Next Article in Journal
Interpretability Analysis of Convolutional Neural Networks for Crack Detection
Previous Article in Journal
An Environmental Quality Assessment of Office Buildings: The Impact of a Glass Façade on Users in Different Climate Contexts in Croatia
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Experimental Thermal Conductivity Measurement of Hollow-Structured Polypropylene Material by DTC-25 and Hot Box Test

Mechanical and Aerospace Engineering, University of Missouri, Columbia, MO 65211, USA
*
Author to whom correspondence should be addressed.
Buildings 2023, 13(12), 3094; https://doi.org/10.3390/buildings13123094
Submission received: 20 November 2023 / Revised: 10 December 2023 / Accepted: 11 December 2023 / Published: 13 December 2023
(This article belongs to the Section Building Materials, and Repair & Renovation)

Abstract

:
In this study, experimental measurements of porous polypropylene (PP) using the DTC-25 TA laboratory equipment and a hot box test are compared. The thermal conductivity of building materials indicates their insulation capability. Excellent building materials will have a lower thermal conductivity value as well as other building insulator performance metrics. While results show that increasing the volume fraction of fluid in porous PP has an inverse association with the thermal conductivity of the material, as predicted by porous media theories, there is a marked difference in the measured values of thermal conductivity using the two methods. The thermal conductivity values of porous PP from the DTC-25 and the hot box test were 0.21 and 0.0033 W/mK, respectively. The difference in the thermal conductivity values was due to the misapplication of the Fourier’s guarded heat flow model in the DTC-25 device to a convective fluid porous medium.

1. Introduction

In the last decade, attention toward energy and the environment has grown extensively and many international and national policies have been developed to achieve a more sustainable future for the environment while also focusing on reducing the cost of energy. Across the world, buildings seem to consume the most energy. In Europe, 40% of the total energy utilization is attributed to the building sector [1]. The United States Energy Information Administration (EIA) estimated that about 39% (or about 38 quadrillion British thermal units) of the total US energy consumption was consumed by the residential and commercial building sector in 2017 [2]. Residential buildings have more than a 50 percent share of the total building energy utilization in the United States.
One way to reduce this vast amount of energy utilization in buildings, which will lower the cost of energy utilization and reduce the greenhouse effect [3,4,5], is to engage in energy-saving practices. Energy savings are implemented through energy efficiency practices and through the judicious use of the energy available to our residential building envelope. Since a considerable amount of energy consumption in residential buildings is expended in HVAC systems [6,7], windows and door insulation [8,9], thermal bridge losses [10] and external walls [11,12], any design or material improvement in terms of energy utilization would lead to more energy-efficient housing.
The development of external walls has greatly contributed to significant energy savings and presents research opportunities in the development of new and advanced insulation materials. In future buildings, external wall insulation should guarantee optimum performance throughout the entire life cycle of the building. Other non-thermal considerations when choosing insulation materials are the environmental safety of the material [13], fire resistivity [14], sound shielding [15,16], mechanical impact tolerance and cost effectiveness [17].
While inroads have been made in newer innovative building insulation such as vacuum insulation panels (VIPs), gas-filled panels (GFPs), phase change materials (PCMs), dynamic insulators, etc., and in meeting the goal of saving more energy than the conventional insulating materials such as glass wool, cellulose, polystyrene etc., conventional insulation is still very popular in old and recent residential buildings due to the high cost and advanced manufacturing processing of innovative insulating materials [18].
This work explored alternative material designs that would increase energy savings. We have selected the material polypropylene because it potentially provides a mix of good insulation, impact resistance and recyclability. Polypropylene exhibits good building insulator characteristics, as outlined by Al-Homoud [19]. Polypropylene board insulation has the edge over conventional insulations like blanket and batt, blown and spray in its eco-friendliness, and it poses no harm to humans and animals. Its high impact resistance [20,21] has the potential to enhance the hail storm resistance of buildings.
A multi-hollowed polypropylene plastic is manufactured through 3D printing to create a foam-like macrostructure. The efficient thermal insulation of polar bears’ hair [22,23,24,25] is, in part, explained by their hollow macrostructure. Hollow hair reduces thermal conductivity in severe cold weather, thereby enhancing thermal insulation. Dohrn et al. [26] note that the insulation capacity of the polyurethane (PUR) form is primarily due to the gases contained in the foam, which significantly improve the thermal insulation associated with vacuum insulations.
Effective thermal conductivity represents the mean thermal conductivity value for a composite material or porous medium. Effective thermal conductivity is an important heat transfer property of materials. Kulkani and Doraiswamy [27] found that the effective thermal conductivity of a porous material is influenced by the thermal conductivities of the solid, gas, porosity, emissivity of the particles’ surface and temperature. Similarly, other research studies [28,29,30] have established that effective thermal conductivity is determined by factors such as the properties of the particles and matrix, and the microstructure of a composite polymer. Effective thermal conductivity can be calculated analytically [31,32], assessed through numerical simulation [33,34] or measured experimentally [35,36]. In this paper, an experimental method was used to measure the effective thermal conductivity of multi-hollowed polypropylene board using the DTC-25 and hot box testing. The thermal conductivity measurements of samples were obtained by means of the DTC-25 conductivity meter from TA Instruments, which uses the guarded heat flow method to obtain thermal conductivity measurements [37]. The hot box test follows the ASTM C1046 [38] standard practice for in situ measurement of heat flux and temperature in building envelope components. This method establishes the heat flux measurements through localized materials and between the hot box’s internal and external environmental temperatures using heat flux sensors and temperature transducers. While the hot box test method can provide accurate measurements of the thermal conductivity of building materials in real-world usage [39], it requires an understanding of a complex data-logging tool for post-test data processing.

