Songyang Liuabc and
Huaming Yang*abc
aCentre for Mineral Materials, School of Minerals Processing and Bioengineering, Central South University, Changsha 410083, China. E-mail: hmyang@csu.edu.cn; Fax: +86-731-88710804; Tel: +86-731-88830549
bKey Laboratory for Mineral Materials and Application of Hunan Province, Central South University, Changsha 410083, China
cState Key Laboratory of Powder Metallurgy, Central South University, Changsha 410083, China
First published on 29th April 2016
This paper aimed to develop a novel form-stable composite phase change material (PCM) by infiltrating molten Na2SO4 into a mullite-corundum porous ceramic preform (M-PCP). Sufficient coal-series kaolinite (Kc), aluminum hydroxide, aluminum fluoride and graphite were mixed and subsequently heated in air at 1450 °C to produce M-PCP. The microstructure, thermal properties and thermal reliability of the composite PCMs were characterized by thermogravimetric and differential scanning calorimetry, X-ray diffraction and scanning electron microscopy. The results indicate that the M-PCP/Na2SO4 was 54.33 J g−1 at its melting temperature of 882.17 °C. Impregnation experiments and numerical simulation demonstrated high-temperature chemical compatibility and wettability between molten Na2SO4 and M-PCP. The M-PCP/Na2SO4 composite showed good thermal stability after 30 thermal shock cycles, and could potentially be used in the thermal energy storage field.
Among the PCMs investigated, Na2SO4 has been found to exhibit many desirable characteristics, such as high latent heat, good thermal reliability after a large number of melt/freeze cycles and commercial availability at reasonable cost.39,40 Despite these desirable properties, the application of Na2SO4, has been hampered in many ways by the complicated energy storage systems and the leakage of the PCMs. Therefore, various methods have been developed to solve these problems.
Recently, a new type of PCM/ceramic composite material, based on a porous ceramic preform with the phase change material distributed in the pores, has aroused much attention for its attractive properties. It not only integrates the advantages of traditional phase change materials and the sensible heat of the ceramics, but is also expected to solve the application problem, particularly because of its encapsulation, low thermal conductivity and the fact that it is not corrosive to the preform. Most previous research was aimed at studying the impregnation of PCMs into the porous ceramic preform prepared by either mixed sintering41,42 or the sol–gel method.43,44 But its lower phase transition temperature, specific heat capacity and its poor thermal shock resistance to a large extent limit its application as a high temperature thermal energy storage material.45–48 Therefore, it is necessary to fabricate novel composites which can improve the thermal energy storage effectively and which are expected to show high performance.
Mullite (3Al2O3·2SiO2) and mullite-based ceramic materials have a close thermal expansion match, low density, and excellent high-temperature strength.49,50 Kc has been studied extensively in the last few decades for serving the conventional ceramics industry and there has been renewed interest in the conversion of Kc to mullite, due to its low cost and relatively low sintering temperature. It is worth mentioning that the mullite made by the conventional method is generally dense on account of the excess SiO2 existing in Kc. Therefore, the addition of a certain amount of Al(OH)3 powder can reduce the amount of glass phase (SiO2), increase the amount of mullite51 and improve the overall properties of the composite.
In the present work, we aimed to prepare an M-PCP/Na2SO4 composite to be used for thermal energy storage. Kc and aluminum hydroxide were the main raw materials. Aluminum fluoride and graphite power were used as the cosolvent and pore former. A corundum-mullite porous ceramic preform (C-PCP) was also prepared by adding different amounts of Kc to investigate the effect of the Kc content on the properties of the M-PCP/Na2SO4 composite. In addition, the structure, thermal stability, and high temperature thermal performance of the composites were investigated systematically.
