Next Article in Journal
Exploring Perceptions toward Emotional Intelligence in Chilean Construction Using a Qualitative Approach
Previous Article in Journal
Sustainability-Based Value Engineering Management as an Integrated Approach to Construction Projects
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Integration of Phase Change Materials in Service Areas of Building Envelopes for Improved Thermal Performance: An Experimental Study in Saudi Arabia

1
Department of Industrial Engineering, College of Engineering, Alfaisal University, Riyadh 11533, Saudi Arabia
2
Department of Mechanical Engineering, College of Engineering, Alfaisal University, Riyadh 11533, Saudi Arabia
*
Author to whom correspondence should be addressed.
Buildings 2024, 14(4), 904; https://doi.org/10.3390/buildings14040904
Submission received: 24 January 2024 / Revised: 18 March 2024 / Accepted: 19 March 2024 / Published: 27 March 2024
(This article belongs to the Section Building Energy, Physics, Environment, and Systems)

Abstract

:
This experimental study explores the integration of Phase Change Materials (PCMs) within building envelopes. The research specifically centers on the utilization of two microencapsulated paraffin-based PCMs with melting points of 37 °C and 43 °C. The study assesses their performance within cement and gypsum-based PCM composites, concentrating on service areas often overlooked in thermal analysis, including underground garages, staircases, and utility rooms. The experimental setup included constructing three chambers inside an underground garage during the hot months of June and July in Saudi Arabia. Two chambers were assigned to integrate the PCM, while the third chamber served as a control without PCM. The experiment unfolds in two phases. In the initial phase, the objective was to determine which PCM is more effective in reducing the heat load inside the chambers. This led to the adoption of the 43 °C PCM for the subsequent stage. The adoption of the 43 °C PCM resulted in a fourfold decrease in heat compared to the 37 °C PCM. The second phase investigates the integration of the selected PCM with cement and gypsum composites. The percentage of PCM incorporated into the concrete and gypsum composites was determined experimentally. For cement-based composites, the identified percentage that maintains material integrity is 20%, and for gypsum-based composites, it is 22%. The findings demonstrate a significant reduction in cooling load with PCM incorporation, with cement-based composites exhibiting superior thermal performance compared to gypsum-based alternatives and reducing the heat load by approximately 63%. Additionally, it was observed that concrete reduced the highest temperature during the day by 5.2 °C, which equates to about a 10% reduction, further enhancing comfort. Conducted over the course of two summer seasons, this study contributes valuable insights toward improving the quality of life for building occupants, considering various factors such as their living environment.

1. Introduction

In recent years, there has been a growing demand to improve the quality of life for citizens across the globe. This has resulted in countries initiating various programs to enhance their citizens’ quality of life and rank higher in the global quality of life index. The Kingdom of Saudi Arabia launched the Saudi Vision 2030 program in 2016, which includes several initiatives aimed at improving the quality of life in the country. One important factor that contributes to a better quality of life is the living environment, and thermal comfort is a key indicator of satisfaction with the living environment [1].
However, the escalating issue of high temperatures on a global scale, with the year 2023 experiencing exceptionally high temperatures worldwide, has placed additional pressure on the pursuit of thermal comfort. Saudi Arabia has faced the challenges posed by soaring temperatures, with the mercury hitting the 50s in the past two years.
Buildings contribute to a high percentage of overall energy consumption worldwide. The total energy used in buildings for heating, cooling, and ventilation in the European Union reaches 40% [2]. Most of the required energy is utilized to achieve the level of thermal comfort by heating and cooling the environment. It becomes rather necessary to improve the energy efficiency of buildings nationally and internationally [3].
The integration of phase change materials (PCMs) with building materials has been investigated in recent years. The PCM is used as a thermal storage that acts as a passive cooling and heating agent to achieve thermal comfort without excessive energy consumption. Many researchers developed different building prototypes using different types of PCM materials where the focus is shifting the peak load of energy demand [4,5,6]. In the summer, the electrical energy demand increases during the day to cool the building; it is more profitable to store it and release it at nighttime when the price of electricity is lower. The discomfort caused by high daytime temperature is eased and transferred to the nighttime when the outside temperature is cooler [7]. Moreover, the PCM integration is expected to lower the overall energy consumption and eventually reduce the carbon footprint [8]. Such a system is greatly useful in areas where the temperature gradient between day and night is quite high [9].
PCM works by absorbing/releasing a sufficient amount of heat, known as latent heat, to undergo a phase change. Thermal and mechanical stability, high thermal conductivity, non-corrosiveness, and minimal volume change during expansion are among the essential properties that ideal PCMs should possess to function effectively [10]. Despite the availability of different types of PCM materials, paraffin wax is mainly used as it exhibits most of the favored PCM properties [11].
PCMs are primarily integrated into building components, including construction materials, for their effective utilization [11]. These materials incorporate PCMs into building envelopes, such as walls, windows, ceilings, and floors [12,13]. It has been incorporated into construction materials through direct impregnation with wood [14], plaster [15], cement [16], and concrete [17]. PCM leakage is a significant problem with impregnation, which has been resolved by incorporating PCM directly into materials such as cement mortar [18] and insulation materials [19] through direct mixing. The drawback of direct mixing is the potential degradation of concrete properties that arise from the integration of PCM, such as strength and chemical structure [10,20].
To achieve optimal thermal efficiency, a substantial quantity of PCM is often incorporated, leading to an increasing risk of flammability, particularly when paraffin is employed. Various methods, including microencapsulation and nanoencapsulation, have been explored to mitigate this risk [21,22]. However, the cost considerations associated with nanoencapsulation versus microencapsulation PCMs significantly constrain their widespread adoption [23]. In addition, the incorporation of phase change materials into the building envelope is gradually replacing conventional, costly, and heavier construction methods aimed at improving thermal inertia [24].
The utilization of PCM in building applications for thermal management has been examined across various global regions. Li et al. [25] compiled studies on PCM applications in buildings conducted in different countries, with China, the United States, and Italy representing most of the research.
Our research explores the integration of PCMs within the service areas, such as underground garages of building envelopes in Saudi Arabia as well as investigates the integration of the selected PCM with cement and gypsum composites. The recent trend of increasing the size of commercial buildings in Saudi Arabia is accompanied by a corresponding surge in the size of underground garages required to service these developments [26]. This investigation is prompted by the substantial electricity consumption required for cooling buildings. Reports indicate that residential buildings in Saudi Arabia account for approximately 50% of the total electricity consumption [27]. Another incentive is the considerable temperature difference between day and night within the kingdom, making PCM a viable candidate for passive cooling. For instance, in May, temperatures can fluctuate from 48 °C during the day to 25 °C at night [28]. Lastly, the abundance of paraffin-based PCM materials in the kingdom, one of the primary byproducts of the oil industry, further drives this research effort [29].