2. Theory

Wang et al. [40] define a hollow structure as a solid structure with voids. The existence of voids breaks the continuity of heat transport pathways in hollow-structured materials [41]. Figure 1 shows the morphology of a hollow-structured material with three immiscible phases. Phase I is a solid for continuous heat transport pathway. Phase II is the void. Phase III is an interfacial bridge which could be a solid component. The multi-hollow polypropylene polymer was modeled as a pore structure. Pore defects in polymer composites are capable of lowering the effective thermal conductivities of polymers [28] and improving their insulation properties.
Many researchers have developed models to explain the effective thermal conductivity of porous materials [36,42,43,44,45,46]. The multi-hollow polypropylene polymer was modeled as a pore structure. Pore defects in polymer composites are capable of lowering the effective thermal conductivities of polymers [28] and improving their insulation properties. Wang and Pan [47] reviewed the effective thermal conductivity models that are applied to multiphase materials according to the size and structure of the pores. The basic structural model includes the series, parallel, Maxwell–Eucken and effective medium theory (EMT) models shown in Equations (1)–(4), respectively. The series and parallel models assume the simplest structure with heat flow. In the series model, heat flows across the pore, while the parallel model assumes heat from along the porous media. The Maxwell–Eucken model proposes a spherical dispersed phase in a continuous solid phase with no contact. The EMT model supposes that a void or filler is surrounded by a homogeneous, effective medium.
k e = 1 ( 1 θ 2 ) / k I + θ 2 / k 2
k e = k 1 1 θ 2 + k 2 θ 2
k e = k 1 2 k 1 + k 2 2 k 1 k 2 θ 2 2 k 1 + k 2 + k 1 k 2 θ 2
1 θ 2 k 1 k e k 1 + 2 k e + θ 2 k 2 k e k 2 + 2 k e = 0
where k e is the effective thermal conductivity of the porous medium, k 1 and θ 1 are the thermal conductivity and phase volume of the solid medium, respectively, and k 2 and θ 2 are, respectively, the thermal conductivity and volume fraction of the fluid medium.
Utilizing the local thermal equilibrium hypothesis, which supposes that the temperature (T) is the same for both the solid and the fluid phase, the heat transfer equation is derived from the solid and fluid energy mixture. The energy quantity present in the fluid and solid phases is given by Equations (5) and (6).
( ρ C P ) e T t + ρ f C P . f u T + . q = Q
q = k e T
where ( ρ C P ) e = θ s ρ s C p , s + ε p ρ f C p . f
ρ f = f l u i d   d e n s i t y
C P . f = f l u i d   h e a t   c a p a c i t y   a t   c o n t a n t   p r e s s u r e
( ρ C P ) e = e f f e c t i v e   v o l u m e t r i c   h e a t   c a p a c i t y   a t   c o n s t a n t   p r e s s u r e
θ s = s o l i d   m a t r i x   v o l u m e   f r a c t i o n
ρ s = s o l i d   m a t r i x   d e n s i t y
C p , s = s o l i d   m a t r i x   h e a t   c a p a c i t y
k e = e f f e c t i v e   t h e r m a l   c o n d u c t i v i t y
q = c o n d u c t i v e   h e a t   f l u x
u = v e l o c i t y   f i e l d ,   interpreted as Darcy velocity
Q = h e a t   s o u r c e   o r   s i n k
The term ( ρ C P ) e T t vanishes for a steady state problem. Since the heat flows in series across the porous medium, with all the heat flux passing through both solid and fluid, the effective thermal conductivity refers to the weighted harmonic mean of conductivities k f and k s .
1 k e = θ s k s   +   ε p k f
Equation (7) provides a lower bound for the effective thermal conductivity of porous media. Equation (7) is a modification of Equation (1), which considers the impact of porosity and convection in the fluid moving through porous media.
The experiment performed by Liang [36] to measure the thermal conductivity of polypropylene–hollow glass bead composites underscore the fact that the effective thermal conductivity decreases linearly with the increase in the volume fraction of hollow glass beads (HGBs). Other materials also display similar behavior. For example, Nait-Ali et al. [48] determined the thermal conductivity of highly porous zirconia ceramic with nanometric-sized grains and found that its effective thermal conductivity decreases as the pore volume fraction increases. The finding of Liang’s experiment demonstrates that creating a void in polymer composites could decrease the thermal conductivity of the polymer concerned. Figure 2 shows the correlation between effective thermal conductivity ( K e f f ) and the volume fraction ( ϕ f ) of a polypropylene–hollow glass bead (HGB) composite (PP/TK35) supplied by the Molus company in Germany [36]. From the plot, it can be observed that the effective thermal conductivity K e f f and volume fraction ϕ f of HGB show an inverse relationship.

3. Methodology

3.1. Experimental Methods

We measured the effective thermal conductivity of multi-hollow polypropylene board by employing the discovery thermal conductivity tester (DTC-25) manufactured by TA Instruments and the hot box built in Dr. Sanjeev’s laboratory at the University of Missouri. DTC-25 is suited for quality control and accurately determining the thermal conductivity of polymers, metals, ceramics, composites, glass, rubber and graphite products in the laboratory.
The hot box test follows the ASTM C1046 standard [38] practice for in situ measurement of heat flux and temperature in building envelope components. This hot box is a more robust measurement of the thermal conductivity of building materials as it closely reflects reality in application. The method involves measuring heat flux through localized materials and the hot box’s internal/external environmental temperatures with heat flux transducers and temperature transducers, respectively. Heat flux is the heat energy transfer rate through a given material. It is defined as the energy flow per unit area per unit of time. The SI unit is given by watts per square meter (W/m2). The heat flux determination through the PP insulation board using the hot box test provides a basis for the calculation of the thermal conductivity of the porous PP board at steady state, via the formula of Fourier’s law of heat transfer given in Equation (8).
q = k e T
q = k e Δ T Δ x
q = h e a t   f l u x
k e = e f f e c t i v e   t h e r m a l   c o n d u c t i v i t y
T = v e c t o r   d i f f e r e n t i a l   o p e r a t o r   o n   t e m p e r a t u r e
Δ T = c h a n g e   i n   t e m p e r a t u r e
Δ x = t h i c k n e s s

