door between the sunspace and bedroom is opened, and the window of the sunspace is closed. DCWO means that the door between the sunspace and bedroom is closed, and the window of the sunspace is opened. Table 2 shows the monitoring schedule.

2.2. Simulation Study

The energy performance of the studied sunspace was evaluated by using the proposed method: the parametric performance design. This method combines parametric design and performance evaluation. It is an ‘all-in-one’ toolbox for non-professionals to assess and optimize the energy-saving performance of a sunspace. Developed from the Ladybug plugin, the simulation method is operated on the EnergyPlus engine in the grasshopper platform of Rhino. A parametric model of the study room was built and simulated by using Ladybug Tools. A notable feature of EnergyPlus is that heat and mass transfer models can be integrated to measure the air movement between different zones. Compared to the Airflow network model, ‘ZoneCrossMixing’ is a simplified interzone airflow model in the software. It is ideal to simulate the airflow exchange between two zones. Therefore, the ‘ZoneCrossMixing’ model was applied to simulate the air exchange between the bedroom and the sunspace instead of the airflow network model. In the ‘ZoneCrossMixing’ model, Delta temperature was the key parameter that was used to control the airflow from the sunspace to Bedroom 1. Because Delta temperature cannot be a negative value, there would be two conditions. If the Delta temperature was positive, it means that the sunspace temperature was higher than the bedroom temperature, and the warm air was being drawn from the sunspace to the bedroom. This can lead to a mixing of the warm and cold air in the bedroom. If the Delta temperature was zero, the warm and cold air was still mixed regardless of the relative temperatures of the two rooms.
During the heating period, the heating system (running 24 h) was used to maintain the bedroom temperature at 18 °C. Table 3, Table 4, Table 5 and Table 6 show the structural parameters of the sunspace as a parametric simulation model. The typical meteorological year (TMY) data of Qingdao (Qingdao_TMYx.2007–2021) was used for the simulation. The basic climatic parameters including global horizontal radiation, dry bulb temperature, dew point temperature, relative humidity, wind direction, and wind speed are shown in Table 7. The simplified dynamic parametric simulation model is shown in Figure 5. Using this model, heat collection and energy savings were simulated to assess the different configurations in various parameter settings. The optimized model was also coupled with the DT-controlled ventilation device to compare and evaluate the energy-saving potential of the optimized ventilated sunspace.
Table 8 shows the parameters used in the simulation programs. There are eight parameters, including θ (slope angle of the roof), WRR (window to roof ratio), RGT (roof glazing type), FGT (façade glazing type), WWR (window to wall ratio), depth (width of the sunspace), FR (flow rate), and DT (delta temperature). The input range of θ is between 0° to 60°, with a total of seven values at intervals of 10°. The range of WRR is from 0% to 90% with a total of nine values, at an interval of 10%. RGT and FGT both have two types of settings: single glazing and double low-E glazing. The input range of the WWR is from 10% to 90% with a total of nine values at intervals of 10%. The input range of depth is from 0.6 m to 1.5 m, with a total of ten values at intervals of 0.1 m. The input range of FR is from 0.00 m3/s to 0.06 m3/s, with a total of seven values at intervals of 0.01 m3/s. The input range of DT is between 1 and 6 °C, with a total of seven values at intervals of 1 °C. A total of 1,058,400 configurations were calculated to explore and evaluate the energy-saving potential of the sunspace in different physical forms. Table 9 shows the characteristics of the glazing materials of the sunspace.