2. Materials and Methods

2.1. Methods

The research approach involves a two-stage process. In the first stage, pure containment is employed to compare the performance of PCM 37(Average Particle size 19 microns, Heat capacity ≥ 190 J/g, and Melting temperature: 37 °C) and PCM 43 (Average Particle size 23 microns, Heat capacity ≥ 180 J/g, and Melting temperature: 43 °C), both grades are made of paraffin mix and encapsulated with Melamine Formaldehyde as shell material. In the second stage, the chosen PCM is integrated with both gypsum and cement separately to assess and compare the thermal effectiveness of each integration. Table 1 shows the thermal conductivity of PCM, gypsum, and cement. The cooling impact during exceptionally hot weather is determined by employing natural convection through an air cavity to remove the heat trapped in the melted PCM. Temperature differences between the indoor and ambient air are used to quantify the cooling impact, manifesting as a reduction in cooling load through a decreased heat transfer rate and a temporal lag in the peak interior temperature.

2.1.1. Selection of Experimental Parameters

In the first stage, a screening process was conducted to identify the appropriate PCM material for the experiment. Two types of PCM with melting points of 37 and 43 °C were used in experiments carried out in June and July of 2021. Thin, small containers with high thermal conductivity were designed to hold the PCM material, as shown in Figure 1a. The experiment utilized three identical chambers. The front wall was composed of 15 cm of concrete. A down-scaled test chamber measuring 40 cm × 40 cm × 40 cm was constructed behind the walls to simulate an indoor space. The test chambers were made of a 20 cm thick polystyrene sheet sealed with epoxy resin to achieve adiabatic boundary conditions, as illustrated in Figure 1a. Temperature measurements were taken using an Elitech GSP-6 device with a measurement error of 0.3 °C. The Elitech GSP-6 was installed inside the test chamber, and temperatures were recorded within the chamber, as shown in Figure 1b. The chambers were constructed in an internal underground parking garage beneath the main building of Alfaisal University in Riyadh, Saudi Arabia, as depicted in Figure 1c. The chamber dimensions are shown in Figure 1d.
In the second stage of the experiment, all three chambers were employed. Chamber I contained gypsum with PCM 43, Chamber II contained cement with PCM 43, and Chamber III served as the control without adding cement, gypsum, or PCM. Data collection took place during June and July 2022.

2.1.2. Specimen Preparation

Gypsum Specimen

The mixing of gypsum and PCM was performed manually for a period of 3 min, during which the powders were combined and slowly added to water. It should be noted that multiple attempts were made to identify an appropriate weight of gypsum and PCM to minimize any potential mechanical stress caused by the expansion of the PCM on the gypsum layer. After mixing, the resulting slurry was poured into a Styrofoam mold with dimensions of 40 cm × 40 cm × 2 cm and left to dry for 4 days. The gypsum wall was then removed from the mold and allowed to dry for an additional period of 3 days. The final composition is listed in Table 2.