3.2. Method for DTC-25 Test

The three-step procedure in the DTC-25 test includes sample preparation, sample loading and an experimental test. We prepared a multi-hollow polypropylene sample measuring 50 mm in diameter and 10 mm in thickness for the experimental measurement of thermal conductivity. The sample (Figure 3) was cut away from the 3D-printed, hollow-structured PP block shown in Figure 4. Three different porous PP samples with hole sizes of 4 mm, 2 mm and 1 mm were produced for this experimental test. It is necessary that test samples are smooth and flat on both surfaces. Before beginning the experimental run, a heat sink compound or thermal paste should be applied to both surfaces to maximize heat transfer and dissipation between the test stack carrying heat and the polypropylene board. Silicone paste was used as a thermal paste for this test. The sample is loaded into the test stack with the aid of a free/move switch on the electronic enclosure, which allows for the up/down movement of the stack rod. The test pressure should be kept between 40 and 60 psi, ensuring that the sample is in line with the top and bottom sample plate before it is locked for testing.
The temperature controller setpoint should be maintained at 45 °C and should be adjusted whenever it deviates from the required temperature. Experimental reading takes 30 min, which allows the voltages on the LCD display to reach equilibrium. The thermistor output voltages taken for experimental analysis included reference voltage, upper voltage, lower voltage and heat sink voltage. The voltages were converted to temperature in degrees Celsius with help from computer software.

3.3. Method for Hot Box Test

Thermal conductivity is known to be affected by thermal bridging, changes in material property with temperature, airflow in and through the building envelope, internal convection, phase changes and the hygric storage effect [49], especially in contemporary insulation materials and how they control heat flow. To obtain a more accurate predictive thermal conductivity of a material, a physical facility is needed to replicate field usage conditions. A hot box was deployed for the experimental measurement of the hollow-structured PP’s thermal conductivity. A hot box apparatus is generally large and expensive [49,50]. A few of them can be found in the US National Laboratories. The hot box utilized was constructed in Professor Sanjeev Khanna’s laboratory at the University of Missouri by a group of students in 2015. Figure 5 shows the constructed hot box for measuring the heat flux through building materials. It measures 50 in length by 50 in width by 75 in height and is completely insulated except for two metering windows, which allow for the measurement of heat flux passing through materials. Two incandescent light sources provide the heating for the enclosure. A datalogger, CR1000, manufactured by Campbell Scientific, provides a connection for measuring the heat flux and the hot box’s internal and external environmental temperature. A datalogger software, PC400, accompanies the RC1000 for post-processing recorded measurements.
The hot box is well insulated with two windows measuring 14 in by 14 in on both sides of the building envelope enclosure. An extruded polystyrene (XPS) insulator with a known thermal conductivity is placed on one window while the other window is covered with porous PP material. We are able to establish the PP’s thermal conductivity if the sensors and data-logging system correctly determined the known thermal conductivity value of the XPS insulator. Cai et al. [51] reviewed the thermal conductivities of XPS to be 0.026–0.040 W/mK. These thermal conductivity values were confirmed in a previous technical report on the comparison of expanded and extruded foam polystyrene insulation in roadway and airport embankments [52].
One HFP01 sensor is attached to the outer surface of the XPS insulator. Similarly, the other HFP01 sensor is attached to the outer surface of the PP’s insulation. The Campbell Scientific datalogger system with connected heat flux and temperature sensors was used to log the data for 5 h.
Circling back to Equation (8), XPS and hollow-structured PP thermal conductivity can be calculated when the heat flux is measured using the sensors. Figure 6 shows a schematic of the experimental setup. Figure 7 illustrates the physical test showing the Campbell Scientific datalogger system.

