Phase Change UDF Curve Fitting

Download as pdf or txt
Download as pdf or txt
You are on page 1of 25

CORE Metadata, citation and similar papers at core.ac.

uk
Provided by Brunel University Research Archive

Improved simulation of phase change processes in applications where

conduction is the dominant heat transfer mode

B.L. Gowreesunker *, S.A. Tassou, M. Kolokotroni

Howell Building, Mechanical Engineering, School of Engineering and Design, Brunel


University,Uxbridge, Middlesex, UB8 3PH, UK

* Corresponding Author: [email protected]

United Kingdom

Abstract:

This paper reports on the development, experimental validation and application of a semi-

empirical model for the simulation of the phase change process in phase change materials

(PCM). PCMs are now increasingly being used in various building materials such as

plasterboard, concrete or panels to improve thermal control in buildings and accurate

modelling of their behaviour is important to effectively capture the effects of storage on

indoor thermal conditions. Unlike many commercial simulation packages that assume very

similar melting and freezing behaviour for the PCM and no hysteresis, the methodology

employed treats the melting and freezing processes separately and this allows the inclusion

of the effect of hysteresis in the modelling. As demonstrated by the results in this paper,

this approach provides a more accurate prediction of the temperature and heat flow in the

material, which is of particular importance in providing accurate representation of indoor

thermal conditions during thermal cycling. The difference in the prediction accuracy of the

two methods is a function of the properties of the PCM. The smaller the hysteresis of the

PCM, the lower will be the prediction error of the conventional approach, and solution time

will become the determining factor in selecting the simulation approach in practical

applications.

1
Keywords: Phase Change Materials (PCM), Heat source/sink, Temperature hysteresis

2
Nomenclature:

t Time (s)

ρ Density (kg/m3)

H Enthalpy (J/kg)

uj Velocity vector

λ Thermal conductivity (W/m K)

T Temperature (K)

Direction vector

β Liquid Fraction

cp Specific heat capacity (J/Kg K)

L Latent heat capacity (J/kg)

Δt Time step (s)

href Reference enthalpy (J/kg)

f(T) User Defined Cumulative energy v/s


Temperature relation

Subscripts:

sol Solidus

liq Liquidus

m Melting

f Freezing

t Current time

t-1 Previous time

ref Reference values

3
1.0 Introduction

Phase change materials (PCM), in the context of buildings, refer to materials with enhanced

heat storage capabilities in a specific temperature range through accessing the latent

capacity of the materials. Conventional building materials such as bricks and concrete

provide thermal mass by sensible processes, that is, through changes in temperature.

However the growing trend by architects to design aesthetically pleasing buildings, with

maximum exposure to the outdoors and maximum sunlight, produces thermally less

massive buildings and therefore reduces the influence of the thermal inertia of the building

on indoor environment, leading to higher energy consumption.

Various studies have portrayed PCMs as a very effective way of enhancing the thermal

inertia properties of lightweight building materials [1,2]. Kuznik et al. [3] showed that a 5

mm thick 60% micro-encapsulated paraffin PCM wallboard stored energy equivalent to 8 cm

thick concrete. The increased interest in PCMs has led to various companies developing

commercial products for new or retrofit applications. Such products include BASF Micronal(r)

PCM boards, DuPont Energain(r) the EBB Clay PCM building boards, as well as a growing

number of experimental prototypes [12].

While the advantages of PCMs are clear, their effectiveness is heavily dependent on the

building fabric, weather conditions and their interaction with heating, ventilation and air-

conditioning equipment. Numerical modelling provides a way of establishing the

performance of PCMs over a wide range of conditions and operating modes. Various

commercial simulation tools incorporate phase change modelling capabilities such as ESP-r

[10]; TRNSYS [11]; and FLUENT, amongst others. However, these simulation tools do not

provide a flexible enough way to introduce custom individual melting and freezing

4
processes. The validation of the heat source/sink method used in this study to overcome

these limitations will be performed with FLUENT, even though it can also be applied using

other commercial CFD packages, via the governing energy equation.