3. Results

3.1. Results of the Experimental Study

Figure 6 shows the relationship between room temperature and solar radiation variations in the ‘Door-closed’ mode during the heating period. It can be seen that the overall temperature of Bedroom 1 (with the sunspace) was higher than that of Bedroom 0 (without the sunspace). The maximum temperature difference reached 1.9 °C. The average temperature of Bedroom 1 was 23.8 °C, which was 1.6 °C higher than the average temperature of Bedroom 0. On 26 December, there was a significant increase in solar radiation. The room temperature of Bedroom 0 reached its peak value in the late afternoon with a maximum temperature of 30.4 °C, creating an uncomfortable indoor environment. In contrast, Bedroom 1 had fewer temperature fluctuations and no overheating in the afternoon.
Figure 7 shows the relationship between temperature and solar radiation variations in the ‘Door-closed’ mode during the non-heating period. The overall temperature of Bedroom 1 remained generally higher than that of Bedroom 0, and the average temperature of Bedroom 1 was 0.8 °C higher than that of Bedroom 0. Unlike the heating period, the temperature difference between Bedroom 1 and Bedroom 0 was smaller during the afternoon hours of the day, and the temperature difference was even more pronounced at night. This indicates that the ‘Door-closed’ mode can reduce the bedroom’s heat loss during the night. In addition, compared with the average bedroom temperature during the heating period, the average temperature of the two bedrooms during the non-heating period decreased. The overall temperature fluctuation of the two bedrooms was small, and the overheating of the two bedrooms disappeared in the afternoon.
In summary, the thermal performance of the studied sunspace was very efficient in winter when the ‘Door-closed’ mode was enabled. The sunspace can reduce the bedroom’s heat loss at night and improve the room’s overall thermal stability. In addition, Bedroom 1 can be overheated because it received a lot of direct solar heat in the afternoon. Using radiators for heating can further lead to overheating in winter. This means that Bedroom 1 can better withstand the effects of the outdoor environment and increase the indoor thermal stability.
Figure 8 shows the relationship between the temperature and solar radiation variations in the ‘Door-open’ mode during the heating period. Compared with the bedroom under the ‘Door-closed’ mode, the temperature of Bedroom 1 in the ‘Door-open’ mode fluctuated more during the day. However, the overall thermal stability of Bedroom 1 was still better than that of Bedroom 0. However, the average temperature of Bedroom 1 was 1 °C lower than that of Bedroom 0. This is especially true at night, as Bedroom 1 saw more heat loss than Bedroom 0.
Figure 9 shows the relationship between the temperature and solar radiation variations in the ‘Door-open’ mode during the non-heating period. The average temperature of Bedroom 1 was 0.98 °C higher than that of Bedroom 0, and the maximum temperature difference between the two rooms reached 1.38 °C. However, because the outdoor temperature and solar radiation dropped dramatically from 12 April to 13 April, the additional heat gain created by the sunspace also decreased and eventually disappeared. This phenomenon was observed as the temperature of Bedroom 1 became the same as that of Bedroom 0. This indicates that the ‘Door-open’ mode of the sunspace can reduce the thermal stability of the adjacent bedroom. When the outdoor temperature becomes warmer, the ‘Door-open’ mode can preserve the heat. However, the bedroom suffered significant heat loss in the ‘Door-open’ mode at night when the outdoor temperature dropped drastically.
Figure 10 shows the relationship between the temperature and solar radiation variations in the DT-controlled ventilation mode during the heating period. The DT-controlled ventilation mode was applied from 18 to 21 December, and the ‘Door-closed’ mode was enabled from 23 to 26 December. Due to the lack of solar radiation and cold weather outside, the sunspace temperature was lower than Bedroom 1′s temperature throughout the day (Figure 6), and, therefore, DT-controlled ventilation was stopped. Furthermore, the average temperature of Bedroom 1 was above 20 °C during the heating period. It was not suitable to use the DT-controlled ventilation mode in the sunspace under the current heating conditions because of the lack of proper temperature difference conditions and thermal comfort requirements.
Figure 11 shows a schematic diagram of the effective working zones in DT-controlled ventilation mode during the non-heating period. Due to the significant temperature difference between Bedroom 1 and the sunspace during the non-heating period, the DT-controlled ventilation mode produced an effective working zone. In the effective working zone, the temperature of Bedroom 1 was slightly increased, proving that the DT-controlled ventilation had some positive heating effect on Bedroom 1.
In order to understand and avoid the indoor overheating that may be caused by the sunspace in summer, three modes of use of the sunspace were tested: the ‘Door-closed-window-closed’ mode, ‘Door-closed-window-open’ mode, and ‘Door-open-window-closed’ mode. Among them, Figure 12 shows the relationship between the indoor and outdoor temperatures and the solar radiation variations in the summer when the sunspace was in the ‘Door-closed-window-closed’ mode. As shown in the figure, the overall temperature of Bedroom 1 was significantly higher than that of Bedroom 0. The average daily temperature difference between Bedroom 1 and Bedroom 0 was 0.4 °C, and the maximum temperature difference reached 0.6 °C. Therefore, Bedroom 1 was overheated in ‘Door-closed-window-closed’ mode in summer, but the degree of overheating was mild.
Figure 13 shows the relationship between the temperature and solar radiation variations of the sunspace in the ‘Door-closed-window-open’ mode in summer. Compared with the two rooms in the ‘Door-closed-window-closed’ mode, the average temperature difference between Bedroom 1 and Bedroom 0 decreased and was only 0.27 °C. At noon and in the afternoon, the temperature of Bedroom 1 was similar to Bedroom 0. Therefore, in ‘Door-closed-window-open’ mode, opening the external windows of the sunspace for ventilation during the daytime can minimize the heat storage effect and further mitigate the overheating problem of the adjacent bedroom.
Figure 14 shows the relationship between temperature and solar radiation variations in the summer in the ‘Door-open-window-closed’ mode, and the data shows a 0.02 °C positive temperature difference between Bedroom 1 and Bedroom 0. The temperature of Bedroom 1 was lower or equal to Bedroom 0, especially at night. This suggests that opening the door of the sunspace at night can effectively mitigate the overheating of the adjacent bedroom. Therefore, overheating in the summer caused by the sunspace can basically be avoided by opening the external windows of the sunspace for ventilation during the daytime and leaving the door between the sunspace and the bedroom open at night. Bataineh [47] also had a similar conclusion in their study of a sunspace in Amman-Jordan.

3.2. Results of Simulation Study

3.2.1. Model Validation

This study used ASHRAE Guide 14-2014 [48] to calibrate the simulation model. Two dimensionless metrics for determining the error were applied: the MBE (mean bias error) and the CV(RMSE) (coefficient of variation of the root mean square error). Simplified MBE formulas and CV(RMSE) formulas are shown in Equations (1) and (2).
M B E = Σ i = 1 n ( M i S i ) Σ i = 1 n M i × 100 %
C V ( R M S E ) = 1 y ¯ Σ i = 1 n ( M i S i ) 2 n × 100 %
According to ASHRAE guidelines, the simulation model should be calibrated if the hourly MBE value is less than ±10% and the hourly CV (RMSE) value is less than 30%. Figure 15 shows the comparison of simulated and measured temperatures. The results indicate that the simulation data generally match the measured data. The error indicators for each room are shown in Table 10. This indicates that the simulation model is able to generate relatively valid data that match reality.