4. Results and Discussion

For DTC-25, the thermistor output voltages are converted to temperature using a DTC-25 computer software with an algorithm to calculate the thermal conductivity of porous PP. The results are tabulated in Table 1. These values compare to the porous media models described by Wang and Pan [47], showing that an increase in the volume fraction of the material lowers its thermal conductivity. A graph of thermal conductivity of porous PP versus the size of the holes in the material is shown in Figure 8.
To use the hot box, processing of the sensors’ measurements in the datalogger is carried out in the Campbell Scientific PC400 software. The datalogger logs a measurement every 5 s. A snapshot of the sensors’ measurements is shown in Figure 9 after 5 h.
From the results, the heat flux through the PP board is −48.72 W/m2K while the heat flux through the extruded polystyrene is −20.15 W/m2K. The negative values of the heat flux indicate that heat flows from a high to low temperature. The hot box’s interior temperature reading is 142.18 F (61.21 °C), and the average outer temperature reading is 73.35 F (22.97 °C). The PP’s board thickness is 50 mm with a pore size of 4 mm. Using Fourier’s equation of conduction heat transfer in Equation (8), the thermal conductivity of the XPS specimen can be deduced. Figure 10 illustrates the heat flow through the XPS insulation board. In this case, the measured heat transfer is the heat conduction through both the solid and fluid phases. As noted in various publications on heat transfer through metal and insulating foams [53,54,55,56,57], convection and radiation heat transfer are considered negligible to the extent that conduction is the only heat transfer mechanism.
Given:
k e = q Δ x Δ T
Thermal conductivity k of XPS = 20.15   ×   0.05 38.24   = 0.0263 W/mK.
The thermal conductivity value of the extruded polystyrene (XPS) in the hot box test corresponds to the reviewed values in Cai et al. [51] and Connor [52].
With some innovative porous media, the effect of convective heat transfer plays a crucial role in the overall heat transfer through the media, i.e., the total heat transfer is a summation of conductive heat transfer through the solid and convective heat transfer through the pores. Such materials must surpass a critical Rayleigh number for convective heat transfer to show any influence. The Rayleigh number is a dimensionless number that assesses how the effects of buoyancy forces and viscosity forces compare to conductive heat transfer. Harold Jeffreys [58] first calculated the critical Rayleigh number for a layer of fluid to be 1708 in the year 1926 in his widely published original article titled “The stability of a layer of fluid heated below”. Determination of the Rayleigh number will indicate whether the heat transfer mode in the pore of the porous PP is conduction or convection. Equation (9) is a depiction of the Rayleigh number R a problem highlighted in Nygard and Tyvand [59,60,61]. A graph of pore size against the Rayleigh number is shown in Figure 11.
R a = g β Κ Δ Τ h υ κ
R a = R a y l e i g h   n u m b e r
g = a c c e l e r a t i o n   d u e   t o   g r a v i t y
β = e x p a n s i o n   c o e f f i c i e n t   o f   f l u i d
Κ = p e r m e a b i l i t y
Δ Τ = c h a n g e   i n   p o r e   t e m p e r a t u r e
h = c l a s s i c a l   l e n g h t   o f   p o r e
v = t h e r m a l   d i f f u s i v i t y   o f   f l u i d
k = k i n e m a t i c   v i s c o s i t y   o f   f l u i d
R a = 1.5167 × 10 6 implies that the critical Rayleigh number is exceeded, and convection must take precedence in the porous PP. The analysis infers that a two-dimensional heat transfer mode, conduction and convection, is present in the PP porous media if we consider the effect of radiative heat transfer negligible. Following this, Fourier’s conduction heat transfer equation is insufficient to model the total heat transfer through the porous PP media to determine the thermal conductivity of the hollow-structured material. The guarded heat flow technique utilizes the Fourier steady heat transfer model to measure the thermal conductivity of materials in many laboratory apparatuses such as DTC-25. With porous media, a critical Rayleigh number for the fluid media must be determined to understand if the heat flow in the pores is conduction or convection. There is a potential measurement error in determining the thermal conductivity of porous media when the total heat flow through the material is a combination of conduction and convection. A more accurate model for such porous media was put forward by Adrian Bejan [62,63]. The Bejan model demonstrates how the heat flux q relates to the Rayleigh number R a in the porous medium shown in Equation (10).
q k e Δ T = 0.319 R a 1 / 2
where:
q = 48.72   W / m 2 K
Δ T = 38.24   K
R a = 1.5167 × 10 6
The effective thermal conductivity k e of the porous PP can be deduced from the above equation and determined to be 0.00325   W / m K . This effective thermal conductivity value of porous PP is about eight times less than the thermal conductivity of conventional building insulators. Figure 12 compares experimental thermal conductivity tests using DTC-25 and the hot box.
There is a marked difference in the effective thermal conductivity measurement of porous PP between the DTC-25 and the hot box test. The hot box thermal conductivity measurement is 70 times lower than DTC-25. This difference is due to the exclusion of convective heat transfer in the porous PP, when it is in fact present when using the DTC-25. The guarded heat flow method that considers only conductive heat flow through the solid and fluid phase is the operational principle of the DTC-25, whereas the hot box allows us to consider the influence of convective heat flow through the porous PP medium, as demonstrated in our approach. Selected building insulators’ thermal conductivities are compared with the thermal conductivity of our porous PP in Table 2. In Figure 13, we created a bar chart of the table for better data visualization.

5. Conclusions

Experimental measurements of porous PP using the DTC-25 laboratory equipment from TA Instruments and a hot box test have been compared. The thermal conductivity of building materials indicates their insulation capability. Excellent building materials will have a lower thermal conductivity value as well as other building insulator performance metrics. While results show that increasing the volume fraction of fluid in the porous PP has an inverse association with the thermal conductivity of the material, as predicted by porous media theories [47], there is a marked difference in the measured values of thermal conductivity using the two methods. The DTC-25 test utilizes Fourier’s guarded heat flow model, which assumes conductive heat transfer throughout the multiphase material. The assessment of the Rayleigh number of the fluid phase confirmed that a critical Rayleigh number has been exceeded. When the Rayleigh number is below a critical value for a fluid, there is no flow of heat and heat is transferred purely by conduction, which is sufficient for the operation of the DTC-25 device. A deviation from the pure conductive heat model is attained beyond a critical Rayleigh number, which triggers heat transfer by natural convection. Our porous PP medium is an example of a material that has a high Rayleigh number. An accurate model developed by the acclaimed authority on thermodynamics and heat transfer, Adrian Bejan, to predict a combined heat transfer by conduction and convection in porous PP was employed in the studies. The combined heat flow model was applied in the hot box test. The thermal conductivity value of the porous PP derived from the heat flux measurement in the hot box experiment was 0.0033 W/mK. This value exceeds the performance of conventional building insulators compared in Figure 13. The insulating metric exceeds the performance of traditional building insulators by a multiple of eight and compares favorably with the most innovative insulators, vacuum insulation panels (VIPs). Porous PP materials exhibit superior thermal insulation attributes due to the presence of a convective porous structure that inhibits the direct flow of heat across the porous PP board. The results demonstrate that porous PP has an efficient thermal insulation capability. There is the potential to use hollow-structured polypropylene as insulation for buildings’ external walls or exterior cladding to enhance building insulation, especially for renovated homes. This could save more energy cost than conventional insulators and is a more environmentally friendly option.
Future work should focus on a scaled production of 3D-printed, multi-hollow-structured polypropylene board, or manufacturing techniques for the production of multi-hollow polypropylene. The thermal conductivities of other insulators with multi-hollow structures should be investigated using the combined DTC-25 and hot box method to determine the impact of the Rayleigh number on their thermal conductivities.

Author Contributions

Conceptualization, O.O. and S.K.; methodology, O.O.; software, O.O.; validation, O.O.; formal analysis, O.O.; investigation, O.O.; resources, O.O. and S.K.; writing—original draft preparation, O.O.; writing—review and editing, S.K.; supervision, S.K.; All authors have read and agreed to the published version of the manuscript.

Funding

This research received no external funding.