Computational Fluid Dynamics (CFD) simulation environments usually possess a default

melting/solidification model but this does not provide the flexibility to vary the enthalpy-

temperature relationships and the possibility of introducing temperature hysteresis [4]. The

default enthalpy-porosity method used, developed from the Stefan Problem, does not

explicitly track the solid-liquid interface, but rather, a parameter known as the liquid

fraction (β), which indicates the fraction of liquid in a specific cell in the modelling domain

[5]. The liquid fraction allows the computation of the change in enthalpy from the energy in

the material during phase change as follows [6].

Eq. (1)

Eqs. (2)

Where;

Eq. (3)

The solver constantly iterates between Eq. (1) and (3) to determine the temperature of each

cell. In conduction dominant materials where advective movements within the liquid phase

are negligible, Eq. (1) can be simplified to:

5
(4)

The limitations of the default enthalpy-porosity model presented above are that the

calculation of the liquid fraction is done via the lever rule [6] (that is, the

enthalpy/temperature relationships are assumed linear), and that the solver assumes that

the melting and solidification enthalpy-temperature relationships are similar. Previous

observations show that these are not always the case, and can result in discrepancies

between experimental and numerical results [4, 7, 9].

This paper presents an approach by which the process of melting and solidification are

treated separately by using specific enthalpy-temperature relationships of melting and

solidification within the CFD model. This provides a better representation of the processes

involved and more accurate simulation results. This method is inspired by similar works

performed on evaporation/ condensation [8]. The concept relies on the addition of a heat

source term SE in equation (4) to mimic the melting and solidification process. The works in

ref. [8] focus on a single phase change temperature, incorporating source terms in the

energy and momentum equations to simulate phase change in pure water. This study, on

the other hand, focuses on phase change temperature ranges (mushy regions) and only

influences the energy equation.

Contrary to the methods in commercial software, such as FLUENT which uses the enthalpy-

porosity method, the method proposed here calculates the energy stored/released directly

from the temperature in the form of a user defined function (UDF). The liquid fraction is

calculated as an extra parameter, also from a UDF, but is not used in the simulation process.

6
This eliminates the dependency of the enthalpy function to obey the lever rule and provides

a more flexible approach.

This method may be considered as an extrapolation of the enthalpy-porosity method, even

though it uses a heat source/sink as opposed to the explicit enthalpy-temperature relation.

The reason is that the quantification of the heat source/sink is dependent on the energy or

enthalpy change during the phase change process. This, therefore, is a semi-empirical

method requiring the experimental enthalpy-temperature relation to be determined though

thermal techniques such as Differential Scanning Calorimetry (DSC) [4] or the T-history

method [13].

The approach presented in this study is valid for materials where advective movements

inside the PCM are negligible and the dominant heat transfer mode is conduction. This is the

case for PCMs impregnated in matrices such as plasterboard or tiles. In applications where

large quantities of PCM are used, for example PCM in large tank for thermal storage

applications, movement of the liquid phase will introduce heat transfer by convection. In

these cases, the source term in Eq. 1 must be modified to accommodate for the appropriate

convection effects.

7
2.0 Description of the model

A crucial aspect of the model is to differentiate between melting and freezing, so that the

solver uses the appropriate heat source function. Melting is an endothermic process, i.e.

absorbing heat as the temperature of the material increases, and freezing is an exothermic

process, releasing heat as the temperature decreases. As a result, melting will be mimicked

through a heat sink, while freezing as a heat source, with the corresponding change in

temperature, incorporated in the source term SE in Eq. (4).

Fig. 1 - Differences between the step function used in commercial software to describe melting and
freezing, actual data from DSC results, and data produced by the modelling approach proposed in
this paper.

It can be seen from Fig. 1 that the default CFD curve is a step function, approximating the

DSC curve. The default CFD curve aims at equalising the area between the onset and the

end points, and is not an accurate representation of the actual DSC curve. On the other

hand, the UDF (SE) is a more flexible, accurate representation of the DSC curve. By

8
appropriately adjusting the regression coefficient of the UDF polynomial curve, an exact

representation of the DSC can be obtained, depending on the material properties.

For melting, the simulation cell temperature at the previous time-step should be lower than

at the ‘actual’ time-step, and vice-versa for freezing. This concept is therefore used to

provide the solver with the required information to determine whether the material in a

specific cell is melting or freezing, and consequently follow the appropriate enthalpy-

temperature curve. Additionally, the function of the liquid fraction can be defined based on

the curvature of the integral of the UDF DSC curve, and the onset and end temperatures of

the phase change.