3.2.2. Efficiency Analysis

The sunspace roof was first optimized by using a software simulation. Figure 16 shows the results of different combinations of WRR (window-to-roof ratio) and θ (roof tilt angle) configurations. Figure 16a shows the average temperature of the sunspace without heating in winter; Figure 16b shows the amount of energy used for heating Bedroom 1.
The results show that when the roof was flat (θ = 0°) and the WRR increased from 0% to 90%, the average temperature of the sunspace decreased by 0.1 °C, and the heating load decreased by 1 kWh/m². This may be because the increasing WRR can increase the incident solar radiant heat and the heat conduction in the sunspace. While increasing the size of the skylight, more solar radiant heat entered Bedroom 1 through the inner wall glazing, and therefore the energy consumption decreased. However, in general, there was no additional heat gain if the roof of the sunspace was a skylight with single glazing. When the roof was tilted at certain angles, increasing the WRR increased the heat gain of the sunspace. When the WRR was greater than or equal to 30%, the larger the roof tilt angle, the more heat gain the sunspace can attain. Bastien [49] also obtained a similar result. Therefore, a tilted roof was suitable for a larger WRR. Because the height of the sunspace should be more than 1.8 m, the maximum roof tilt angle was 40°. When the WRR was 90% and the roof tilt angle was increased from 0° to 40°, the heating load was significantly reduced by 4 kWh/m2.
Figure 17 shows the simulation results of different combinations of the WRR (window-to-roof ratio) and θ (roof tilt angle) of the roof configurations when the roof glazing type was changed to double low-E glazing. Among them, Figure 17a shows the average temperature of the sunspace without heating, and Figure 17b shows the energy consumption for heating in Bedroom 1. Better thermal performance of the sunspace can be achieved by optimizing the roof material. More heat gain can be achieved when the roof glazing type is changed to double low-E glazing. The results show that increasing the WRR of a flat roof can significantly increase the heat gain. Similar to single glazing, increasing WRR or θ can also improve the thermal performance and heat gain of the sunspace. Moreover, the overall thermal performance of the sunspace was substantially improved compared to single glazing. When the WRR was set to 90% and θ was 40° and the double low-E glazing was applied, the heating load of the bedrooms was reduced by 21.11 kWh/m2 and the average temperature of the sunspace was increased by 6 °C. The simulation indicates that the best sunspace roof configuration was the double low-E glazing with a 40° tilt angle and 90% WRR.
The impact of different façade types and depths on the thermal performance of the sunspace was further investigated by using the software simulation. Figure 18 shows the results of different combinations of depth and WWR (window-to-wall ratio) of the sunspace. Figure 18a shows the average temperature of the sunspace without a heating system; Figure 18b shows the amount of energy used for heating Bedroom 1. The results show that the sunspace’s heat gain was positively correlated with WWR and negatively correlated with depth when the roof was an opaque, flat roof. Bataineh [47] also had a similar conclusion. However, as the WWR increased, the additional heat gain rate kept decreasing when the depth remained unchanged. In particular, when the WWR was greater than 60%, the increasing rate of heat gain decreased as the WWR increased. Furthermore, increasing the depth can lead to inefficient thermal performance. This might be due to the extra shadow casting created by the untransparent roof cover.
To eliminate the negative impact of the opaque roof shading on the sunspace, the roof configuration was set to the optimized roof parameters: θ was set to 40°, WRR was 90%, and RGT was double low-E glazing. Then, the thermal performance of the sunspace was simulated when different depths and WWRs (window-to-wall ratios) of the sunspace were combined. The results are shown in Figure 19. It can be seen that the heat gain of the sunspace was negatively correlated to WWR, and the heat gain of the sunspace was positively correlated to depth. This was the opposite of the previous result of the opaque roof. Increasing the depth can increase the heat gain of the sunspace. This is largely because increasing the depth allows more solar radiation to penetrate the roof and enter the sunspace. Furthermore, increasing the WWR can decrease the thermal performance of the sunspace when the depth remains unchanged. This is because the large window opening area can increase the thermal conductivity of the sunspace, resulting in weak thermal insulation. When the roof tilt angle θ was 40°, the optimal depth of the sunspace was 1.1 m. Moreover, this study assumed that the minimum WWR should be no less than 50%, as a small WWR can affect indoor daylight. Therefore, in the simulation, when the sunspace depth was reduced from 1.2 m to 1.1 m and the WWR was reduced from 70% to 50%, the heating load of the bedroom was reduced by 0.85 kWh/m2 and the average temperature of the sunspace was increased by 0.5 °C.
The thermal performance of the sunspace can be further optimized by changing the façade glazing type from single glazing to double low-E glazing. In Babaee’s study [35], they also obtained a similar result. Figure 20 shows the thermal performance of the sunspace configurations with different combinations of depth and WWR (window-to-wall ratio). The data show that the range of WWR values decides whether increasing depth has a positive or negative impact on the sunspace. When the WWR was less than 50%, the heat collection capacity was positively correlated to the depth. This is because increasing depth can increase the glazing area of the façade, and this means that the sunspace can receive more solar radiation and heat. When the WWR was 50%, the effect of the depth on the sunspace was negligible. When the WWR was greater than 50%, the thermal performance was negatively correlated to depth. This result might be generated because of the excessive window-opening area of the façade, which increased the thermal conductivity of the sunspace. In addition, the high-performance façade glazing greatly increased the heat collection and insulation of the sunspace. When the WWR was set to 50% and the depth was 1.1 m, changing the façade glazing type from single glazing to double low-E glazing reduced the heating load by 11.92 kWh/m² and increased the average temperature of the sunspace by 9.64 °C. By optimizing the structural configuration of the sunspace, the thermal performance of the sunspace was significantly improved. The average temperature of the optimized sunspace reached 25.38 °C in winter when double low-E glazing was used.
Figure 21a shows the average temperature of Bedroom 1 without heating by using different combinations of FR (flow rate) and DT (delta temperature). Figure 21b shows the amount of energy consumed for heating Bedroom 1 by using different combinations of FR and DT. The results show that the larger the FR, the more energy was consumed to heat the bedroom when DT was equal to 0° and DT-controlled ventilation mode was enabled. When DT was larger than 0°, the energy consumption decreased significantly, and more specifically, it decreased as the FR increased. Moreover, the energy consumption for heating was the lowest when DT was set to 1 °C, resulting in the optimal 0.03 m3/s air flow rate. Under these parameter settings, using the DT-controlled ventilation mode reduced the heating load by 0.58 kWh/m2 compared to the ‘Door-closed’ mode. Compared with the unoptimized conditions, the heating load was reduced by 38.82 kWh/m2.
The optimized sunspace shows a significant improvement in energy performance and saving. Table 11 shows some optimized sunspace configurations and their energy savings. These results can provide data support for the renovation and design of the sunspace. Conf_0 represents the unoptimized model. By optimizing the configuration of the sunspace and ventilation mode, the energy used for heating the bedroom can be reduced by up to 99%.