Data Availability Statement

Data is contained within the article.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Schiavoni, S.; D’Alessandro, F.; Bianchi, F.; Asdrubali, F. Insulation Materials for the Building Sector: A Review and Comparative Analysis. Renew. Sustain. Energy Rev. 2016, 62, 988–1011. [Google Scholar] [CrossRef]
  2. U.S. Energy Information Administration. Monthly Energy Review—April 2018. 2018, 244, 30. Washington DC, USA. DOE/EIA-0035(2018/4). Available online: http://large.stanford.edu/courses/2018/ph241/green1/docs/doe-eia-0035-2018-4.pdf (accessed on 10 December 2021).
  3. Khan, M.Z.A. Causes and Consequences of Greenhouse Effect & Its Catastrophic Problems for Earth. Int. J. Sustain. Manag. Inf. Technol. 2017, 3, 34. [Google Scholar] [CrossRef]
  4. Santamaría, J.; Campano, M.A.; Girón, S. Method for the Economic Profitability of Energy Rehabilitation Operations: Application to Residential Dwellings in Seville. Procedia Comput. Sci. 2016, 83, 742–749. [Google Scholar] [CrossRef]
  5. Sierra-Pérez, J.; Boschmonart-Rives, J.; Gabarrell, X. Environmental Assessment of Façade-Building Systems and Thermal Insulation Materials for Different Climatic Conditions. J. Clean. Prod. 2016, 113, 102–113. [Google Scholar] [CrossRef]
  6. Abu Bakar, N.N.; Hassan, M.Y.; Abdullah, H.; Rahman, H.A.; Abdullah, M.P.; Hussin, F.; Bandi, M. Energy Efficiency Index as an Indicator for Measuring Building Energy Performance: A Review. Renew. Sustain. Energy Rev. 2015, 44, 1–11. [Google Scholar] [CrossRef]
  7. Kassai, M. Prediction of the HVAC Energy Demand and Consumption of a Single Family House with Different Calculation Methods. Energy Procedia 2017, 112, 585–594. [Google Scholar] [CrossRef]
  8. Baldinelli, G.; Asdrubali, F.; Baldassarri, C.; Bianchi, F.; D’Alessandro, F.; Schiavoni, S.; Basilicata, C. Energy and Environmental Performance Optimization of a Wooden Window: A Holistic Approach. Energy Build. 2014, 79, 114–131. [Google Scholar] [CrossRef]
  9. Gugliermetti, F.; Bisegna, F. Saving Energy in Residential Buildings: The Use of Fully Reversible Windows. Energy 2007, 32, 1235–1247. [Google Scholar] [CrossRef]
  10. Ascione, F.; Bianco, N.; De Masi, R.F.; Mauro, G.M.; Musto, M.; Vanoli, G.P. Experimental Validation of a Numerical Code by Thin Film Heat Flux Sensors for the Resolution of Thermal Bridges in Dynamic Conditions. Appl. Energy 2014, 124, 213–222. [Google Scholar] [CrossRef]
  11. Asdrubali, F.; D’Alessandro, F.; Baldinelli, G.; Bianchi, F. Evaluating in Situ Thermal Transmittance of Green Buildings Masonries—A Case Study. Case Stud. Constr. Mater. 2014, 1, 53–59. [Google Scholar] [CrossRef]
  12. Sim, J.; Sim, J. The Effect of External Walls on Energy Performance of a Korean Traditional Building. Sustain. Cities Soc. 2016, 24, 10–19. [Google Scholar] [CrossRef]
  13. Densley Tingley, D.; Hathway, A.; Davison, B. An Environmental Impact Comparison of External Wall Insulation Types. Build. Environ. 2015, 85, 182–189. [Google Scholar] [CrossRef]
  14. Iringova, A. Revitalisation of External Walls in Listed Buildings in the Context of Fire Protection. Procedia Eng. 2017, 195, 163–170. [Google Scholar] [CrossRef]
  15. Galbrun, L.; Scerri, L. Sound Insulation of Lightweight Extensive Green Roofs. Build. Environ. 2017, 116, 130–139. [Google Scholar] [CrossRef]
  16. Hongisto, V.; Mäkilä, M.; Suokas, M. Satisfaction with Sound Insulation in Residential Dwellings—The Effect of Wall Construction. Build. Environ. 2015, 85, 309–320. [Google Scholar] [CrossRef]
  17. INCA-Technical-Guide-02-Impact-Resistance-of-EWI-Systems.Pdf. Available online: https://jrmpl2t1z5gzqpmv2vjf0mqh-wpengine.netdna-ssl.com/wp-content/uploads/2017/07/INCA-Technical-Guide-02-Impact-Resistance-of-EWI-Systems.pdf (accessed on 10 December 2021).
  18. Aerogel Insulation: The Materials Science of Empty Space. Energy.gov. Available online: https://www.energy.gov/eere/amo/articles/aerogel-insulation-materials-science-empty-space (accessed on 10 December 2021).
  19. Al-Homoud, M.S. Performance Characteristics and Practical Applications of Common Building Thermal Insulation Materials. Build. Environ. 2005, 40, 353–366. [Google Scholar] [CrossRef]
  20. Furlan, L.G.; Ferreira, C.I.; Dal Castel, C.; Santos, K.S.; Mello, A.C.E.; Liberman, S.A.; Oviedo, M.A.S.; Mauler, R.S. Effect of Processing Conditions on the Mechanical and Thermal Properties of High-Impact Polypropylene Nanocomposites. Mater. Sci. Eng. A 2011, 528, 6715–6718. [Google Scholar] [CrossRef]
  21. Kozderka, M.; Rose, B.; Bahlouli, N.; Kočí, V.; Caillaud, E. Recycled High Impact Polypropylene in the Automotive Industry—Mechanical and Environmental Properties. Int. J. Interact. Des. Manuf. IJIDeM 2017, 11, 737–750. [Google Scholar] [CrossRef]
