Enhanced mechanical and thermal properties of hollow wood composites filled with phase-change material

Enhanced mechanical and thermal properties of hollow wood composites filled with phase-change material

Journal Pre-proof Enhanced mechanical and thermal properties of hollow wood composites filled with phase-change material Chusheng Qi, Feng Zhang, Jun ...

4MB Sizes 0 Downloads 54 Views

Journal Pre-proof Enhanced mechanical and thermal properties of hollow wood composites filled with phase-change material Chusheng Qi, Feng Zhang, Jun Mu, Yang Zhang, Zhiming Yu PII:

S0959-6526(20)30420-0

DOI:

https://doi.org/10.1016/j.jclepro.2020.120373

Reference:

JCLP 120373

To appear in:

Journal of Cleaner Production

Received Date: 18 November 2019 Revised Date:

20 January 2020

Accepted Date: 31 January 2020

Please cite this article as: Qi C, Zhang F, Mu J, Zhang Y, Yu Z, Enhanced mechanical and thermal properties of hollow wood composites filled with phase-change material, Journal of Cleaner Production (2020), doi: https://doi.org/10.1016/j.jclepro.2020.120373. This is a PDF file of an article that has undergone enhancements after acceptance, such as the addition of a cover page and metadata, and formatting for readability, but it is not yet the definitive version of record. This version will undergo additional copyediting, typesetting and review before it is published in its final form, but we are providing this version to give early visibility of the article. Please note that, during the production process, errors may be discovered which could affect the content, and all legal disclaimers that apply to the journal pertain. © 2020 Published by Elsevier Ltd.

Credit Author Statement Chusheng Qi: Conceptualization, methodology, investigation, writing-original draft, writingreview & Editing Feng Zhang: Formal analysis, investigation, writing-original draft Jun Mu: Funding acquisition, methodology, supervision Yang Zhang: Resources, validation Zhiming Yu: Funding acquisition, resources

Enhanced Mechanical and Thermal Properties of Hollow Wood Composites Filled with Phase-Change Material

Chusheng Qi 1,2, Feng Zhang 1,2, Jun Mu* 1,2, Yang Zhang 1,2, Zhiming Yu 1,2

1. Key Laboratory of Wood Material Science and Utilization of Ministry of Education, Beijing Forestry University, Beijing, 100083, P. R. China. 2. Beijing City Key Laboratory of Wood Science and Engineering, Beijing Forestry University, Beijing, 100083, P. R. China.

Corresponding author: Dr. Jun Mu Professor Key Laboratory of Wood Material Science and Utilization of Ministry of Education, Beijing Forestry University, Beijing, 100083, P. R. China. Email: [email protected]

1

Graphica abstract

1

Enhanced Mechanical and Thermal Properties of Hollow Wood Composites Filled

2

with Phase-Change Material

3 4

Wordcount: 5153 words with 28 pages

5

Abstract

6

To obtain lightweight wood building materials with good thermal insulation,

7

energy-saving properties, and satisfying mechanical properties, low-density fiberboard

8

and hollow wood composites (HWC) embedded polyvinyl chloride tubes were

9

fabricated by hot-pressing. Polyethylene glycol was used as the phase-change material

10

to fill polyvinyl chloride tubes and obtain phase-change hollow wood composite

11

(PHWC). The physical and mechanical properties of HWC and PHWC were tested, and

12

their thermal properties were analyzed and simulated. The results showed that the

13

thermal conductivities of low-density fiberboard, HWC, and PHWC ranged from

14

0.06-0.07 W/(m·K), indicating they had sufficient physical and mechanical properties to

15

be used as thermal insulation building materials. The combination of series and parallel

16

models accurately predicted the thermal conductivity of HWC and PHWC, whose

17

structures were similar to a series structure. The addition of polyethylene glycol into

18

HWC allowed the PHWC to store latent heat and reduce indoor temperature fluctuations.