4. Discussion

This paper explores the energy-saving potential of a sunspace in an old residential apartment by using six modes of user behavior. The thermal performance of the sunspace in winter was investigated by enabling the ‘Door-closed’ mode, ‘Door-open’ mode, and DT-controlled ventilation mode, and the sunspace overheating in summer was also studied by enabling the ‘Door-closed-window-closed’ mode, ‘Door-closed-window-open’ mode, and ‘Door-open-window-closed’ mode. By using the parametric performance design method, 1,058,400 possible configuration combinations were simulated. Among them, 468 options were selected to evaluate the energy-saving effect of the sunspace in the six ventilation modes.
The thermal performance of a sunspace with different modes of user behavior in winter was different. During the heating period, the average temperature of the studied bedroom with the sunspace in ‘Door-closed’ mode was 1.6 °C higher than the bedroom without the sunspace. In the non-heating period, the temperature difference between the two conditions was only 0.8 °C. In ‘Door-open’ mode, the average temperature of the bedroom with the sunspace was 1 °C lower than the bedroom without the sunspace during the heating period, and the temperature difference between the two was 0.98 °C in the non-heating period. This means that it is not suitable to use the DT-controlled ventilation mode when district heating is applied in Qingdao. Therefore, it is necessary to integrate the HVAC system with the DT-controlled ventilated sunspace in a more flexible way. For example, in order to achieve more energy-efficient performance, the HVAC system can be set to ‘Low’ when using the DT-controlled ventilated sunspace.
From this study, two major suggestions are provided for further studies. First and foremost, a series of parametric modeling tools in modular form should be further explored from engineering and economic perspectives. These tools can be upgraded with deep reinforcement learning functions to automatically gather the best combinations of various parameters of a subject. Moreover, due to the recent research trend being more focused on the development of multifunctional sunspaces, this research recommends inventing new types of multifunctional sunspaces that can integrate sunspaces with other ‘green’ functions at the design stage, such as humidification and dehumidification, air purification, oxygen supply, carbon dioxide removal (photosynthesis), power generation, natural lighting, and sound insulation.

5. Conclusions

The differing thermal performance of a sunspace in various modes of user behavior was tested in this study, and the energy-saving potential of the ventilated sunspace was further evaluated and optimized by using a proposed parametric performance design method. By optimizing the configuration of the sunspace and ventilation mode, the energy consumed for heating in the studied Bedroom 1 can be reduced by up to 99%.
The results of DT-controlled ventilation show that the energy used for heating Bedroom 1 was the lowest when the Delta temperature was set to 1 °C. In this setting, the larger the flow rate, the more energy that can be saved. Applying the DT-controlled ventilation mode reduced the heating load by 0.58 kWh/m2 compared to the ‘Door-closed’ mode. Compared with its unoptimized condition, the heating load was reduced by 38.82 kWh/m2. Roof configuration has a great impact on the thermal performance of a sunspace. Simply optimizing the roof configuration of a non-ventilated sunspace can significantly improve its heat gain and reduce the energy consumed for heating the bedroom by 26 kWh/m2.
Increasing the θ or WRR of the sunspace can also improve the thermal performance and heat gain of the sunspace. Glazing material also shows a great impact on the thermal performance of a sunspace. The best sunspace roof configuration was the double low-E glazing with a 40° tilt angle and 90% WRR. When using the double low-E glazing on the roof of the sunspace, the heating load of the bedroom can be reduced by 21.11 kWh/m2. When the roof was an opaque, flat roof, the heat gain of the sunspace was negatively correlated to depth and positively correlated to WWR. When the roof tilt angle θ was 40°, the optimal depth of the sunspace was 1.1 m. When the sunspace depth was reduced from 1.2 m to 1.1 m and WWR was reduced from 70% to 50%, the average temperature of the sunspace was increased by 0.5 °C and the heating load of the bedroom was reduced by 0.85 kWh/m². When using double low-E glazing on the façade, the heating load of the bedroom can be reduced by 11.92 kWh/m2, and the average temperature of the sunspace can be increased by 9.64 °C. In summer, the overheating problem of the bedroom in Qingdao can be basically solved by opening the exterior windows for ventilation during the day and opening the door of the sunspace at night.

Author Contributions

Conceptualization, Q.M. and X.W.; methodology, Q.M. and W.G.; software, Q.M. and C.X.; validation, C.X.; writing—original draft preparation, X.C. and Q.M.; writing—review and editing, X.C. and Q.M.; visualization, C.X.; funding acquisition, Q.M. All authors have read and agreed to the published version of the manuscript.

Funding

This research is supported by grants from the National Natural Science Foundation of China (No.52108015) and the Natural Science Foundation of Shandong Province (No. ZR201910280141).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

The data presented in this study are available upon request from the corresponding author.