  22. Jia, H.; Guo, J.; Zhu, J. Comparison of the Photo-Thermal Energy Conversion Behavior of Polar Bear Hair and Wool of Sheep. J. Bionic Eng. 2017, 14, 616–621. [Google Scholar] [CrossRef]
  23. Simonis, P.; Ratta, M.; Oualim, E.M.; Mouhse, A.; Vigneron, J.-P. Radiative Contribution to Thermal Conductance in Animal Furs and Other Woolly Insulators. Opt. Express 2014, 22, 1940–1951. [Google Scholar] [CrossRef]
  24. Metwally, S.; Martínez Comesaña, S.; Zarzyka, M.; Szewczyk, P.K.; Karbowniczek, J.E.; Stachewicz, U. Thermal Insulation Design Bioinspired by Microstructure Study of Penguin Feather and Polar Bear Hair. Acta Biomater. 2019, 91, 270–283. [Google Scholar] [CrossRef]
  25. Cui, Y.; Gong, H.; Wang, Y.; Li, D.; Bai, H. A Thermally Insulating Textile Inspired by Polar Bear Hair. Adv. Mater. 2018, 30, 1706807. [Google Scholar] [CrossRef]
  26. Dohrn, R.; Fonseca, J.M.; Albers, R.; Kušan-Bindels, J.; Marrucho, I.M. Thermal Conductivity of Polyurethane Foam Cell Gases: Improved Transient Hot Wire Cell—Data of Isopentane + n-Pentane Mixtures—Extended Wassiljewa-Model. Fluid Phase Equilibria 2007, 261, 41–49. [Google Scholar] [CrossRef]
  27. Kulkarni, B.D.; Doraiswamy, L.K. Estimation of Effective Transport Properties in Packed Bed Reactors. Catal. Rev. 1980, 22, 431–483. [Google Scholar] [CrossRef]
  28. Zhai, S.; Zhang, P.; Xian, Y.; Zeng, J.; Shi, B. Effective Thermal Conductivity of Polymer Composites: Theoretical Models and Simulation Models. Int. J. Heat Mass Transf. 2018, 117, 358–374. [Google Scholar] [CrossRef]
  29. Han, Z.; Fina, A. Thermal Conductivity of Carbon Nanotubes and Their Polymer Nanocomposites: A Review. Prog. Polym. Sci. 2011, 36, 914–944. [Google Scholar] [CrossRef]
  30. Cao, J.-P.; Zhao, X.; Zhao, J.; Zha, J.-W.; Hu, G.-H.; Dang, Z.-M. Improved Thermal Conductivity and Flame Retardancy in Polystyrene/Poly(Vinylidene Fluoride) Blends by Controlling Selective Localization and Surface Modification of SiC Nanoparticles. ACS Appl. Mater. Interfaces 2013, 5, 6915–6924. [Google Scholar] [CrossRef]
  31. Sadati, S.E.; Rahbar, N.; Kargarsharifabad, H.; KhalesiDoost, A. Low Thermal Conductivity Measurement Using Thermoelectric Technology—Mathematical Modeling and Experimental Analysis. Int. Commun. Heat Mass Transf. 2021, 127, 105534. [Google Scholar] [CrossRef]
  32. Islam, S.; Alam, S.M.M.; Akter, S. Mathematical Investigation of the Thermal Conductivity of Fabrics Using Thermal Equation. Mater. Today Proc. 2021, 46, 413–424. [Google Scholar] [CrossRef]
  33. Jacquot, A.; Lenoir, B.; Dauscher, A.; Stölzer, M.; Meusel, J. Numerical Simulation of the 3ω Method for Measuring the Thermal Conductivity. J. Appl. Phys. 2002, 91, 4733–4738. [Google Scholar] [CrossRef]
  34. Zhang, Y.-F.; Zhao, Y.-H.; Bai, S.-L.; Yuan, X. Numerical Simulation of Thermal Conductivity of Graphene Filled Polymer Composites. Compos. Part B Eng. 2016, 106, 324–331. [Google Scholar] [CrossRef]
  35. Chung, K.M.; Zeng, J.; Adapa, S.R.; Feng, T.; Bagepalli, M.V.; Loutzenhiser, P.G.; Albrecht, K.J.; Ho, C.K.; Chen, R. Measurement and Analysis of Thermal Conductivity of Ceramic Particle Beds for Solar Thermal Energy Storage. Sol. Energy Mater. Sol. Cells 2021, 230, 111271. [Google Scholar] [CrossRef]
  36. Liang, J.-Z. Estimation of Thermal Conductivity for Polypropylene/Hollow Glass Bead Composites. Compos. Part B Eng. 2014, 56, 431–434. [Google Scholar] [CrossRef]
  37. Yüksel, N. The Review of Some Commonly Used Methods and Techniques to Measure the Thermal Conductivity of Insulation Materials; IntechOpen: London, UK, 2016. [Google Scholar] [CrossRef]
  38. ASTM C1046; Standard Practice for In-Situ Measurement of Heat Flux and Temperature on Building Envelope Components. ASTM International: West Conshohocken, PA, USA, 2021. Available online: https://www.astm.org/c1046-95r21.html (accessed on 26 August 2022).
  39. Lu, X.; Memari, A.M. Comparison of the Experimental Measurement Methods for Building Envelope Thermal Transmittance. Buildings 2022, 12, 282. [Google Scholar] [CrossRef]
  40. Wang, X.; Feng, J.; Bai, Y.; Zhang, Q.; Yin, Y. Synthesis, Properties, and Applications of Hollow Micro-/Nanostructures. Chem. Rev. 2016, 116, 10983–11060. [Google Scholar] [CrossRef]
  41. Hu, F.; Wu, S.; Sun, Y. Hollow-Structured Materials for Thermal Insulation. Adv. Mater. 2019, 31, 1801001. [Google Scholar] [CrossRef]
  42. Loeb, A.L. Thermal Conductivity: VIII, A Theory of Thermal Conductivity of Porous Materials. J. Am. Ceram. Soc. 1954, 37, 96–99. [Google Scholar] [CrossRef]
  43. Mishra, D.; Satapathy, A. Development of Theoretical Model for Effective Thermal Conductivity of Glass Microsphere Filled Polymer Composites. Plast. Polym. Technol. 2013, 2, 9. [Google Scholar]