Cement Specimen

To incorporate a phase change material into cement, the PCM powder or particles can be added during the mixing process. It is important to ensure that the PCM particles are uniformly dispersed in the mix to achieve an even thermal energy storage capacity throughout the resulting cement composite. In this study, the PCM was added directly to the cement mix, requiring careful attention to the mixing process to ensure the proper dispersion of the PCM particles. The amount of PCM that can be added is typically limited by the properties of both the PCM and the cement. Adding an excessive amount of PCM particles can result in reduced strength and durability of the composite material, as shown in Figure 2a. Therefore, multiple experiments were conducted to determine the appropriate amount of PCM to maintain the durability of the cement, as indicated in Table 3. The mixture yielded a solid integrated wall, as depicted in Figure 2b. The durability of the cement composite was determined after 7 days of room temperature drying where the ordinary Portland cement gained about 70% of its maximum strength as per ACI 308.1R.
All specimens were prepared in the laboratory at the constant room temperature (23 °C) to avoid any undesired effects due to the changes in the surrounding environment temperature.

3. Results and Discussion

This section presents the experimental results for all PCM containments, as explained in Section 2.

3.1. DSC

Differential scanning calorimetry (DSC) was employed to assess the thermal properties of PCM 37 and PCM 43, involving their melting points, latent heat values, and freezing. The DSC curves provide detailed information about the enthalpy changes associated with phase transitions, including melting, solidification, and thermal hysteresis exhibited by the materials [30]. Utilizing a heating rate of 5 °C/min, the tests were conducted over a temperature span from −20 °C to 90 °C to capture the complete thermal behavior of the phase change materials. The latent heat values were determined to be 195 J/g for PCM 37 and 210 J/g for PCM 43. In addition to the melting and solidification cycles, freezing cycles were also analyzed to comprehensively understand the thermal performance of the materials. Figure 3 and Figure 4 present the DSC curve depicting these thermal transitions and behaviors, including both melting and freezing phases for PCM 43 and 37, respectively.

3.2. Pure PCM Containment (Phase 1)

The daily temperatures during the two months study periods exhibited minimal variations and remained consistent. Throughout the months of June and July 2021, temperature fluctuations were observed every 15 s within the three chambers. Figure 5 provides a snapshot of these fluctuations at the end of June and the beginning of July. The cycles were repeated, with the charging heat in the chambers beginning early in the morning and reaching its peak around 5:00 p.m. Heat Discharging started as the outside temperature began to decrease. The discharge process continued throughout the night, usually until around 7 or 8 in the morning when the chambers reached similar temperatures around midnight.

PCM Selection

Both PCM 43 and PCM 37 effectively reduced temperatures during charging (heat absorption) during the day. However, they both dissipated heat during the night. The energy savings achieved through PCM utilization were evaluated. A typical cycle, starting from morning and spanning to the next day’s morning, was analyzed. Figure 6 illustrates this typical cycle for both the chamber containing PCM 43 and the control chamber. Similar analyses can be conducted for other cycles. The cycle started at 8:43 a.m. on 29 June 2021 and concluded at 8:58 a.m. on 30 June 2021. Both chambers started at the same temperature of 35.1 °C and gradually increased, reaching 49.1 °C in the chamber without PCM and 43.2 °C in the PCM chamber around 5 p.m. The temperature began to decline in both chambers, with the PCM chamber cooling at a higher rate, and both reached the same temperature around an hour past midnight. Subsequently, the temperature in the control chamber decreased more rapidly compared to the PCM chamber due to the heat dissipating from the PCM during the PCM phase change to a solid state. The heat continued to be released in the control chamber until 7:13 a.m. on June 30, while in the PCM 43 chamber, the heat was discharged until 8:58 a.m., just before the next cycle was about to begin.
Four distinct zones have been identified. Zones I and II represent the temperature difference between the two chambers, with the control chamber having a higher temperature. The heat in Zones I and II signifies the energy stored in the PCM during the phase change from solid to liquid. Zones III and IV represent temperature differences between the two chambers, with the PCM 43 chamber having a higher temperature. The heat in Zones III and IV represents the energy released from the PCM during the phase change from liquid to solid. The amount of heat absorbed or released in each zone is calculated using the sensible heat formula as follows:
Q = ρ v c T
where ρ is the air density (kg/m3), v is chamber volume (m3), c is the specific heat of the air (kJ/kg/°C), and ΔT (°C) is the temperature gradient.
Table 4 presents the amount of heat in each zone and the overall heat savings achieved through PCM 43. Similarly, the energy savings obtained by utilizing PCM 37 are computed and displayed in Figure 7 and Table 5.
Compared to PCM 37, the use of PCM 43 results in a significantly higher net energy savings. In fact, PCM 43 saves over four times as much energy as PCM 37, with a daily savings of 0.65 kJ compared to only 0.15 kJ for PCM 37. Furthermore, the PCM 43 chambers maintain a consistently cooler temperature throughout the cycle, as shown in Figure 6. The temperature in the PCM 43 chambers remains lower from 7:30 in the morning until 1:30 after midnight, a period of approximately 18 h, at which point the PCM 37 chambers begin to cool more effectively. This is caused by the early solidification of the PCM 43, which causes the discharge of heat over a period of six hours. By comparison, the temperature in the PCM 37 chambers does not decrease significantly during the night, and the maximum temperature drop is only around 1.0 °C (ΔT2), as opposed to the approximately 4 °C drop observed in the PCM 43 chambers (ΔT1) during the day which is about a four times temperature drop compared to PCM 37 as shown Figure 8. The results are similar across the summer season, particularly in June and July, as depicted in the box plot Figure 9, where PCM 43 demonstrates a higher temperature drop compared to PCM 37 composite during this time. For these reasons, all PCM integrations were tested using PCM 43.