Acknowledgments

We would like to express our gratitude to the editors and reviewers for their thoughtful comments and constructive suggestions on improving the quality of the paper.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Wang, D.; Hu, L.; Du, H.; Liu, Y.; Huang, J.; Xu, Y.; Liu, J.J.R.; Reviews, S.E. Classification, experimental assessment, modeling methods and evaluation metrics of Trombe walls. Renew. Sustain. Energy Rev. 2020, 124, 109772. [Google Scholar] [CrossRef]
  2. Ma, Q.; Chen, X.; Wang, X.; Gao, W.; Wei, X.; Fukuda, H. A review of the application of sunspace in buildings. Energy Sources Part A Recovery Util. Environ. Eff. 2021. [Google Scholar] [CrossRef]
  3. Carpio, M.; Hubner, J.; Osorio, E. The influence of sunspaces on the design of thermal envelopes in energy demand in southern chile. Appl. Mech. Mater. 2018, 878, 152–158. [Google Scholar] [CrossRef]
  4. Kolokotsa, D.; Santamouris, M.; Synnefa, A.; Karlessi, T. Passive Solar Architecture. In Comprehensive Renewable Energy; Sayigh, A., Ed.; Elsevier: Oxford, UK, 2012; Volume 637–665. [Google Scholar]
  5. Cui, Y.; Sun, N.; Cai, H.; Li, S. Indoor Temperature Improvement and Energy-Saving Renovations in Rural Houses of China’s Cold Region—A Case Study of Shandong Province. Energies 2020, 13, 870. [Google Scholar] [CrossRef] [Green Version]
  6. Gainza-Barrencua, J.; Odriozola-Maritorena, M.; Hernandez_Minguillon, R.; Gomez-Arriaran, I. Energy savings using sunspaces to preheat ventilation intake air: Experimental and simulation study. J. Build. Eng. 2021, 40, 102343. [Google Scholar] [CrossRef]
  7. Kisilewicz, T. Glazed building wall as a solar thermal collector. Arch. Civ. Mech. Eng. 2009, 9, 83–99. [Google Scholar] [CrossRef]
  8. Sun, H.; Leng, M. Analysis on building energy performance of Tibetan traditional dwelling in cold rural area of Gannan. Energy Build. 2015, 96, 251–260. [Google Scholar] [CrossRef]
  9. Hilliaho, K.; Lahdensivu, J.; Vinha, J. Glazed space thermal simulation with IDA-ICE 4.61 software—Suitability analysis with case study. Energy Build. 2015, 89, 132–141. [Google Scholar] [CrossRef]
  10. Saleh, P.H. Thermal performance of glazed balconies within heavy weight/thermal mass buildings in Beirut, Lebanon’s hot climate. Energy Build. 2015, 108, 291–303. [Google Scholar] [CrossRef]
  11. Oliveti, G.; Simona, M.; Ruffolo, S. Solar Contribution Evaluation for Building Attached Sunspace in the Mediterranean Climate. In Proceedings of the International Conference Passive and Low Energy Cooling for the Built Environment, Santorini, Greece, 19–21 May 2005. [Google Scholar]
  12. Hilliaho, K.; Mäkitalo, E.; Lahdensivu, J. Energy saving potential of glazed space: Sensitivity analysis. Energy Build. 2015, 99, 87–97. [Google Scholar] [CrossRef]
  13. Suárez, M.J.; Gutiérrez, A.J.; Pistono, J.; Blanco, E. CFD analysis of heat collection in a glazed gallery. Energy Build. 2011, 43, 108–116. [Google Scholar] [CrossRef]
  14. Bakos, G.C. Improved energy management method for auxiliary electrical energy saving in a passive-solar-heated residence. Energy Build. 2002, 34, 699–703. [Google Scholar] [CrossRef]
  15. Asdrubali, F.; Cotana, F.; Messineo, A. On the evaluation of solar greenhouse efficiency in building simulation during the heating period. Energies 2012, 5, 1864–1880. [Google Scholar] [CrossRef] [Green Version]
  16. Ma, L.; Zhang, X.; Li, D.; Arıcı, M.; Yıldız, Ç.; Li, Q.; Zhang, S.; Jiang, W. Influence of sunspace on energy consumption of rural residential buildings. Sol. Energy 2020, 211, 336–344. [Google Scholar] [CrossRef]
  17. Bakos, G.C.; Tsagas, N.F. Technology, thermal analysis and economic evaluation of a sunspace located in northern Greece. Energy Build. 2000, 31, 261–266. [Google Scholar] [CrossRef]
  18. Gainza-Barrencua, J.; Odriozola-Maritorena, M.; Barrutieta, X.; Gomez-Arriaran, I.; Hernández Minguillón, R. Use of sunspaces to obtain energy savings by preheating the intake air of the ventilation system: Analysis of its main characteristics in the different Spanish climate zones. J. Build. Eng. 2022, 62, 105331. [Google Scholar] [CrossRef]
  19. Wang, W.; Yuan, M.; Li, Y.-Z.; Li, C. Numerical investigation on the impact of an on-top sunspace passive heating approach for typical rural buildings in northern China. Sol. Energy 2019, 186, 300–310. [Google Scholar] [CrossRef]
  20. Gupta, R.; Tiwari, G.N. Effect of latitude on weighted solar fraction of north partition wall for various shapes of solarium. Build. Environ. 2004, 39, 547–556. [Google Scholar] [CrossRef]
  21. Tiwari, G.N.; Gupta, A.; Gupta, R. Evaluation of solar fraction on north partition wall for various shapes of solarium by Auto-Cad. Energy Build. 2003, 35, 507–514. [Google Scholar] [CrossRef]
  22. Ignjatović, D.; Jovanović Popović, M.; Kavran, J. Application of sunspaces in fostering energy efficiency and economical viability of residential buildings in Serbia. Energy Build. 2015, 98, 3–9. [Google Scholar] [CrossRef]
  23. Allesina, G.; Ferrari, C.; Muscio, A.; Pedrazzi, S. Easy to implement ventilated sunspace for energy retrofit of condominium buildings with balconies. Renew. Energy 2019, 141, 541–548. [Google Scholar] [CrossRef]
  24. Ulpiani, G.; Summa, S.; di Perna, C. Sunspace coupling with hyper-insulated buildings: Investigation of the benefits of heat recovery via controlled mechanical ventilation. Sol. Energy 2019, 181, 17–26. [Google Scholar] [CrossRef]