For completeness, it is not sufficient to define the heat source term (SE) only inside the

phase change temperature range; it should also be defined for all the conditions in Table 1.

(Note that Tsol < Tliq)

9
Heat sink conditions for Melting ( )

( Tt ≤ Tliq, m ) & ( Tt –1 ≥ Tsol, m ) SE, m = - Eq.(5)

( Tt > Tliq, m ) & ( Tt –1 < Tsol, m ) SE, m = - Eq.(6)

( Tsol, m ≤ Tt ≤ Tliq, m ) & ( Tt –1 < Tsol, m ) SE, m = - Eq.(7)

( Tt > Tliq, m ) & ( Tsol, m ≤ Tt –1 ≤ Tliq, m ) SE, m = - Eq.(8)

Heat source conditions for Freezing (Tt –1 > Tt )

( Tt ≥ Tsol, f ) & ( Tt –1 ≤ Tliq, f ) SE, f = - Eq.(9)

( Tt < Tsol, f ) & (Tt –1 > Tliq, f ) SE, f = - Eq.(10)

( Tsol, f ≤ Tt ≤ Tliq, f ) & ( Tt –1 > Tliq, f ) SE, f = - Eq.(11)

( Tt < Tsol, f ) & ( Tsol, f ≤ Tt –1 ≤ Tliq, f ) SE, f = - Eq.(12)

Liquid Fraction (β)Conditions

Melting & Freezing: ( Tsol ≤ Tt ≤ Tliq )

Table 1- Conditions required to fully defining the heat source/sink User Defined Source term. f(T)m and f(T)f are
the equations of the cumulative energy (J/kg) against temperature relationships from the UDF DSC curve in
Fig. 1, from the onset of melting and freezing, respectively.

3.0 Experimental Setup

The experimental setup consisted of macro-encapsulating the PCM material in a 100mm ×

70 mm × 80mm aluminium box as shown in Fig. 2. The box consisted of approximately 0.5 kg

of the composite PCM. The material investigated is a composite of low density polyethylene

and organic PCM, with phase change behaviour as shown in Fig. 4. The experiment was

conducted in an environmental chamber, with 4 T-type thermocouples (located as shown in

10
Fig. 2) and a PICO data logger, calibrated for the temperature range encountered in the

experiment.

Fig. 2 - Box with thermocouple locations

Fig. 3 - Experimental Air Temperature variations

11
Fig. 4 - DSC result of PCM composite

The shaded areas underneath the DSC curves represent the latent heat energy during phase

change used in this study. Note that freezing is slightly underestimated in order to portray

the effect of inaccurately choosing the phase change parameters (onset and end

temperatures and latent heat capacity). More details are given in section 4.0.

The air temperature of the chamber was varied from below the solidus temperature and

above the liquidus temperature as shown in Fig. 3, and the temperature at different points

in the composite was recorded at intervals of 5 s.

4.0 Model inputs

The physical and thermal properties of the composite PCM material shown in Table 2 were

obtained from the manufacturer.

12
Density 840 kg/m3

Specific heat capacity 2400 J/kgK

Thermal conductivity 0.3 W/mK

Table 2 - Manufacturer’s PCM properties

Following the DSC curves in Fig. 4, the cumulative energy for the PCM composite for melting

and freezing are shown in Fig. 5. It is important to note that DSC provides the heat flow

(mW) at each temperature for a specific sample mass, and that these values have to be

properly converted to energy (J/kg) in order to be compatible in this method.

Fig. 5(a) - Experimental, UDF & Default CFD cumulative energy-temperature curves for the melting
process

The cumulative energy curve shown in Fig. 5(a) (equivalent to the enthalpy/energy-

temperature relation) provides the basis for the melting process. The onset of melting is

taken to be 290 K (16.85 oC) and the end temperature 302 K (28.85 oC), with a latent heat

capacity of 44360 J/kg, based on the original DSC curves in Fig 4. The equation (J/kg) of the

UDF curve is given by:

f(T)m = -38.24T 3 + 34027.22T 2 - 10088202.17T + 996484631.17 (R2 = 0.992) - Eq. (13)