  44. Agari, Y.; Tanaka, M.; Nagai, S.; Uno, T. Thermal Conductivity of a Polymer Composite Filled with Mixtures of Particles. J. Appl. Polym. Sci. 1987, 34, 1429–1437. [Google Scholar] [CrossRef]
  45. Suvorov, S.A.; Fishchev, V.N.; Kapustina, S.N.; Kopylova, S.V.; Nekhlopochina, M.G. Thermal Conductivity of Composites with a Filler of Hollow Spherical Corundum Granules. Refractories 1988, 29, 689–693. Available online: https://ur.booksc.eu/book/6507799/7461db (accessed on 9 December 2021). [CrossRef]
  46. Cheng, S.C.; Vachon, R. A Technique for Predicting the Thermal Conductivity of Suspensions, Emulsions and Porous Materials. Int. J. Heat Mass Transf. 1970, 13, 537–546. [Google Scholar] [CrossRef]
  47. Wang, M.; Pan, N. Predictions of Effective Physical Properties of Complex Multiphase Materials. Mater. Sci. Eng. R Rep. 2008, 63, 1–30. [Google Scholar] [CrossRef]
  48. Nait-Ali, B.; Haberko, K.; Vesteghem, H.; Absi, J.; Smith, D.S. Thermal Conductivity of Highly Porous Zirconia. J. Eur. Ceram. Soc. 2006, 26, 3567–3574. [Google Scholar] [CrossRef]
  49. Schumacher, C.; Straube, J.; Ober, D.; Grin, A. Development of a New Hot Box Apparatus to Measure Building Enclosure Thermal Performance; RDH Building Science: Burnaby, BC, Canada, 2013; Available online: https://www.rdh.com/resource/development-of-a-new-hot-box-apparatus-to-measure-building-enclosure-thermal-performance/ (accessed on 28 August 2022).
  50. Seitz, S.; Macdougall, C. Design of an Affordable Hot Box Testing Apparatus. In Proceedings of the 16th NOCMAT Non-Conventional Materials and Technologies, Winnipeg, MB, Canada, 10–13 August 2015. [Google Scholar]
  51. Cai, S.; Zhu, W.; Cremaschi, L. Pipe Insulation Thermal Conductivity under Dry and Wet Condensing Conditions with Moisture Ingress A Critical Review. HVAC&R Res. 2014, 20, 458–479. [Google Scholar]
  52. Connor, B. Comparison of Polystyrene Expanded and Extruded Foam Insulation in Roadway and Airport Embankments; University of Alaska: Fairbanks, Alaska, 2019; p. 37. [Google Scholar]
  53. Bhattacharya, A.; Calmidi, V.V.; Mahajan, R.L. Thermophysical Properties of High Porosity Metal Foams. Int. J. Heat Mass Transf. 2002, 45, 1017–1031. [Google Scholar] [CrossRef]
  54. Paek, J.; Kang, B.; Kim, S.; Hyun, J. Effective Thermal Conductivity and Permeability of Aluminum Foam Materials 1. Int. J. Thermophys. 2000, 21, 453–464. [Google Scholar] [CrossRef]
  55. Calmidi, V.V.; Mahajan, R.L. The Effective Thermal Conductivity of High Porosity Fibrous Metal Foams. J. Heat Transf. 1999, 121, 466–471. [Google Scholar] [CrossRef]
  56. Dyga, R.; Witczak, S. Investigation of Effective Thermal Conductivity Aluminum Foams. Procedia Eng. 2012, 42, 1088–1099. [Google Scholar] [CrossRef]
  57. Boomsma, K.; Poulikakos, D. On the Effective Thermal Conductivity of a Three-Dimensionally Structured Fluid-Saturated Metal Foam. Int. J. Heat Mass Transf. 2001, 44, 827–836. [Google Scholar] [CrossRef]
  58. Jeffreys, H. LXXVI. The Stability of a Layer of Fluid Heated Below. Lond. Edinb. Dublin Philos. Mag. J. Sci. 1926, 2, 833–844. [Google Scholar] [CrossRef]
  59. Nygård, H.S.; Tyvand, P.A. Onset of Thermal Convection in a Vertical Porous Cylinder with a Partly Conducting and Partly Penetrative Cylinder Wall. Transp. Porous Media 2011, 86, 229–241. [Google Scholar] [CrossRef]
  60. Nygård, H.S.; Tyvand, P.A. Onset of Convection in a Porous Rectangle with Buoyancy Along an Open Sidewall. Transp. Porous Media 2011, 90, 403–420. [Google Scholar] [CrossRef]
  61. Nygård, H.; Tyvand, P. Onset of Convection in a Porous Box with Partly Conducting and Partly Penetrative Sidewalls. Transp. Porous Media 2010, 84, 55–73. [Google Scholar] [CrossRef]
  62. Adrian, B. Convection Heat Transfer, 4th ed.; Wiley: Hoboken, NJ, USA, 2013; Available online: https://www.wiley.com/en-us/Convection+Heat+Transfer%2C+4th+Edition-p-9780470900376 (accessed on 21 January 2022).
  63. Nield, D.A.; Bejan, A. Heat Transfer Through a Porous Medium. In Convection in Porous Media; Nield, D.A., Bejan, A., Eds.; Springer: New York, NY, USA, 2013; pp. 31–46. [Google Scholar] [CrossRef]
  64. Thermal Conductivity Coefficient of Rock Wool; Vila Rock Wool Company: Vilashahr, Najafabad, 1399; Available online: https://www.vilainsulgroup.com/thermal-conductivity-coefficient-of-rock-wool/?lang=en (accessed on 3 February 2023).
  65. Jeon, C.-K.; Lee, J.-S.; Chung, H.; Kim, J.-H.; Park, J.-P. A Study on Insulation Characteristics of Glass Wool and Mineral Wool Coated with a Polysiloxane Agent. Adv. Mater. Sci. Eng. 2017, 2017, e3938965. [Google Scholar] [CrossRef]