  25. Owrak, M.; Aminy, M.; Jamal-Abad, M.T.; Dehghan, M. Experiments and simulations on the thermal performance of a sunspace attached to a room including heat-storing porous bed and water tanks. Build. Environ. 2015, 92, 142–151. [Google Scholar] [CrossRef]
  26. Li, Q.; Wang, Y.; Ma, L.; Arıcı, M.; Li, D.; Yıldız, Ç.; Zhu, Y. Effect of sunspace and PCM louver combination on the energy saving of rural residences: Case study in a severe cold region of China. Sustain. Energy Technol. Assess. 2021, 45, 101126. [Google Scholar] [CrossRef]
  27. Grudzińska, M. Glazed balconies as passive greenhouse systems—Potential of their use in Poland. Build. Serv. Eng. Res. Technol. 2016, 37, 555–572. [Google Scholar] [CrossRef]
  28. Arıcı, M.; Bilgin, F.; Nižetić, S.; Karabay, H. PCM integrated to external building walls: An optimization study on maximum activation of latent heat. Appl. Therm. Eng. 2020, 165, 114560. [Google Scholar] [CrossRef]
  29. Sánchez-Ostiz, A.; Monge-Barrio, A.; Domingo-Irigoyen, S.; González-Martínez, P. Design and experimental study of an industrialized sunspace with solar heat storage. Energy Build. 2014, 80, 231–246. [Google Scholar] [CrossRef]
  30. Suárez López, M.J.; Gutiérrez Trashorras, A.J.; Parrondo Gayo, J.L.; Blanco Marigorta, E. Analysis of an Attached Sunspace with a Thermal Inertia Floor. Energies 2018, 11, 1136. [Google Scholar] [CrossRef] [Green Version]
  31. Aelenei, D.; de Azevedo Leal, H.; Aelenei, L. The use of attached-sunspaces in retrofitting design: The case of residential buildings in Portugal. Energy Procedia 2014, 48, 1436–1441. [Google Scholar] [CrossRef] [Green Version]
  32. Bourdeau, L. How people use their sunspaces. Batim. Int. Build. Res. Pract. 1988, 16, 167–171. [Google Scholar] [CrossRef]
  33. Liu, Z.; Wu, D.; Li, J.; Yu, H.; He, B. Optimizing Building Envelope Dimensions for Passive Solar Houses in the Qinghai-Tibetan Region: Window to Wall Ratio and Depth of Sunspace. J. Therm. Sci. 2019, 28, 1115–1128. [Google Scholar] [CrossRef]
  34. Ulpiani, G.; Giuliani, D.; Romagnoli, A.; di Perna, C. Experimental monitoring of a sunspace applied to a NZEB mock-up: Assessing and comparing the energy benefits of different configurations. Energy Build. 2017, 152, 194–215. [Google Scholar] [CrossRef]
  35. Babaee, F.; Fayaz, R.; Sarshar, M. The optimum design of sunspaces in apartment blocks in cold climate. Archit. Sci. Rev. 2016, 59, 239–253. [Google Scholar] [CrossRef]
  36. Mihalakakou, G.; Ferrante, A. Energy conservation and potential of a sunspace: Sensitivity analysis. Energy Convers. Manag. 2000, 41, 1247–1264. [Google Scholar] [CrossRef]
  37. Fernández-González, A. Analysis of the thermal performance and comfort conditions produced by five different passive solar heating strategies in the United States midwest. Sol. Energy 2007, 81, 581–593. [Google Scholar] [CrossRef]
  38. Oliveti, G.; De Simone, M.; Ruffolo, S. Evaluation of the absorption coefficient for solar radiation in sunspaces and windowed rooms. Sol. Energy 2008, 82, 212–219. [Google Scholar] [CrossRef]
  39. Zhu, X.; Liu, J.; Yang, L.; Hu, R. Energy performance of a new Yaodong dwelling, in the Loess Plateau of China. Energy Build. 2014, 70, 159–166. [Google Scholar] [CrossRef]
  40. Ma, Q.; Gao, W.; Fukuda, H.; Wei, X.; Chen, X. Heat Utilization of a Ventilated Sunspace in a Non-Insulated Apartment. Energy Sources Part A Recovery Util. Environ. Eff. 2020. [Google Scholar] [CrossRef]
  41. Chiesa, G.; Simonetti, M.; Ballada, G. Potential of attached sunspaces in winter season comparing different technological choices in Central and Southern Europe. Energy Build. 2017, 138, 377–395. [Google Scholar] [CrossRef]
  42. Monge-Barrio, A.; Sánchez-Ostiz, A. Energy efficiency and thermal behaviour of attached sunspaces, in the residential architecture in Spain. Summer Conditions. Energy Build. 2015, 108, 244–256. [Google Scholar] [CrossRef]
  43. Testo-174H. Available online: http://www.fireball.net.cn/pro_info.php?id=532 (accessed on 12 January 2021).
  44. Testo-175T3. Available online: https://www.testo.com/en-SG/testo-175-t3/p/0572-1753 (accessed on 12 January 2021).
  45. TC96-V DT Controller. Available online: https://item.taobao.com/item.htm?spm=a1z0d.6639537/tb.1997196601.4.3b2174845Y1oJU&id=542390866220 (accessed on 12 January 2021).
  46. Vantage Pro2 Weather Station. Available online: https://ishare.iask.sina.com.cn/f/63596817.html (accessed on 12 January 2021).
  47. Bataineh, K.M.; Fayez, N. Analysis of thermal performance of building attached sunspace. Energy Build. 2011, 43, 1863–1868. [Google Scholar] [CrossRef]
  48. ASHRAE. ASHRAE Guideline 14–2014. Measurement of Energy, Demand, and Water Savings; ASHRAE: Atlanta, GA, USA, 2014. [Google Scholar]
  49. Bastien, D. Analysis of the Solar Radiation Distribution and Passive Thermal Response of an Attached Solarium or Greenhouse. In Proceedings of the International High Performance Buildings Conference, West Lafayette, IN, USA, 12–15 July 2010. [Google Scholar]
Figure 1. The flowchart of the paper structure.
Figure 1. The flowchart of the paper structure.
Sustainability 15 00176 g001
Figure 2. The studied old apartment with the sunspace is indicated in the red circle.
Figure 2. The studied old apartment with the sunspace is indicated in the red circle.
Sustainability 15 00176 g002
Figure 3. The floor plan of the study room and the locations of the thermometers (dimensions in mm).
Figure 3. The floor plan of the study room and the locations of the thermometers (dimensions in mm).
Sustainability 15 00176 g003
Figure 4. Indoor condition of Bedroom 1, which was controlled by the Delta-temperature-controlled ventilation system.