13
Fig. 5 (b) - Experimental, UDF & Default CFD cumulative energy-temperature curves for freezing
process

Similarly, the onset of freezing is taken to be 298.15 K (25.00 oC) and the end temperature at

289 K (15.85 oC), producing an average hysteresis of ≈ 2 K, based on the energy curves. The

latent heat capacity for freezing is 44117 J/kg. The equation of the UDF curve (J/kg) was

determined to be:

- f(T)f = - 0.10T 3 + 394.48T 2 - 201586.68T + 27601496.40 (R2 = 0.997) - Eq. (14)

The liquid fraction equations for melting and freezing respectively are:

βm = - 0.001 T 3 + 0.767 T 2 - 227.412 T + 22463.113 - Eq. (15)

βf = 0.007 T 2 - 3.769 T + 537.874 - Eq. (16)

5.0 Validation & Discussion

Validation of the model was done in two parts; melting and freezing. This is to provide

greater flexibility in the simulations, as well as to ease the investigation of the model in

regards to melting and freezing separately. The modelling results are obtained with a mesh-

14
converged model of 12,320 hexahedral elements, a converged time-step of 10s and all

simulations iteratively converged. The validation results are shown for a single point, in the

material domain, although the same trends and explanations apply to all other points.

5.1 Melting Process

For melting, the PCM box is initialised at 8oC and the air temperature gradually increases

with the same trend as in the experiment. A convective heat transfer coefficient of 5 W/m2K

between the box and the surrounding air is used for validation. The temperatures at the

four locations were monitored. The validation results for location A are shown in Fig. 6.

Fig. 6 - Experimental, UDF model and Default CFD model temperatures


of point A during the melting process

The results in Fig. 6 show the differences in using the different models in relation to the

experiment. The UDF model predicts the temperature trend at both points more accurately

than the default FLUENT model. From the energy-temperature curve in Fig. 5(a), the UDF

model produces a gradual energy change, as opposed to the default model where the

energy change is constant. Furthermore, the onset and the end temperatures of melting on

both models are accurate, confirming the reliability in the approximation of the DSC energy

15
curve. The difference towards the end of the simulation time (> 9000s) can be attributed to

an increase in specific heat capacity as the material changes from solid to liquid phase. This

change in cp was not incorporated in the models.

The default FLUENT model predicts lower temperatures for both locations during melting.

Close attention to the models’ results in Fig. 6 shows that if the initial temperature trends

during melting are flawed, then the error is carried into the simulation affecting results at

later times, providing incorrect system thermal dynamics.

For ease of comparison, the Root Mean Square (RMS) errors relative to the experimental

data were calculated for each location. For the case of melting, the UDF model produced an

error of 1.3%, while the Default CFD model produced a much higher error of 12.3%, and

therefore confirming the prediction accuracy of the UDF model.

5.2 Freezing Process

For freezing, the PCM box is initialised at 59oC and the air temperature gradually decreases

with the same trend as in the experiment. A convective heat transfer coefficient of 8 W/m2K

between the box and the surrounding air, and a specific heat capacity of 2550 J/kgK are used

for validation. The validation results for locations C are shown in Fig. 7.

16
Fig. 7 - Experimental, UDF and Default CFD models’ temperatures of
point C during the freezing process

The results from Fig. 7 show that the experimental approximation of the energy-

temperature curve in Fig. 5 (b) is not accurate enough to exactly predict the freezing

temperature trends in the composite. This is due to the fact that a slightly inaccurate

freezing onset temperature and energy-temperature curve were deliberately chosen from

the DSC data (Fig. 4). Therefore, the discrepancy at the initial stages of freezing in the UDF

model in Fig. 7, is due to a sharper gradient in the UDF energy-temperature curve, relative

to the DSC results, at the onset of freezing (298 K) in Fig. 5(b). The temperature trend in Fig.

7 of the UDF model is however a good representation of the UDF energy-temperature curve

input.

Conversely, the default CFD model predicts a temperature trend with an average error of

approximately 2oC over the freezing range. Thus, even with a relatively inaccurate selection

of the UDF enthalpy-temperature relationship, the UDF model provides a more accurate

prediction than the default model. In this case, the RMS error for the UDF model is 0.5%,

while the default model produces a much higher error of 6.3%, over the phase change

range.