Samples | Kc | Al(OH)3 | AlF3/Kc + Al(OH)3 | Graphite/Kc + Al(OH)3 |
---|---|---|---|---|
1 | 70 | 30 | 10 | 10 |
2 | 30 | 70 | 10 | 10 |
![]() | (1) |
![]() | (2) |
Continuity:
∂t(ρ) + ∂i(ρui) = 0 | (3) |
Momentum:
∂t(ρui) + ∂j(ρuiuj) = μ∂jjui − ∂iP + ρgi + Si | (4) |
Thermal energy:
∂t(ρh) + ∂t(ρΔH) + ∂i(ρuih) = ∂i(k∂iT) | (5) |
In these relations, ui is the fluid velocity, ρ is the PCM's density, μ is the dynamic viscosity, P is the pressure, g is the gravitational acceleration, k is the thermal conductivity and h is the sensible enthalpy which is defined as follows:
![]() | (6) |
The enthalpy, H, is therefore:
H = h + ΔH | (7) |
![]() | (8) |
In eqn (2), Si are the Darcy's law damping terms (as source term) that are added to the moment equation due to phase change effect on convection. It is defined as follows:
![]() | (9) |
The coefficient C is a mushy zone constant which is fixed at a value of 105 kg m−3 s−1 for the present study. The model for the numerical study is created using pre-processor software GAMBIT 2.3.16. Meshing of the numerical model is generated and the boundary is applied at an appropriate surface. In this study, the grid is implemented using the GAMBIT software and then exported to FLUENT for problem solving. The numerical solution is implemented using a commercial CFD program Fluent 6.3.26.
![]() | ||
Fig. 1 Photographs of (a) mixed raw powders, (b) green compact, (c) M-PCP and (d) M-PCP/Na2SO4 composite, (e) pore size distribution of M-PCP and C-PCP, (f) XRD patterns of the samples. |
![]() | ||
Fig. 2 XRD patterns of (a) M-PCP and C-PCP, SEM images of (b) C-PCP, (c) M-PCP, (e) C-PCP/Na2SO4 composite, (f) M-PCP/Na2SO4 composite and (d) corresponding EDS spectra of C-PCP, M-PCP. |
The thermal properties of the composite PCMs mainly consist of thermal storage and release performance, and thermal reliability or stability after a long period of utilization. The phase change temperature and phase change enthalpy of Na2SO4, and C-PCP/Na2SO4 and M-PCP/Na2SO4 composites were investigated by DSC thermal analysis (Fig. 3a). The DSC curve of Na2SO4 is used as a reference to evaluate the variations in thermal properties of the two composites. The phase change temperature (Tm) and latent heat (ΔH) of Na2SO4 are 882.31 °C and 116.7 J g−1. Compared with Na2SO4, the phase change temperatures of the C-PCP/Na2SO4 and M-PCP/Na2SO4 composites are 881.96 °C and 882.17 °C, respectively. The latent heats of the C-PCP/Na2SO4 and M-PCP/Na2SO4 composites are 49.52 J g−1 and 54.33 J g−1, respectively. Table 2 shows that the open porosity of the M-PCP is larger than that of C-PCP, which indicates that M-PCP has more adsorption sites than C-PCP. The infiltration rate of M-PCP is also larger than that of C-PCP, indicating that the M-PCP has the better infiltration effect. The more PCM is filled into the porous ceramics preform, the greater the improvement in latent heat achieved. Therefore, the latent heat of the M-PCP/Na2SO4 composite is larger than that of the C-PCP/Na2SO4 composite. The thermal properties should be stable and exhibit no or minimal change after long periods of use. Therefore, thermal cycling tests were conducted to investigate the thermal reliability of the composite PCMs. There is almost no change in the appearance of the M-PCP/Na2SO4 composite phase change material after 30 thermal cycles (Fig. 3c). By contrast, there are a lot of cracks after 30 thermal cycles for the C-PCP/Na2SO4 composite, which indicates that the PCM Na2SO4 may be leaking from the C-PCP. This result shows that the thermal reliability of the M-PCP/Na2SO4 is superior to that of the C-PCP/Na2SO4 composite. This is mainly because the excess SiO2 in Kc would be consumed in forming the needle-like structure of mullite. Meanwhile, the mullite with its rigid frame structure will improve the thermal shock resistance of the samples.