3.3. PCM with Cement Mix (Phase 2)

The temperature cycles for the chambers without PCM and with PCM 43 mixed with cement are illustrated in Figure 10. These cycles were repeated, with temperatures ranging from the mid-30s to the upper 40s in June. Similarly, temperature fluctuations in July followed the same range as in June, cooling off towards the end of the month. A steady-state cycle is depicted in Figure 11. The net energy savings, derived from the quantification of sensible heat, are calculated and presented in Table 6. The heat was attentively quantified, contributing to the accurate assessment of net heat savings, where positive values represent the heat stored in the PCM during the day, and negative values indicate heat release from the PCM during the night.

3.4. PCM with Gypsum Mix (Phase 2)

The temperature cycles for the chamber without PCM and the chamber with PCM 43 mixed with gypsum are displayed in Figure 12. The temperatures inside follow a pattern similar to that in the cement chamber. A steady cycle is presented in Figure 13, and the quantity of heat saved through PCM utilization is detailed in Table 7.

3.5. Comparison between Cement and Gypsum

Both cement and gypsum integration have demonstrated the ability to store heat during the day and release it during the night, resulting in a relatively small and equivalent net heat gain. The PCM cement integration exhibits a high storage capability during the day but also releases a high amount of heat during the night, equivalent to approximately 85% of what was stored during the day, due to the high thermal conductivity of the cement. On the other hand, the PCM integrated into gypsum stored less heat during the day and released less heat during the night, with a percentage release of about 50%, which is attributed to the low thermal conductivity of gypsum.
In addition, integrating PCM with cement resulted in a prolonged period of lower temperatures lasting for 16 h, as shown in Figure 12. The maximum temperature drop using cement was 5.2 °C, compared to only 2 °C with gypsum. This difference was nearly regained during the night and early morning of the following day. The higher heat absorbed by cement integration shifts from the daytime peak to the night, where there is a lower demand for electricity. Moreover, cement is known for its durability and aligns with the common practice of using cement for final wall finishes in Saudi buildings. Consequently, we recommend adopting the cement integration method. Table 8 summarizes the superior performance of the cement–PCM 43 composites compared to the gypsum composite. The results can be generalized across the summer seasons, particularly in June and July, as depicted in the box plot Figure 14, where cement composite demonstrates a higher temperature drop compared to gypsum composite during this time. The practical ratios for cement, sand, and PCM were determined, as indicated in Table 3. The practical ratios for cement, sand, and PCM, as outlined in Table 3, represent the optimal combination to achieve maximum heat reduction. It is essential to note that surpassing these specified ratios degrades the strength of the cement composition.
These findings are consistent with prior research, as evidenced in studies conducted by [31,32,33,34,35,36,37], highlighting the effectiveness of PCM integration into building materials for lowering indoor temperatures and reducing energy consumption while preserving thermal comfort. In our study, the cement composite with 20% PCM 43 demonstrated a reduction in maximum temperature by 5.2 °C. Table 9 provides a summary of the literature findings.

4. Conclusions

In this study, microencapsulated paraffin-based PCM integrated with service areas of building envelopes was investigated to improve the thermal performance of service areas in buildings in Saudi Arabia. The experiment was conducted during the hottest months of June and July. PCM integration with building envelopes can significantly reduce indoor temperature and shift the cooling load from peak hours to off-peak hours, thereby reducing energy consumption and cost. In this study, we found that the use of 43 °C PCM integrated with cement can decrease the maximum temperature by 5.2 °C during the day, while the use of 43 °C PCM integrated with gypsum performs better at night by releasing less heat. However, the use of cement is recommended due to its better thermal and physical properties and its commonly used wall finish in Saudi Arabia. Additionally, the cooling load was shifted from peak hours to off-peak hours, reducing energy consumption during high-demand hours. Further research can investigate other PCM materials and integration methods to improve the thermal performance of service areas in building envelopes. Finally, we note that the integration of PCM with cement can decrease the cooling load for up to 16 h, making it a highly effective solution for improving thermal performance in service areas of building envelopes in hot climates. As a potential avenue for future research, investigating the use of PCM during winter could offer a viable solution for heating buildings in Saudi Arabia. Given the substantial temperature gradients between day and night in the region, PCM stands out as a suitable candidate for passive heating.