19

Heat transfer simulations showed that when used as a non-structural building wall

20

material, the PHWC wall had a better energy efficiency compared with a concrete wall.

21

Thus, PHWC has potential applications as thermal insulation and phase-change building

22

material.

23

Keywords: Hollow wood composites; Lightweight wood composites; Phase-change

24

material; Mechanical properties; Thermal conductivity 2

25 26

1. Introduction The energy consumption of buildings accounts for more than 30% of total global

27

energy usage (Berardi, 2017), and thermal insulation of building envelopes is one of the

28

main techniques to reduce energy consumed by air conditioning. Thermal insulation

29

products such as fiberglass, mineral wool, and polyurethane foams have low thermal

30

conductivities in the range 0.02-0.05 W/(m·K), but they pose environmental and health

31

hazards (Corporation, 2004). Additionally, polyurethane foams are highly flammable.

32

Building materials should meet relevant requirements for structural safety, quality of life,

33

energy efficiency, cost, fire resistance, and durability (Matalkah et al., 2017). Wood is a

34

natural and renewable material that has be used in many applications, such as bioenergy

35

(Mardoyan and Braun, 2015; Pradhan et al., 2018), biorefining (Akim, 2016; MarouÅ

36

and Žák, 2015) and biochar (Agegnehu et al., 2017; Maroušek et al., 2019).

37

Wood-based composites with simple processing technologies and low energy and

38

financial requirements are one of the most important uses of wood. Lightweight wood

39

composites (LWC) with thermal and sound insulation properties have great application

40

prospects for use in building envelopes. Xie et al. (2011) found that ultra-low density

41

fiberboard had a very low thermal conductivity and a high sound reduction coefficient.

42

Many other studies have shown that thermal conductivity decreases with the density of

43

natural fiber composites because gases have a lower thermal conductivity and far slower

44

heat transfer than solids (Binici et al., 2012). Additionally, the shortage of forestry

45

resources (Li et al., 2017), forest destruction due to wildfires and the prohibition of the

46

logging of natural forests in some countries have made it necessary to develop LWC

47

which consuming fewer wood resources.

48

Much research has been devoted to developing many types of LWC. For example, 3

49

Hussain et al. (2019) fabricated wood-based sandwich panels with wood-based cores

50

and face sheets composed of glass-fiber-reinforced polymer with a density of 0.21

51

g/cm3. Monteiro et al. (2019) produced particleboards with a density range of 0.32-0.54

52

g/cm3 using sour cassava starch as the adhesive and foam. The formation of hollow

53

wood composites can also reduce the density of a product while retaining its satisfying

54

performance. For example, Voth et al. (2015) fabricated wood-stand sandwich panels

55

with hollow-core interiors and a density of around 0.3 g/cm3. The interior cavity of

56

hollow wood composites can be used to house thermal and sound insulation materials,

57

but they must be further developed for filling with phase-change materials (PCM) due

58

to PCM leakage after melting.

59

Energy demand today is rapidly increasing around the world (Dag et al., 2019), and

60

global buildings were responsible for about 32% of energy consumption and 19% of

61

energy-related greenhouse gases emission in 2010 (Abanda and Byers, 2016). This has

62

caused the various governments to focus on the research and development of

63

energy-efficient buildings. Phase-change materials are used to reduce energy

64

consumption by decreasing indoor temperatures, decreasing indoor temperature

65

fluctuations, and shifting loads away from peak usage times. The incorporation of PCM

66

in building components may reduce indoor temperature fluctuations because PCM can

67

store large amounts of latent heat within a small temperature range associated with a

68

phase change. The application of PCM in buildings has been conducted to reduce

69

energy usage and increase energy efficiency (Ascione et al., 2014; Li et al., 2015;

70

Shafie-khah et al., 2016). The high economic value of PCM for reducing energy

71

consumption in a typical multistory office building was demonstrated by Mi et al.