Samples | Open porosity (wt%) | Infiltration rate (wt%) |
---|---|---|
C-PCP/Na2SO4 | 53.33 | 45.16 |
M-PCP/Na2SO4 | 56.21 | 49.51 |
The thermal stabilities of Na2SO4 and M-PCP/Na2SO4 were also analyzed by TG (Fig. 3b). The Na2SO4 and M-PCP/Na2SO4 are stable from room temperature to high temperatures up to 1000 °C. Based on this analysis, we may conclude that the M-PCP/Na2SO4 composite has a good thermal stability. It is observed that the phase change temperatures of the sample are very close to the original composite phase change materials after 30 thermal cycles (Fig. 3d), and that the latent heats of the samples are slightly lower than those of the original materials. Meanwhile, no leakage of Na2SO4 was observed after thermal cycling, which can be explained by the fact that the Na2SO4 is held in the network of the M-PCP by capillary force and surface tension force. Fig. 3d shows the SEM image of the M-PCP/Na2SO4 composite after 30 thermal cycles. It is observed that the morphology of M-PCP/Na2SO4 has no obvious variation in comparison with the PCM composite before (Fig. 2f) and after 30 thermal cycles (Fig. 3d). In conclusion, the M-PCP/Na2SO4 composite shows excellent thermal reliability.
We have studied the high temperature stability of Na2SO4 and M-PCP/Na2SO4 experimentally. We also studied the thermodynamic feasibility experimentally. The vapor pressure of Na2SO4 is low at high temperature, which shows that the volatilization of the phase-change material is very weak. In the gas–liquid–solid system, the overall possible chemical reactions in the process of impregnation are expressed by eqn (10)–(13).
![]() | (10) |
![]() | (11) |
![]() | (12) |
![]() | (13) |
To illustrate the theoretical feasibility of the route proposed below, the thermodynamic parameters at 1300 K in the process of impregnation were calculated. From the results of the thermodynamic calculation, the Gibbs free energy changes (ΔG) are 371.90 and 349.19 kJ mol−1 > 0 for eqn (10) and (11), respectively. So, eqn (10) and (11) cannot proceed, which indicates the PCM Na2SO4 has good thermodynamic thermal stability. On this basis we analyzed the stability of the MCPCP/Na2SO4 composite. The Gibbs free energy changes (ΔG) are 15205.05 and 186.21 kJ mol−1 > 0 for eqn (12) and (13), respectively, so the equation cannot proceed spontaneously in the impregnation process, which means that the mixed system (PCM Na2SO4 and M-PCP preform) is thermodynamically stable.
The infiltration ratio and the relative density are the two main factors used to evaluate the high temperature chemical compatibility of Na2SO4 and the M-PCP/Na2SO4 composite and to investigate the infiltration effects of the molten Na2SO4 into the porous ceramic preform. Fig. 4 shows the infiltration process of the PCM Na2SO4 being impregnated into the M-PCP. The M-PCP/Na2SO4 composite was fabricated by a spontaneous infiltration method. At the beginning, the M-PCP was covered by a certain amount of Na2SO4, and the PCM Na2SO4 would melt gradually and impregnate into the pores of M-PCP through the effects of surface tension and capillarity when the temperature moved beyond its melting point. Parts of the pores in the M-PCP would be blocked by the melting Na2SO4. Finally, the PCM would be gradually solidified as the temperature decreased. Fig. 5 indicates the effects of the infiltration process on the infiltration ratio and the relative density process on the ratio and the relative density of the composite. The high temperature chemical compatibility between Na2SO4 and the M-PCP composite were analyzed. In order to investigate the effects of the infiltration temperature on the infiltration ratio and relative density of the M-PCP/Na2SO4 composite, the infiltration temperature was varied from 900 °C to 1100 °C with the infiltration time fixed at 1 h. Similarly, in order to study the effects of the infiltration time on the infiltration ratio and relative density of the M-PCP/Na2SO4 composite, the infiltration time was varied from 0.5 h to 2.5 h with the infiltration temperature fixed at 1000 °C.