Author Contributions

Conceptualization, A.A., A.M.G. and K.H.A.; methodology, A.A., A.M.G. and K.H.A.; software, A.A., A.M.G., K.H.A. and A.N.D.; validation, A.A., A.M.G., K.H.A. and A.N.D.; formal analysis, A.A. and A.M.G.; investigation, K.H.A. and A.N.D.; resources, K.H.A. and A.N.D.; data curation, A.M.G. and K.H.A.; writing—original draft preparation, A.A. and A.M.G.; writing—review and editing A.A. and A.M.G.; visualization, K.H.A. and A.N.D.; supervision, A.A.; project administration, A.A. All authors have read and agreed to the published version of the manuscript.

Funding

The authors extend their appreciation to the Office of Research & Graduate Studies for their support under grant (IRG 21205), Alfaisal University.

Data Availability Statement

The original contributions presented in the study are included in the article, further inquiries can be directed to the corresponding authors.

Conflicts of Interest

The authors declare no conflicts of interest.

References

  1. Alijani, S.; Pourahmad, A.; Nejad, H.H.; Ziari, K.; Sodoudi, S. A new approach of urban livability in Tehran: Thermal comfort as a primitive indicator. Case study, district 22. Urban Clim. 2020, 33, 100656. [Google Scholar] [CrossRef]
  2. European Union. Directive 2010/31/EU of the European Parliament and of the Council of 19 May 2010 on the Energy Performance of Buildings; European Union: Maastricht, The Netherlands, 2010. [Google Scholar]
  3. Lu, S.; Li, Y.; Kong, X.; Pang, B.; Chen, Y.; Zheng, S.; Sun, L. A review of PCM energy storage technology used in buildings for the global warming solution. In Energy Solutions to Combat Global Warming; Zhang, X., Dincer, I., Eds.; Springer International Publishing: Cham, Switzerland, 2017; Volume 336, pp. 611–644. [Google Scholar] [CrossRef]
  4. Khudhair, A.M.; Farid, M.M. A review on energy conservation in building applications with thermal storage by latent heat using phase change materials. Energy Convers. Manag. 2004, 45, 263–275. [Google Scholar] [CrossRef]
  5. Hauer, A.; Mehling, H.; Schossig, P.; Yamaha, M.; Cabeza, L.F.; Martin, V.; Setterwall, F. International Energy Agency Implementing Agreement on Energy Conservation through Energy Storage Annex 17 ‘Advanced Thermal Energy Storage through Phase Change Materials and Chemical Reactions-Feasibility Studies and Demonstration Projects’ Final Report; International Energy Agency: Paris, France, 2005. [Google Scholar]
  6. Zhou, G.; Zhang, Y.; Wang, X.; Lin, K.; Xiao, W. An assessment of mixed type PCM-gypsum and shape-stabilized PCM plates in a building for passive solar heating. Sol. Energy 2007, 81, 1351–1360. [Google Scholar] [CrossRef]
  7. Fokaides, P.A.; Christoforou, E.; Ilic, M.; Papadopoulos, A. Performance of a Passive House under subtropical climatic conditions. Energy Build. 2016, 133, 14–31. [Google Scholar] [CrossRef]
  8. Diaconu, B.M.; Cruceru, M. Novel concept of composite phase change material wall system for year-round thermal energy savings. Energy Build. 2010, 42, 1759–1772. [Google Scholar] [CrossRef]
  9. Mathieu-Potvin, F.; Gosselin, L. Thermal shielding of multilayer walls with phase change materials under different transient boundary conditions. Int. J. Therm. Sci. 2009, 48, 1707–1717. [Google Scholar] [CrossRef]
  10. Regin, A.F.; Solanki, S.; Saini, J. Heat transfer characteristics of thermal energy storage system using PCM capsules: A review. Renew. Sustain. Energy Rev. 2008, 12, 2438–2458. [Google Scholar] [CrossRef]
  11. Akgün, M.; Aydın, O.; Kaygusuz, K. Experimental study on melting/solidification characteristics of a paraffin as PCM. Energy Convers. Manag. 2007, 48, 669–678. [Google Scholar] [CrossRef]
  12. Hawes, D.; Feldman, D.; Banu, D. Latent heat storage in building materials. Energy Build. 1993, 20, 77–86. [Google Scholar] [CrossRef]
  13. Souayfane, F.; Fardoun, F.; Biwole, P.H. Phase change materials (PCM) for cooling applications in buildings: A review. Energy Build. 2016, 129, 396–431. [Google Scholar] [CrossRef]
  14. Zalba, B.; Marín, J.M.; Cabeza, L.F.; Mehling, H. Review on thermal energy storage with phase change: Materials, heat transfer analysis and applications. Appl. Therm. Eng. 2003, 23, 251–283. [Google Scholar] [CrossRef]