72

(2016). Potential leakage is the primary problem with the use of solid-to-liquid PCM 4

73

because they exist as liquids above their melting temperature. An inexpensive and

74

effective solution to this problem is to contain PCM in hollow plastic tubes.

75

The objective of this research is to develop lightweight wood composites with

76

hollow thermoplastic tubes, which were then filled with PCM to obtain final composites

77

with thermal insulation and thermal absorption properties. The physical and mechanical

78

properties and the thermal performance of the obtained composites were analyzed and

79

simulated.

80

2 Materials and methods

81

2.1 Materials

82

Poplar wood fibers were obtained from a local market in China and then ground to

83

80 mesh with a moisture content of 6.0 %. Polyvinyl chloride (PVC) tubes with an outer

84

diameter of 7.0 mm, a wall thickness of 0.5 mm, and a solid density of 1.2 g/cm3 were

85

purchased from Dongguan Taolue Electronic Products Co., Ltd. Liquid isocyanate

86

adhesive (MDI, PM200) with an NCO content of 30.5 % and a viscosity of 250 MPa·s

87

(25oC) was obtained from Wanhua Chemicals. Acetone purchased from Beijing

88

Chemicals was used as a diluent for MDI. Polyethylene glycol (PEG) from Jiangsu

89

Haian Chemical Plant was used as the PCM. PEG-800 with a molecular weight of

90

760-840, a density of 1.27 g/cm3, and a viscosity of 2.3 MPa·s at 25 oC was used in this

91

study.

92

2.2 Fabrication of hollow wood composites

93

Hollow wood composites (HWC) with dimensions of 350 × 350 × 20 mm3 were

94

fabricated with a target density of 0.3 g/cm3. Four percent (based on the oven-dried

95

weight of wood fibers) of MDI was used as the adhesive, and it was diluted with 5

96

acetone in a ratio of 1:1 to reduce its viscosity before spraying onto wood fiber in a

97

roller. The wood fiber containing MDI was divided into three portions in a mass ratio of

98

3:4:3. The first 30% of the wood fiber was evenly placed on a caul in a frame with a

99

size of 350× 350 mm2. Parallel hollow PVC tubes settled by a self-made clamp (Fig.

100

1(a)) were placed on the first layer of wood fiber. Then, 40% of wood fiber was added,

101

and another layer of parallel hollow PVC tubes was placed on the second layer of the

102

wood fiber. Finally, the last 30% of the wood fiber was evenly placed on top. The

103

formed mat was placed between two aluminum cauls, and silicon paper was placed

104

between the caul and mat to prevent adhesion between the caul and the final HWC. The

105

mat was pressed for 8 min at 180 oC under a pressure of 1 MPa in a hot press machine,

106

and the thickness of final products was controlled by a thickness gauge. Figure 1(b)

107

shows the final HWC. The low-density fiberboard (LDF) without hollow PVC tubes

108

was also fabricated using the same hot press schedule as the reference material. All

109

samples were equilibrated at 25 oC and a humidity of 65% for at least one week before

110

performance testing, and six replicates of each test were performed.

111

The air volume fraction of HWC was adjusted by changing the number of PVC

112

tubes at a constant HWC bulk density of 0.3 g/cm3. The relationship between the

113

number of tubes and the volume ratios are listed in Eq. 1 - Eq. 4: λ +λ +λ +λ =1 λ = λ = λ =

V × 100% = V

V /

(1)

× 100%

(2)

× 100%

(3)

× 100%

(4)

114 115

where λ , λ , λ , and λ are the volume ratios of the PVC tube cavity, solid PVC, 6

116

wood fiber cell wall, and air inside the wood fiber, respectively; V,

117

the volumes (cm3) of HWC, the PVC tube cavity, solid PVC, and wood fiber cell wall,

118

respectively; n is the number of PVC tubes,

119

fiber (g); and

120

λ and λ is the air volume ratio of HWC (λ ). Different λ and λ values were

121

obtained by changing the number of PVC tubes, and Table 1 shows the experimental

122

design and corresponding parameters.