![]() | ||
Fig. 5 The effects of the infiltration temperature (a) and infiltration time (b) on the ratio and relative density of the M-PCP/Na2SO4 composite. |
With the increase in infiltration temperature, the infiltration ratio and relative density decrease slightly from 900 °C to 1000 °C (Fig. 5a), and then decrease obviously when the infiltration temperature is between 1000 °C and 1100 °C. According to previous research,52 along with the increase in the infiltration temperature, the surface tension, viscosity and the wetting angle of the molten salts decrease, whereas there is little change in viscosity and the wetting angle of the molten salts. In consequence, under certain infiltration times, the surface tension of the molten Na2SO4 has become a principal factor for infiltration temperatures between 900 °C and 1100 °C. With the increase in temperature, the surface tension of the molten salts decreases. This results in a decrease in the high temperature chemical compatibility between Na2SO4 and the M-PCP composite. To prevent the blockage of the pores by the solidification of the molten Na2SO4 in the process of infiltration, 900–1000 °C should be chosen as the optimum range according to the infiltration experiments. Further research is needed to ascertain at what temperature the optimal high temperature chemical compatibility between Na2SO4 and the M-PCP/Na2SO4 composite can be obtained. It can be seen (Fig. 5b) that the infiltration ratio and relative density increase rapidly with an increase in the infiltration time. The infiltration ratio and relative density tend to be constant after 1 h. Therefore, 1 h was chosen as the optimal infiltration time.
With the aim of producing thermal control applications of solid liquid phase change materials, the heat transfer numerical simulation between phase change materials and phase change composite materials is indispensable. The study of the heat transfer process between them plays an important part in determining the structure which will optimize the design and the effective utilization of thermal control systems for phase change composite materials. The numerical simulation of heat transfer in the solid–liquid phase change process for porous mediums includes the heat transfer calculation for the phase change materials and porous mediums. Fig. 6 shows the temperature profiles in the M-PCP/Na2SO4 composite during heat storage and heat retrieval. The temperature field changes from the inside to the outside during heat storage (Fig. 6a–d). By comparison and analysis, we find that the changes in the temperature gradients are less marked from 0.75 h to 1.5 h, mainly because the temperature in the furnace is lower than the melting temperature of Na2SO4. Then, the PCM Na2SO4 will melt gradually and impregnate the pores of M-PCP when the temperature in the furnace is above its melting point. The temperature gradients changed significantly from 1.5 h to 3 h. This phenomenon can be attributed to the melting characteristics of Na2SO4. At the beginning of the phase transition, melting of pure Na2SO4 was accelerated during heat storage (melting), due to the intensive natural convection in the melted Na2SO4. The existence of natural convection can provide the PCM composite with a large heat transfer coefficient and small thermal resistance, and this makes the change in temperature gradients obvious. The temperatures in the furnace will reach 950 °C and 1000 °C when the melting times are 2.25 and 3 h, respectively. The heat transfer performance of the M-PCP composite at the heating time of 3 h is higher than that of 2.25 h, and it will obtain a higher impregnation rate of Na2SO4. This means that the high temperature chemical compatibility between Na2SO4 and M-PCP/Na2SO4 will achieve better effects at 1000 °C.
By contrast, Fig. 6e–h show that for the temperature field changes from the outside to the inside during heat retrieval, the temperature gradients of each layer are becoming more stable. In the heat retrieval (cooling) duration, since Na2SO4 is impregnated in the pores of M-PCP, natural convection is consequently hampered. Therefore, freezing of the M-PCP/Na2SO4 composite is typically dominated by thermal conduction. While modelling the melting and freezing of the M-PCP/Na2SO4 composite, the natural convection and thermal conduction were taken into account at different durations. By comparison and analysis, the optimal heat transfer performance of the M-PCP/Na2SO4 composite can be achieved at 1000 °C.
This journal is © The Royal Society of Chemistry 2016 |