  15. Lin, X.; Jia, S.; Liu, J.; Wang, W.; Cao, H.; Guo, X.; Sun, W. Fabrication of thermal energy storage wood based on graphene aerogel encapsulated polyethylene glycol as phase change material. Mater. Res. Express 2020, 7, 095503. [Google Scholar] [CrossRef]
  16. Athienitis, A.K.; Liu, C.; Hawes, D.; Banu, D.; Feldman, D. Investigation of the thermal performance of a passive solar test-room with wall latent heat storage. Build. Environ. 1997, 32, 405–410. [Google Scholar] [CrossRef]
  17. Li, H.; Liu, X.; Fang, G. Preparation and characteristics of n-nonadecane/cement composites as thermal energy storage materials in buildings. Energy Build. 2010, 42, 1661–1665. [Google Scholar] [CrossRef]
  18. Hadjieva, M.; Stoykov, R.; Filipova, T. Composite salt-hydrate concrete system for building energy storage. Renew. Energy 2000, 19, 111–115. [Google Scholar] [CrossRef]
  19. Zhou, D.; Zhao, C.Y.; Tian, Y. Review on thermal energy storage with phase change materials (PCMs) in building applications. Appl. Energy 2012, 92, 593–605. [Google Scholar] [CrossRef]
  20. Cunha, S.; Leite, P.; Aguiar, J. Characterization of Innovative Mortars with Direct Incorporation of Phase Change Materials. J. Energy Storage 2020, 30, 101439. [Google Scholar] [CrossRef]
  21. Boussaba, L.; Lefebvre, G.; Makhlouf, S.; Grados, A.; Royon, L. Investigation and Properties of a Novel Composite bio-PCM to Reduce Summer Energy Consumptions in Buildings of Hot and Dry Climates. Sol. Energy 2020, 214, 119–130. [Google Scholar] [CrossRef]
  22. Akeiber, H.; Nejat, P.; Majid, M.Z.A.; Wahid, M.A.; Jomehzadeh, F.; Famileh, I.Z.; Calautit, J.K.; Hughes, B.R.; Zaki, S.A. A Review on Phase Change Material (PCM) for Sustainable Passive Cooling in Building Envelopes. Renew. Sustain. Energy Rev. 2016, 60, 1470–1497. [Google Scholar] [CrossRef]
  23. Zhang, Y.; Zhou, G.; Lin, K.; Zhang, Q.; Di, H. Application of latent heat thermal energy storage in buildings: State-of-the-art and outlook. Build. Environ. 2007, 42, 2197–2209. [Google Scholar] [CrossRef]
  24. Huang, M.; Eames, P.; Hewitt, N. The application of a validated numerical model to predict the energy conservation potential of using phase change materials in the fabric of a building. Sol. Energy Mater. Sol. Cells 2006, 90, 1951–1960. [Google Scholar] [CrossRef]
  25. Li, F.; Wang, X.; Wu, D. Fabrication of multifunctional microcapsules containing n-eicosane core and zinc oxide shell for low-temperature energy storage, photocatalysis, and antibiosis. Energy Convers. Manag. 2015, 106, 873–885. [Google Scholar] [CrossRef]
  26. Alghamdi, M.S.; Beach, T.H.; Rezgui, Y. Reviewing the effects of deploying building information modelling (BIM) on the adoption of sustainable design in Gulf countries: A case study in Saudi Arabia. City Territ. Archit. 2022, 9, 18. [Google Scholar] [CrossRef]
  27. Raja, I.A.; Nicol, J.F.; McCartney, K.J.; Humphreys, M.A. Thermal comfort: Use of controls in naturally ventilated buildings. Energy Build. 2001, 33, 235–244. [Google Scholar] [CrossRef]
  28. Cui, Y.; Xie, J.; Liu, J.; Wang, J.; Chen, S. A review on phase change material application in building. Adv. Mech. Eng. 2017, 9, 1687814017700828. [Google Scholar] [CrossRef]
  29. Felimban, A.; Prieto, A.; Knaack, U.; Klein, T.; Qaffas, Y. Assessment of current energy consumption in residential buildings in Jeddah, Saudi Arabia. Buildings 2019, 9, 163. [Google Scholar] [CrossRef]
  30. Kuznik, F.; David, D.; Johannes, K.; Roux, J.-J. A review on phase change materials integrated in building walls. Renew. Sustain. Energy Rev. 2010, 15, 379–391. [Google Scholar] [CrossRef]
  31. Dafalla, M.A.; Al Shuraim, M.I. Efficiency of polystyrene insulated cement blocks in arid regions. Int. J. GEOMATE 2017, 13, 35–38. [Google Scholar] [CrossRef]
  32. Kosny, J.; Shukla, N.; Fallahi, A. Cost Analysis of Simple Phase Change Material-Enhanced Building Envelopes in Southern U.S. Climates; Building Technologies Program, U.S. Department of Energy, DOE/GO-102013-3692; National Renewable Energy Lab. (NREL): Golden, CO, USA, 2013.
  33. Marani, A.; Madhkhan, M. Thermal performance of concrete sandwich panels incorporating phase change materials: An experimental study. J. Mater. Res. Technol. 2021, 12, 760–775. [Google Scholar] [CrossRef]