123

Table 1 Experimental design and parameters of hollow wood composites

,

, and V

are

represents the dry weight of wood

is the density of the wood fiber cell wall (1.5 g/cm3). The sum of

Sample

Number of PVC tubes

λ

λ

λ

(%)

(%)

HWC-8

19

7.67

HWC-9

22

HWC-10

25

Design density (g/cm3)

Average HWC density (g/cm3)

λ

λ

(%)

(%)

(%)

2.77

17.78

71.78

79.45

0.289

8.88

3.21

17.43

70.48

79.36

0.274

10.09

3.65

17.08

69.18

79.27

0.308 0.3

HWC-11

28

11.30

4.08

16.73

67.89

79.19

0.292

HWC-12

31

12.52

4.52

16.38

66.58

79.10

0.289

HWC-13

34

13.73

4.96

16.03

65.28

79.01

0.280

124 125 126

2.3 Preparation of phase-change hollow wood composites Polyethylene glycol was first melted at 50oC in a water bath, and then 7.7 g of PEG

127

was injected via syringe into each hollow PVC tube inside the HWC. Then, the two

128

sides of the hollow PVC tube were sealed using a thermosetting resin to obtain

129

phase-change hollow wood composite (PHWC) (Fig. 1(c)). The melt PEG was cooled to

130

20oC before testing, and six replicates were performed for each test.

131

2.4 Physical and mechanical property evaluation 7

132

The modulus of rupture (MOR), modulus of elasticity (MOE), internal bond

133

strength (IB), thickness swelling (TS), and water absorption (WA) of the samples were

134

tested based on the Chinese Standard GB/T17657-2013. Nine specimens were tested for

135

each measurement.

136

The thermal conductivity and thermal resistance of HWC and PHWC were

137

measured using a thermal conductivity analyzer (DRH-300) according to standard GB/T

138

10294-2008. The cold plate and hot plate temperatures were set to 23oC and 43oC,

139

respectively, with an ambient temperature of 21oC.

140

A self-made temperature detector (Fig. 2) was developed and used to test the heat

141

transfer characteristics of LDF, HWC, and PHWC. Samples were placed on a hot plate

142

set to 50oC and surrounded and covered by a 5 mm thick polystyrene foam board to

143

prevent heat exchange with the environment. Nine thermocouples were placed on the

144

top surface to monitor the temperature changes over time. The samples were first heated

145

at 50oC for 1 h and then cooled naturally for another 1 h.

146

2.5 Vertical density profile analysis

147

The vertical density profile (VDP) of lightweight wood composites, HWC, and

148

PHWC was measured by an X-ray profile density analyzer (GreCon DAX 6000) with a

149

scan speed of 0.5 mm/s and a specimen size of 50 × 50 × 20 mm3.

150

2.6 DSC analysis

151

Differential scanning calorimetry (DSC) was used to evaluate the thermal

152

properties and energy storage of PEG from -20 oC - 100 oC at a heating rate of 5 oC /min

153

under a nitrogen (N2) atmosphere. The temperature and enthalpy accuracies of the DSC

154

were 0.1oC and 1%, respectively. The measured heat enthalpy and specific heat of PEG 8

155

were used as the material parameters for heat transfer simulations.

156

2.7 Heat transfer simulations

157

The heat transfer of HWC and PHWC was simulated using ANSYS Workbench

158

software with a simplified model size of 50 × 50 × 20 mm3. The heat transfer of a

159

building wall made of PHWC with a thickness of 240 mm was also simulated and

160

compared with a building wall made of concrete. An outdoor temperature of 40 oC, a

161

convective heat transfer coefficient of 20 W/(m2·K) for the inner surface, and an initial

162

environmental temperature of 20 oC were used to simulate a typical summer

163

environment in southern China. The material properties used in the simulation were

164

experimentally determined and are listed in Table 2.