The main observation from sections 5.1 and 5.2 is that the errors with the UDF model are

lower than the default CFD model for both the melting and freezing cases.

5.3 Validation Contour plots of UDF model for Melting and Freezing

Figs. 8 and 9 show the contour plots of the PCM composite during the melting and freezing

processes using the UDF model. Melting completes after 12000s while freezing is a much

17
slower process (due to slower reduction in the surrounding experimental air temperature),

and completes in more than 20000s. The phase change effects start at the corners and

edges due to higher heat transfer rates and progresses towards the bulk of the composite.

The contour results, depicted in Figs. 8 and 9 for both the temperatures and the liquid

fraction for melting and freezing, respectively, are in accordance with the user defined

functions. Figs. 8 and 9 show five plane sections of the entire 3-D PCM model simulated, and

confirm the validity of the UDF model in regards to visualisation of heat transfer. In this case

where the advective movements are neglected, the liquid fraction does not affect the

momentum of the liquid phase of the PCM. It is simply aids visualisation of the mushy parts.

Fig. 8 - Temperature and Liquid Fraction contours during the melting process

Fig. 9 - Temperature and Liquid Fraction contours during the freezing process

18
6.0 Application of heat source/sink model

PCMs are a growing technology used in buildings to provide passive thermal control of the

indoor environment, by introducing additional thermal inertia in the internal space to lower

the air temperature swing, and maintain thermal comfort. Commercial products include

BASF Micronal(r) PCM boards, DuPont Energain(r) and the EBB Clay PCM building boards. To

test the performance of such systems, numerical simulations are often employed, but as

shown in Figs. 6 and 7, using the default enthalpy-porosity models of phase change often

leads to errors in the dynamic temperature predictions. The following sub-sections will

portray the use of the heat source/sink model, in comparison to the default CFD model for

an average lightweight external wall.

The wall consists of 19mm render, 200mm concrete and 13mm plaster, and according to

CIBSE [14], has a decrement factor of 0.42 and decrement delay of 6.5 hrs. To this wall, a

20mm solid PCM layer was applied on the internal side, and the thermal response of the

wall examined with both the heat source/sink model and the default enthalpy-porosity

model. The walls were investigated in FLUENT by applying a sinusoidal air temperature

excitation: varying from peak temperatures of 273.15K and 318.15K, over a period of 24

hours, with heat transfer coefficient of 7.7 W/m2K on the internal surface [14], and the

external surfaces maintained as adiabatic, as shown in Fig. 10.

19
Fig. 10 – Simulation Wall Description

The PCM used for the validation of the heat source/sink model (Fig. 4) is used for this non-

linear energy-temperature relationship investigation. The PCM thermal properties can be

obtained from Figs. 4 and 5, and Table 2. Because of the inability to specify separate melting

and freezing temperatures and enthalpies in the default model, the solidus temperature

was taken to be 289.5 K, the liquidus temperature as 300.1 K and the latent heat capacity as

44240 J/kg, calculated as the average values for both melting and freezing.

Fig. 11 shows the temperature variation with time at the midpoints of each wall layer,

predicted by the default method in the CFD software and from the User Defined Model, for

one air temperature cycle.

20
Fig. 11(a) – PCM temperature development Fig. 11(b) – Plaster temperature development

Fig. 11(c) – Concrete temperature development Fig. 11(d) – Render temperature development
Fig. 11 – Simulated temperature trends at midpoints of simulation wall for validation PCM

Figs. 11 (a) – (d) show that the thermal dynamics of the entire wall is affected by the choice

of the phase change model. The phase change temperature trend in the PCM segment of

the wall (Fig. 11a) are in accordance with the individual model (default and UDF) enthalpy-

temperature curves, but the thermal dynamics of the entire wall is very different for each

model, with peak temperature differences of approximately 4 – 5 K.

As the default model assumes a higher enthalpy change at the beginning of melting (Fig. 5a),

the temperatures in the UDF model (Fig. 11a) are higher, due to the lower latent heat

storage capacity of the PCM. Subsequently, the onset of freezing is clearly shown in Fig.

11(a) at t ≈ 70000s by the change in temperature reduction rate for the UDF model, in

relation to the gradual change in the default model. It is noticed that the initial temperature

trends in the wall are very important in the subsequent development of temperature along

21
the materials. In this regard, the curvatures of the enthalpy-temperature relationships are

very important, as they dictate both the initial and subsequent development of temperature

in not only the PCM layer, but in the entire wall.