  34. Kheradmand, M.; Vicente, R.; Azenha, M.; de Aguiar, J.L. Influence of the incorporation of phase change materials on temperature development in mortar at early ages: Experiments and numerical simulation. Constr. Build. Mater. 2019, 225, 1036–1051. [Google Scholar] [CrossRef]
  35. Frahat, N.B.; Amin, M.; Heniegal, A.M.; Ibrahim, O.M.O. Optimizing microencapsulated PCM ratios of sustainable cement mortar for energy savings in buildings. Constr. Build. Mater. 2023, 391, 131844. [Google Scholar] [CrossRef]
  36. Wang, X.; Yu, H.; Li, L.; Zhao, M. Research on temperature dependent effective thermal conductivity of composite-phase change materials (PCMs) wall based on steady-state method in a thermal chamber. Energy Build. 2016, 126, 408–414. [Google Scholar] [CrossRef]
  37. Marani, A. Hydration Kinetics, Microstructure, and Mechanical Strength Development of Cement-Based Composites Incorporating Phase Change Materials. Electronic Thesis and Dissertation Repository. Ph.D. Thesis, The University of Western Ontario, London, ON, Canada, 2022. Available online: https://ir.lib.uwo.ca/etd/8617 (accessed on 7 July 2022).
Figure 1. (a) PCM pure container (b) Test chamber with polystyrene sheets and concrete wall (c) Three test chambers installed in the underground garage (d) Dimensions of the test chamber.
Figure 1. (a) PCM pure container (b) Test chamber with polystyrene sheets and concrete wall (c) Three test chambers installed in the underground garage (d) Dimensions of the test chamber.
Buildings 14 00904 g001
Figure 2. Specimen preparation (a) with defects due to excessive PCM ratio (b) without defects at the optimal PCM ratio.
Figure 2. Specimen preparation (a) with defects due to excessive PCM ratio (b) without defects at the optimal PCM ratio.
Buildings 14 00904 g002
Figure 3. DSC curve for PCM 43.
Figure 3. DSC curve for PCM 43.
Buildings 14 00904 g003
Figure 4. DSC curve for PCM 37.
Figure 4. DSC curve for PCM 37.
Buildings 14 00904 g004
Figure 5. Temperature fluctuation inside the three chambers (every cycle represents 24 h period).
Figure 5. Temperature fluctuation inside the three chambers (every cycle represents 24 h period).
Buildings 14 00904 g005
Figure 6. One cycle temperature fluctuation for the PCM 43 chamber and the control chamber.
Figure 6. One cycle temperature fluctuation for the PCM 43 chamber and the control chamber.
Buildings 14 00904 g006
Figure 7. One cycle temperature fluctuation for the PCM 37 chamber and the control chamber.
Figure 7. One cycle temperature fluctuation for the PCM 37 chamber and the control chamber.
Buildings 14 00904 g007
Figure 8. Temperature drop Comparison between PCM 43 and PCM 37 during the day and night.
Figure 8. Temperature drop Comparison between PCM 43 and PCM 37 during the day and night.
Buildings 14 00904 g008
Figure 9. The box plot shows the superiority of PCM 43, dropping maximum temperature during June and July of 2022.
Figure 9. The box plot shows the superiority of PCM 43, dropping maximum temperature during June and July of 2022.
Buildings 14 00904 g009
Figure 10. Daily temperature fluctuations in PCM43+cement and control chambers.
Figure 10. Daily temperature fluctuations in PCM43+cement and control chambers.
Buildings 14 00904 g010
Figure 11. One cycle temperature fluctuation for the PCM 43+cement chamber and the control chamber.
Figure 11. One cycle temperature fluctuation for the PCM 43+cement chamber and the control chamber.
Buildings 14 00904 g011
Figure 12. Daily temperature fluctuations in PCM43+gypsum and control chambers.
Figure 12. Daily temperature fluctuations in PCM43+gypsum and control chambers.
Buildings 14 00904 g012
Figure 13. One cycle temperature fluctuation for PCM 43+gypsum chamber and the control chamber.
Figure 13. One cycle temperature fluctuation for PCM 43+gypsum chamber and the control chamber.
Buildings 14 00904 g013
Figure 14. The box plot shows the superiority of the Cement composite dropping maximum temperature during June and July of 2022.
Figure 14. The box plot shows the superiority of the Cement composite dropping maximum temperature during June and July of 2022.