  66. Bianco, L.; Pollo, R.; Serra, V. Wood Fiber vs Synthetic Thermal Insulation for Roofs Energy Retrofit: A Case Study in Turin, Italy. Energy Procedia 2017, 111, 347–356. [Google Scholar] [CrossRef]
  67. XPS and EPS Comparison. Available online: https://www.karplus.com.tr/EN/xpsandepscomparison.html (accessed on 3 February 2023).
  68. Gravit, M.; Kuleshin, A.; Khametgalieva, E.; Karakozova, I. Technical Characteristics of Rigid Sprayed PUR and PIR Foams Used in Construction Industry. IOP Conf. Ser. Earth Environ. Sci. 2017, 90, 012187. [Google Scholar] [CrossRef]
  69. 5—Performance of the Bio-Based Materials. In Performance of Bio-Based Building Materials; Jones, D.; Brischke, C. (Eds.) Woodhead Publishing: Sawston, UK, 2017; pp. 249–333. [Google Scholar] [CrossRef]
  70. Simmler, H.; Brunner, S.; Heinemann, U.; Schwab, H.; Kumaran, K.; Mukhopadhyaya, P.; Quenard, D.; Sallee, H.; Noller, K.; Kuecuekpinar-Niarchos, E.; et al. Vacuum Insulation Panels—Study on VIP-Components and Panels for Service Life Prediction of VIP in Building Applications (Subtask A); U.S. Department of Energy, Office of Scientific and Technical Information: Oak Ridge, TN, USA, 2005.
  71. Yang, W.; Liu, J.; Wang, Y.; Gao, S. Experimental Study on the Thermal Conductivity of Aerogel-Enhanced Insulating Materials under Various Hygrothermal Environments. Energy Build. 2020, 206, 109583. [Google Scholar] [CrossRef]
Figure 1. Morphology of a typical hollow-structured material.
Figure 1. Morphology of a typical hollow-structured material.
Buildings 13 03094 g001
Figure 2. The relationship between effective thermal conductivity and the volume fraction.
Figure 2. The relationship between effective thermal conductivity and the volume fraction.
Buildings 13 03094 g002
Figure 3. A porous PP sample (4 mm hole).
Figure 3. A porous PP sample (4 mm hole).
Buildings 13 03094 g003
Figure 4. A multi-hollow 3D polypropylene block.
Figure 4. A multi-hollow 3D polypropylene block.
Buildings 13 03094 g004
Figure 5. Hot box (outside and inside).
Figure 5. Hot box (outside and inside).
Buildings 13 03094 g005
Figure 6. Schematic of the experimental hot box test.
Figure 6. Schematic of the experimental hot box test.
Buildings 13 03094 g006
Figure 7. The Campbell Scientific datalogger system.
Figure 7. The Campbell Scientific datalogger system.
Buildings 13 03094 g007
Figure 8. A graph of effective thermal conductivity versus hole size of porous PP.
Figure 8. A graph of effective thermal conductivity versus hole size of porous PP.
Buildings 13 03094 g008
Figure 9. Snapshot of the sensors’ measurements on the PC400 software.
Figure 9. Snapshot of the sensors’ measurements on the PC400 software.
Buildings 13 03094 g009
Figure 10. Measuring heat flux through thermal insulation using a heat flux sensor.
Figure 10. Measuring heat flux through thermal insulation using a heat flux sensor.
Buildings 13 03094 g010
Figure 11. A graph of pore size of PP versus Rayleigh number.
Figure 11. A graph of pore size of PP versus Rayleigh number.
Buildings 13 03094 g011
Figure 12. Thermal conductivity measurement by test type.
Figure 12. Thermal conductivity measurement by test type.
Buildings 13 03094 g012
Figure 13. A bar chart comparing building insulators’ thermal conductivity.
Figure 13. A bar chart comparing building insulators’ thermal conductivity.
Buildings 13 03094 g013
Table 1. Measurement values from DTC-25.
Table 1. Measurement values from DTC-25.
PP hole size (mm)1.002.504.00
Upper temperature (°C)41.1041.4441.10
Lower temperature (°C)5.945.625.94
Heat sink temperature (°C)3.683.573.68
Mean sample temperature (°C)23.5223.5323.52
Thermal conductivity (W/mK)0.280.250.21
Table 2. Thermal conductivity values of building insulators.
Table 2. Thermal conductivity values of building insulators.
Building InsulatorThermal Conductivity (W/mK)
Rock wool0.04 [64]
Glass wool0.034 [65]
Wood wool0.04 [66]
EPS0.045 [67]
XPS0.028 [67]
PUR0.024 [68]
Cellulose0.040 [69]
VIPs0.004 [70]
Aerogel0.012 [71]
Porous PP0.0033 (current study)
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

Osaze, O.; Khanna, S. Experimental Thermal Conductivity Measurement of Hollow-Structured Polypropylene Material by DTC-25 and Hot Box Test. Buildings 2023, 13, 3094. https://doi.org/10.3390/buildings13123094

AMA Style

Osaze O, Khanna S. Experimental Thermal Conductivity Measurement of Hollow-Structured Polypropylene Material by DTC-25 and Hot Box Test. Buildings. 2023; 13(12):3094. https://doi.org/10.3390/buildings13123094

Chicago/Turabian Style

Osaze, Osasu, and Sanjeev Khanna. 2023. "Experimental Thermal Conductivity Measurement of Hollow-Structured Polypropylene Material by DTC-25 and Hot Box Test" Buildings 13, no. 12: 3094. https://doi.org/10.3390/buildings13123094

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