165

Table 2 Material properties for heat transfer simulation Properties Density (g/cm3) Specific heat (J/(kg·K)) Thermal conductivity(W/ (m·K))

Wood fiber

PVC plastic

PEG

Concrete (Zhao et al., 2011)

0.3

1.2

1.27

2.4

1173 (Guo et al., 2013)

895.6

3393

960

0.061

0.14 (Zhang et al., 2015)

0.24 (Wu et al., 2019)

1.28

166 167

3. Results and discussion

168

3.1 Physical and mechanical properties of HWC

169 170

Figure 3 (a) shows the MOR and MOE values of LDF and HWC. The low-density fiberboard without added PVC hollow tubes had a MOR of 1.8 MPa and a MOE of 9

171

196.1 MPa. These values were much lower than medium-density fiberboard (MDF),

172

which has a MOR of around 28.0 MPa and MOE of 1.4 GPa because they were much

173

less dense than MDF (around 0.65 g/cm3) (Hussain et al., 2019), but these mechanical

174

properties can support the transportation, hoisting, and installation of LDF and HWC.

175

Panyakaew and Fotios (2011) reported that the MOR and MOE of binderless bagasse

176

boards with a density of 0.25 g/cm3 were 0.43 MPa and 102 MPa, respectively, and

177

these values increased with the density. The MOR and MOE values obtained in this

178

study were higher than those previously reported, which was attributed to the addition

179

of the MDI resin. The isocyanate group of MDI reacted with the hydroxyl group of the

180

wood fiber to form a strong covalent bond that contributed to the good mechanical

181

properties of LDF and HWC. There was an obvious increase in the MOR and MOE

182

when hollow PVC tubes were added into LDF (Fig. 3 (a)), which was attributed to the

183

toughness and stiffness of PVC tubes. As λ increased, the MOR and MOE of HWC

184

slightly fluctuated. A higher λ indicates hollower PVC tubes and less wood fiber in

185

HWC. The PVC tubes had positive effects on MOR and MOE, but decreasing the

186

wood fiber content had an inverse effect.

187

The internal bond strengths of LDF and HWC are given in Fig. 3 (b). LDF had an

188

IB of 0.06 MPa, which is within the range of conventional thermal insulation materials,

189

which typically have IB values of 0.03 - 0.08 MPa (Pfundstein et al., 2012). Kawasaki

190

et al. (1998) reported that low-density fiberboard with a density of 0.3 g/cm3 had an IB

191

of around 0.1 MPa. These results demonstrate that low-density fiberboard has a low IB

192

because the fibers are not close enough to be bonded. However, the IB of LDF was

193

much higher than that of binderless bagasse board (0.01 MPa) with a density of 0.35

194

g/cm3 (Panyakaew and Fotios, 2011), showing that the MDI resin improves the 10

195

mechanical properties. The addition of hollow PVC tubes did not improve the IB,

196

possibly because no interfacial bond was formed between the hydrophobic PVC and

197

hydrophilic wood fiber. This should be improved in future work.

198

The thickness swelling and water absorption of LDF and HWC are shown in Fig. 3

199

(c) and Fig. 3 (d). The thickness swelling of LDF after 2 h and 24 h were 4.4% and

200

9.3%, respectively. For comparison, a 24 h thickness swelling of 13.2% was observed in

201

particleboard with a density of 0.4 g/cm3 (Monteiro et al., 2019), and LDF in this study

202

had better water resistance when considering its low density. Thickness swelling slightly

203

increased when hollow PVC tubes were added due to increased water passage and

204

contact surface since the water can fill the gaps between the hollow PVC tubes and

205

fibers. The water absorption rates of LDF after 2 h and 24 h were 39.2% and 104.5%,

206

respectively, and the 24 h water absorption of HWC had a trend of decrease as λ

207

increased. The water penetrated the voids between the wood fibers and lumen, and the

208

chemical components of the wood cell walls absorbed water and swelled. Therefore,

209

when more voids and lumen are present inside the HWC, and the thickness swelling and

210

water absorption rate are both higher. λ , which indicates the void fraction of lumen

211

inside HWC, decreased as λ increased (Table 1), causing the 24 h thickness swelling

212

and water absorption to decrease with λ .