Buildings 14 00904 g014
Table 1. The thermal conductivity of PCM, gypsum, and cement.
Table 1. The thermal conductivity of PCM, gypsum, and cement.
ComponentsThermal Conductivity (W/m·K)
PCM0.1–0.3
Gypsum0.314
Cement0.874
Table 2. PCM and Gypsum Quantities and percentages achieved After Multiple Trials.
Table 2. PCM and Gypsum Quantities and percentages achieved After Multiple Trials.
ComponentsAmountPercentage
Water2 kg0.44
Gypsum1.5 kg0.33
PCM1 kg0.22
Table 3. PCM and Cement Quantities and percentages achieved After Multiple Trials.
Table 3. PCM and Cement Quantities and percentages achieved After Multiple Trials.
ComponentsAmountPercentage
Sand3.5 kg0.53
Cement1.75 kg0.27
PCM1.32 kg0.20
Table 4. Net heat saved in a 24 h cycle using PCM 43 in pure containment.
Table 4. Net heat saved in a 24 h cycle using PCM 43 in pure containment.
ZoneChamberTfTiHeat (kJ)Heat Stored (+) or Released (−) in the PCM 43 (kJ)
IControl49.1035.101.08+0.46
PCM43.1035.100.62
IIControl39.8049.100.72+0.46
PCM39.8043.100.26
IIIControl33.9039.800.46−0.14
PCM35.7039.800.32
IVControl36.1033.900.17−0.14
PCM36.1035.700.03
Net heat saved in each cycle 0.65
Where Tf is the final Temperature, and Ti is the initial temperature.
Table 5. Net heat saved in a 24 h cycle using PCM 37 in pure containment.
Table 5. Net heat saved in a 24 h cycle using PCM 37 in pure containment.
ZoneChamberTfTiHeat kJHeat Stored (+) or Released (−) in the PCM 37 (kJ)
INo PCM49.1036.600.97+0.17
PCM46.9036.600.80
IINo PCM41.0049.100.63+0.17
PCM41.0046.900.46
IIINo PCM33.9041.000.55−0.09
PCM35.1041.000.46
IVNo PCM36.1033.900.17−0.09
PCM36.1035.100.08
Net heat saved in each cycle 0.15
Table 6. Net heat saved in one cycle using PCM 43+cement.
Table 6. Net heat saved in one cycle using PCM 43+cement.
ZoneChamberTfTiHeat kJHeat Stored (+) or Released (−) in the PCM 43+cement (kJ)
INo PCM49.3034.901.110.40
PCM44.1034.900.71
IINo PCM41.7049.300.590.40
PCM41.7044.100.19
IIINo PCM34.9041.700.53−0.33
PCM39.2041.700.19
IVNo PCM33.5034.900.11−0.22
PCM35.0039.200.33
VNo PCM35.6033.500.16−0.12
PCM35.6035.000.05
Net heat saved in each cycle 0.14
Table 7. Net heat saved in one cycle using PCM 43+gypsum.
Table 7. Net heat saved in one cycle using PCM 43+gypsum.
ZoneChamberTfTiHeat kJHeat Stored (+) or Released (−) in the PCM 43+gypsum (kJ)
INo PCM49.2037.000.940.15
PCM47.2037.000.79
IINo PCM41.3049.200.610.15
PCM41.3047.200.46
IIINo PCM34.2041.300.55−0.07
PCM35.1041.300.48
IVNo PCM36.9034.200.21−0.07
PCM36.9035.100.14
Net heat saved in each cycle 0.17
Table 8. Cement–PCM composite performance vs. Gypsum–PCM composite.
Table 8. Cement–PCM composite performance vs. Gypsum–PCM composite.
PerformanceCement–PCM CompositeGypsum–PCM Composite% Improvement (Cement Composite)
Heat storage during the day0.8 kJ0.3 kJ63%
Temperature
reduction
period/day
16 h12 h25%
Maximum
temperature
reduction
5.2 °C2 °C62%
Table 9. Literature findings on the maximum indoor temperature drop employing various percentages of PCM in cement.
Table 9. Literature findings on the maximum indoor temperature drop employing various percentages of PCM in cement.
ReferencePCM % in Cement CompositeMaximum Indoor Temperature Drop (°C)
[31]10–20%1.9–5.4
[32]20%7.2
[33]50%9.5
[34]40%5.8
[35]40%4
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

Alrashdan, A.; Ghaleb, A.M.; Ahmad, K.H.; Daoud, A.N. Integration of Phase Change Materials in Service Areas of Building Envelopes for Improved Thermal Performance: An Experimental Study in Saudi Arabia. Buildings 2024, 14, 904. https://doi.org/10.3390/buildings14040904

AMA Style

Alrashdan A, Ghaleb AM, Ahmad KH, Daoud AN. Integration of Phase Change Materials in Service Areas of Building Envelopes for Improved Thermal Performance: An Experimental Study in Saudi Arabia. Buildings. 2024; 14(4):904. https://doi.org/10.3390/buildings14040904

Chicago/Turabian Style

Alrashdan, Abdalla, Atef M. Ghaleb, Khalid Haj Ahmad, and Abdel Naser Daoud. 2024. "Integration of Phase Change Materials in Service Areas of Building Envelopes for Improved Thermal Performance: An Experimental Study in Saudi Arabia" Buildings 14, no. 4: 904. https://doi.org/10.3390/buildings14040904

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