Figure 4. Indoor condition of Bedroom 1, which was controlled by the Delta-temperature-controlled ventilation system.
Sustainability 15 00176 g004
Figure 5. A schematic diagram of the sunspace in the parametric simulation model.
Figure 5. A schematic diagram of the sunspace in the parametric simulation model.
Sustainability 15 00176 g005
Figure 6. The relationship between the temperature and solar radiation variations in the ‘Door−closed’ mode during the heating period.
Figure 6. The relationship between the temperature and solar radiation variations in the ‘Door−closed’ mode during the heating period.
Sustainability 15 00176 g006
Figure 7. The relationship between temperature and solar radiation variations in the ‘Door-closed’ mode during the non-heating period.
Figure 7. The relationship between temperature and solar radiation variations in the ‘Door-closed’ mode during the non-heating period.
Sustainability 15 00176 g007
Figure 8. The relationship between the temperature and solar radiation variations in the ‘Door−open’ mode during the heating period.
Figure 8. The relationship between the temperature and solar radiation variations in the ‘Door−open’ mode during the heating period.
Sustainability 15 00176 g008
Figure 9. The relationship between the temperature and solar radiation variations in the ‘Door-open’ mode during the non-heating period.
Figure 9. The relationship between the temperature and solar radiation variations in the ‘Door-open’ mode during the non-heating period.
Sustainability 15 00176 g009
Figure 10. The relationship between the temperature and solar radiation variations in the DT−controlled ventilation mode during the heating period.
Figure 10. The relationship between the temperature and solar radiation variations in the DT−controlled ventilation mode during the heating period.
Sustainability 15 00176 g010
Figure 11. The relationship between the temperature and solar radiation variations in the DT−controlled ventilation mode during the non-heating period.
Figure 11. The relationship between the temperature and solar radiation variations in the DT−controlled ventilation mode during the non-heating period.
Sustainability 15 00176 g011
Figure 12. The relationship between temperature and solar radiation variations in the ‘Door-closed-window-closed’ mode in summer.
Figure 12. The relationship between temperature and solar radiation variations in the ‘Door-closed-window-closed’ mode in summer.
Sustainability 15 00176 g012
Figure 13. The relationship between temperature and solar radiation variations in the ‘Door-closed-window-open’ mode in summer.
Figure 13. The relationship between temperature and solar radiation variations in the ‘Door-closed-window-open’ mode in summer.
Sustainability 15 00176 g013
Figure 14. The relationship between temperature and solar radiation variations in the ‘Door-open-window-closed’ mode in summer.
Figure 14. The relationship between temperature and solar radiation variations in the ‘Door-open-window-closed’ mode in summer.
Sustainability 15 00176 g014
Figure 15. Comparison of the simulated and measured temperatures.
Figure 15. Comparison of the simulated and measured temperatures.
Sustainability 15 00176 g015
Figure 16. The results for different configurations of various combinations of WRR and θ: (a) average temperature of the sunspace without heating; (b) the amount of energy used for heating Bedroom 1.
Figure 16. The results for different configurations of various combinations of WRR and θ: (a) average temperature of the sunspace without heating; (b) the amount of energy used for heating Bedroom 1.
Sustainability 15 00176 g016
Figure 17. The results of different configurations of various combinations of WRR and θ using double low-E glazing on the roof: (a) average temperature of the sunspace without heating; (b) the amount of energy used for heating Bedroom 1.
Figure 17. The results of different configurations of various combinations of WRR and θ using double low-E glazing on the roof: (a) average temperature of the sunspace without heating; (b) the amount of energy used for heating Bedroom 1.
Sustainability 15 00176 g017
Figure 18. The results of different configurations of various combinations of depth and WWR: (a) average temperature of the sunspace without heating; (b) the amount of energy used for heating Bedroom 1.
Figure 18. The results of different configurations of various combinations of depth and WWR: (a) average temperature of the sunspace without heating; (b) the amount of energy used for heating Bedroom 1.
Sustainability 15 00176 g018
Figure 19. The results of different configurations of various combinations of depth and WWR with the optimized roof: (a) average temperature of the sunspace without heating; (b) the amount of energy used for heating Bedroom 1.
Figure 19. The results of different configurations of various combinations of depth and WWR with the optimized roof: (a) average temperature of the sunspace without heating; (b) the amount of energy used for heating Bedroom 1.
Sustainability 15 00176 g019
Figure 20. The results of different configurations of various combinations of depth and WWR using double low-E glazing: (a) sunspace average temperature without heating; (b) the amount of energy consumed for heating Bedroom 1.
Figure 20. The results of different configurations of various combinations of depth and WWR using double low-E glazing: (a) sunspace average temperature without heating; (b) the amount of energy consumed for heating Bedroom 1.
Sustainability 15 00176 g020