213

3.2 Vertical density profile analysis

214

Figure 4 (a) presents the typical vertical density profiles of LDF, HWC, and

215

PHWC in the thickness direction, and the results show that the VDP for LWC was

216

nearly constant with an average of 0.3 g/cm3. However, there were large fluctuations for

217

HWC when hollow PVC tubes were present, and a U-type shape was observed. This

218

mainly occurred because the hollow PVC tube had a U-type shape density profile (Fig. 11

219

4 (b)). The density of PVC is 1.2 g/cm3, which is much higher than LWC, and the

220

vertical density of hollow PVC tube ranged from 0.2 - 1.2 g/cm3 (Fig. 4 (b)), resulting

221

in a lower density (around 0.2 g/cm3) of wood fiber in HWC when the overall density

222

was the same. This low density resulted in a low internal bond strength, as shown in Fig.

223

3 (b). When the hollow PVC tubes inside the HWC were filled with PEG (ρ = 1.27

224

g/cm3), the corresponding VDP of PVC tubes increased, as shown in Fig. 4 (a).

225

3.3 Thermal conductivity and resistance

226

The thermal conductivity (k) is inversely proportional to the thermal resistance (R),

227

and their values of HWC and PHWC are shown in Fig. 5(a). The k and R of LDF were

228

0.06 W/(m·K) and 0.33 m2·K/W at 33 oC, respectively. The thermal conductivity of

229

HWC and PHWC ranged from 0.06-0.07 W/(m·K), indicating that the lightweight wood

230

composites with a density of 0.3 g/cm3 could be used as a thermally insulating building

231

material since its thermal conductivity was below 0.12 W /(m·K). The thermal

232

conductivities of binderless coconut husk and bagasse insulation boards with densities

233

of 250-350 kg/m3 have been reported to rang from 0.046 to 0.068 W/(m·K) (Panyakaew

234

and Fotios, 2011), which are consistent with the results reported here. The addition of

235

hollow PVC tubes into the LWC did not significantly influence their thermal

236

conductivity or resistance since the thermal conductivity of solid PVC (

237

W/(m·K) (Zhang et al., 2015) and higher than that of LWC. This was mainly because

238

both λ

239

had the lowest thermal conductivity of 0.026 W/(m·K) at 33 oC. PHWC had a slightly

240

higher thermal conductivity than HWC since the PEG-800 had a thermal conductivity

241

of 0.24 W/(m·K) (Wu et al., 2019), showing that the addition of PEG-800 increased the

242

thermal conductivity of PHWC. High thermal conductivity helped improve the energy

and λ

) is 0.14

decreased with an increase in λ (Table 1), and air inside the HWC

12

243

change efficiency of phase-change materials.

244

Series model, parallel model, as well as a combined series and parallel model for

245

three phases were employed to predict the thermal conductivity of HWC and PHWC.