22
7.0 Conclusion

The results in the paper show that the heat source/sink method developed is suitable for

the prediction of phase change phenomena and offers more accurate representation of the

temperature during phase change processes compared to the default enthalpy-porosity

method used by many commercial software packages. The importance of energy variation

with temperature during phase change and its direct effect to the solution dynamics in a

wall are portrayed. Furthermore, because the individual melting and freezing effects as well

as hysteresis are incorporated in the same code, this method allows accurate simulation of

the PCM behaviour under cyclic conditions imposed by the variation in external ambient

temperature.

As CFD simulations require detailed inputs, this semi-empirical model is also heavily

dependent on the thermal properties of the PCM, especially the temperatures at the start

and end of the melting and freezing processes, the energy-temperature relations as well as

the specific heat-temperature relations. Variation of the thermal conductivity of the PCM

with temperature may also have an influence if significant changes in its value take place

during phase change. Hence, the input to the simulation is crucial, and in this study, a

minimum regression coefficient (R2) of 0.992 is used in the development of correlations for

the UDF from experimental data.

The higher the hysteresis of the PCM the more pronounced will be the advantages of the

proposed methodology over the default enthalpy-porosity method employed by

commercial software packages.

8.0 Acknowledgements

23
This work was made possible through sponsorship from the Engineering and Physical

Sciences Research Council (EPSRC) of the UK, Grant No: EP/H004181/1.

9.0 References

[1] L. Shilei, F. Guohui, Z. Neng, D. Li, Experimental study and evaluation of latent heat

storage in phase change materials wallboards, Energy and Buildings 39(2007): pp 1088 -

1091

[2] F. Kuznik, J. Virgone, Jean Noel, Optimization of a phase change material wallboard for

building use, Applied Thermal Engineering 29(2008): pp 1291-1298

[3] F. Kuznik, J. Virgone, J.J Roux Energetic efficiency of a room wall containing PCM

wallboard: A full scale experimental investigation, Energy and Buildings, 40(2008): pp 148-

156

[4] G. Susman, Z. Dehouche, T. Cheechern, S. Craig, Tests of prototype PCM ‘sails’ for office

cooling, Applied Thermal Engineering 31(2010): pp 717-726

[5] V.R Voller, L Shadabi, Enthalpy method for tracking a phase change boundary in two

dimensions, International Communications in Heat Mass Transfer 11(1984): pp 239-249

[6] ANSYS FLUENT Theory Guide (Release 13.0, November 2010), Solidification and melting,

Ch 18: pp 569 -576

[7] P.W Egolf, H Manz, Theory and modelling of phase change materials with and without

mush regions, International Journal of Heat and Mass Transfer 37(1994): pp 2917 – 2624

24
[8] S.C.K De Schepper, G.J Heynderickx, G.B Marin, Modeling the evaporation of a

hydrocarbon feedstock in the convection section of a steam cracker, Computers and

Chemical Engineering 33 (2009): pp 122–132

[9] F. Kuznik, J. Virgone, Experimental investigation of wallboard containing phase change

material: Data for validation of numerical modelling, Energy and Buildings 41 (2009): pp

561–570

[10] D. Heim and J. Clarke, Numerical modelling and thermal simulation of phase change

materials within ESP-r, Eighth International IBPSA Conference, Eindhoven, Netherlands,

August 11-14, 2003

[11] J. Bony, S Citherlet , Numerical model and experimental validation of heat storage with

phase change materials, Energy and Buildings 39 (2007): pp 1065–1072

[12] I. Ceróna, J. Neila, M. Khayet , Experimental tile with phase change materials (PCM) for

building use, Energy and Buildings 43 (2011): pp 1869–1874

[13] A. Lazaro, E. Gunther, H. Mehling, S. hiebler, JM. Marin, B. Zalba, Verification of a T-

history installation to measure enthalpy versus temperature curves of phase change

materials, Measurement Science and Technology, 17(2006): pp. 2168-2174

[14] Chartered Institute of Building Services Engineers (CIBSE), Guide A (2006): pp. 3-48

25

You might also like