Figure 21. Results of the optimized sunspace with different combinations of FR and DT: (a) average temperature of Bedroom 1 without heating; (b) the amount of energy used for heating Bedroom 1.
Figure 21. Results of the optimized sunspace with different combinations of FR and DT: (a) average temperature of Bedroom 1 without heating; (b) the amount of energy used for heating Bedroom 1.
Sustainability 15 00176 g021
Table 1. Uncertainties of experimental devices.
Table 1. Uncertainties of experimental devices.
DeviceParameterResolutionRangeAccuracy
Testo-174H [43]Temperature0.1 °C−20 to +70 °C±0.5 °C
Testo-175T3 [44]Temperature0.1 °C−40 to +400 °C±0.5 °C
TC96-V DT controller [45]Delta temperature0.1 °C0 to +60 °C±1 °C
Vantage Pro2 weather station [46]Solar radiation1 W/m²0 to 1800 W/m²±5%
Temperature1 °C−68 to +64 °C±2 °C
Table 2. The monitoring schedule.
Table 2. The monitoring schedule.
TimeMode of User BehaviorDetails
DoorWindowVentilation
Heating2021/12/18~2021/12/21DT-controlled ventilationClosedClosedThe air supply starts when the DT is 3 °C and stops when the DT is 1 °C.
2021/12/23~2021/12/26Door-closedClosedClosedNo ventilation.
2021/12/28~2021/12/30Door-openOpenClosedNo ventilation.
HVAC system shut down 2022/04/17~2022/04/18DT-controlled ventilationClosedClosedThe air supply starts when the DT is 3 °C and stops when the DT is 1 °C.
2022/04/14~2022/04/16Door-closedClosedClosedNo ventilation.
2022/04/10~2022/04/13Door-openOpenClosedNo ventilation.
HVAC system shut down2022/07/05~2022/07/07DCWCClosedClosedNo ventilation.
2022/07/11~2022/07/14DCWOClosedOpenNo ventilation.
2022/07/16~2022/07/19DOWCOpenClosedNo ventilation.
Table 3. Parameters of the structural layers of the wall.
Table 3. Parameters of the structural layers of the wall.
LayerThickness (m)λ (W/m K)ρ (kg/m³)cp (J/kg K)
Stucco0.0250.691858837
Brick0.240.811920790
Stucco0.0250.691858837
Table 4. Parameters of the structural layers of the floor.
Table 4. Parameters of the structural layers of the floor.
LayerThickness (m)λ (W/m K)ρ (kg/m³)cp (J/kg K)
Tile0.0190.06368590
Concrete0.1010.531280840
Gypsum0.0150.16784.9830
Table 5. Parameters of the structural layers of the window.
Table 5. Parameters of the structural layers of the window.
U-Factor (W/m² K)SHGCVT
Window2.9530.390.27
Table 6. Parameters of the structural layers of the roof.
Table 6. Parameters of the structural layers of the roof.
LayerThickness (m)λ (W/m K)ρ (kg/m³)cp (J/kg K)
Concrete0.20.5212805000
Air gap0.18---
Tile0.020.06368590
Table 7. Weather data summary.
Table 7. Weather data summary.
Monthly MeansJanuaryFebruaryMarchAprilMayJuneJulyAugustSeptemberOctoberNovemberDecember
Global Horiz Radiation (Avg Hourly) (Wh/sq.m)297351416443457414365373364341266257
Global Horiz Radiation (Avg Daily Total) (Wh/sq.m)291837194910573863845973518149674466375626822465
Dry Bulb Temperature (Avg Monthly) (°C)0271319222626221681
Dew Point Temperature (Avg Monthly) (°C)−6−50491722211582−5
Relative Humidity (Avg Monthly) (%)625664605776807468626559
Wind Direction (Monthly Mode) (degrees)3303401801801801601601601603400330
Wind Speed (Avg Monthly) (m/s)334434322332
Table 8. Input parameters of the sunspace in the simulation program.
Table 8. Input parameters of the sunspace in the simulation program.
ParametersRangeIncrementAmount
θ (slope angle of the roof)0~60°10°7
WRR (window-to-roof ratio)0~90%10%10
RGT (roof glazing type)Single glazing/double low-E glazing-2
FGT (façade glazing type)Single glazing/double low-E glazing-2
WWR (window-to-wall ratio)10~90%10%9
Depth0.6~1.5 m0.1 m10
FR (flow rate)0.00~0.06 m3/s0.01 m3/s6
DT (delta temperature)0~8 °C1 °C9
Table 9. Characteristics of the glazing materials of the sunspace.
Table 9. Characteristics of the glazing materials of the sunspace.
Glazing TypeThermal Conductivity (W/m²·K)SHGCVT
Single glazing4.60.280.6
Double low-E glazing1.480.670.78
Table 10. Simulation error index of each room.
Table 10. Simulation error index of each room.
RoomMBECV (RMSE)
Sunspace−3%6%
Bedroom 12%2%
Bedroom 01%1%
Table 11. Optimized sunspace configurations and their energy savings.
Table 11. Optimized sunspace configurations and their energy savings.
Configurationsθ (o)WRRRGTFGTWWRDepth (m)FR (m3/s)DT (°C)Heating Loads (kWh/m2)Efficiency
Conf_000%NoneSingle glazing70%1.20-39.15-
Conf_1090%Single glazingSingle glazing70%1.20-38.242%
Conf_22090%Single glazingSingle glazing70%1.20-36.117%
Conf_34090%Single glazingSingle glazing70%1.20-34.2513%
Conf_44090%Double low-E glazingSingle glazing70%1.20-13.1466%
Conf_54090%Double low-E glazingSingle glazing50%1.10-12.8367%
Conf_64090%Double low-E glazingDouble low-E glazing50%1.10-0.9198%
Conf_74090%Double low-E glazingDouble low-E glazing50%1.10.0310.3399%
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

Ma, Q.; Xu, C.; Chen, X.; Gao, W.; Wei, X. Experimental and Simulation Research on the Energy-Saving Potential of a Sunspace—Taking an Apartment in Qingdao as an Example. Sustainability 2023, 15, 176. https://doi.org/10.3390/su15010176

AMA Style

Ma Q, Xu C, Chen X, Gao W, Wei X. Experimental and Simulation Research on the Energy-Saving Potential of a Sunspace—Taking an Apartment in Qingdao as an Example. Sustainability. 2023; 15(1):176. https://doi.org/10.3390/su15010176

Chicago/Turabian Style

Ma, Qingsong, Cui Xu, Xiaofei Chen, Weijun Gao, and Xindong Wei. 2023. "Experimental and Simulation Research on the Energy-Saving Potential of a Sunspace—Taking an Apartment in Qingdao as an Example" Sustainability 15, no. 1: 176. https://doi.org/10.3390/su15010176

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