246

The HWC was assumed to be composed of three phases: wood fiber (including wood

247

fiber lumen and voids between wood fibers), solid PVC, and cavity of the hollow PVC

248

tubes. For PHWC, the PEG phase was instead of the PVC tube cavity. The series model

249

(Eq. 5), parallel model (Eq. 6), and their combined model (Eq. 7) can be described as

250

follows: 1

= ( = ε

= ( 251

where

+

+ (1 − ε +

+ +

+

+

(5) )

+

(6)

+

(7)



is the thermal conductivity of the final HWC or PHWC;

is the

252

thermal conductivity of the wood fiber phase, and its experimentally determined value

253

was 0.163 W/(m·K);

254

hollow PVC tube for HWC and PEG (0.24 W/(m·K)) in PHWC;

255

and λ in Table 1, and the sum of

256

factor of the series model to the parallel model and is decided by the material structure

257

and can be calculated by linear fitting.

is the thermal conductivity of air (0.026 W/(m·K)) inside the

,

, and

is the sum of λ

is one, and ε signifies the ratio

258

Figure 5 (b) and (c) compare the experimental data and predicted results of HWC

259

and PHWC. The parallel model predicted the maximum thermal conductivity when the

260

series model predicted the minimum level, and the experimental data fall between these

261

two extremes. ε was found to be 0.79 for HWC and 0.94 for PHWC with an R2 of 0.99, 13

262

indicating that the structures of HWC and PHWC were very similar to a serial structure,

263

especially for PHWC with PEG-filled PVC tubes, which can be further shown in Fig. 1

264

(b) and Fig. 1 (c). The combined series and parallel model accurately predicted the

265

thermal conductivity of HWC and PHWC.

266

3.4 The thermal properties of polyethylene glycol

267

The heat flow of PEG during heating and cooling is shown in Fig. 6 (a). The results

268

indicated that the onset melting temperature was 23.4oC, with a melting peak at 29.6oC,

269

which was consistent with previous research (Sánchez et al., 2007) and was in an ideal

270

temperature range for a phase change material for building insulation. The PEG began

271

to release heat when the temperature decreased to 22.0oC with a heat release peak at

272

17.3oC, which will help keep a building warm. The latent heat of PEG was calculated as

273

125.7 J/g, and its heat enthalpy at 0oC, 20oC, 28oC, and 35oC were 3.18× 10! ,

274

1.60× 10" , 2.49× 10" , and 2.76× 10" J/m3, respectively. These values were used for

275

later heat transfer simulations.

276

3.5 Heat transfer characteristics

277

Figure 6 (b) and Fig. 6 (c) present the top surface temperatures of HWC and

278

PHWC when their bottom temperatures were remained at 50oC for 1 hour, followed by

279

natural cooling for another 1 h. The top surface temperatures of HWC with different λ

280

were similar (Fig. 6 (b)), mainly because the HWC density was the same (0.3 g/cm3) as

281

λ and λ changed, and λ

282

the addition of hollow PVC tubes had almost no effect on heat transfer, and their

283

thermal conductivities showed the same trend (Fig. 5(a)). The top surface temperatures

284

of PHWC at different λ were notably different, and a lower top surface temperature

and λ decreased when λ and λ increased. Therefore,

14

285 286

was observed at higher λ since more heat was absorbed by the PEG during melting. Figure 6 (c) compares the top surface temperature of LDF, HWC, and PHWC, and

287

the results show that LDF and HWC had similar temperatures, but PHWC had a much

288

lower temperature than LDF and HWC during the heating stage and a higher

289

temperature during the cooling stage. This indicates that the addition of PEG into the

290

hollow PVC tubes improved the energy efficiency of HWC, which has also been

291

reported by other researchers (Li et al., 2015).

292

Figure 7 (a) compares the simulated temperature distributions at the same heat

293

transfer experimental testing conditions. The simulation results show that PHWC had a

294

notably lower temperature than HWC, and PEG delayed temperature increases, which

295

was consistent with the experimental data.

296

3.6 Thermal transition simulation of a building wall

297

Figure 7(b) and (c) compare the indoor temperatures of a building wall with a

298

thickness of 240 mm made of PHWC and concrete. The results show that the indoor

299

temperature of the room with a PHWC wall was 23.0 oC, which was 3.5 oC lower than

300

the concrete wall after 5 h insulation, indicating that lightweight wood building

301

materials with good thermal insulation, energy-saving properties, and satisfying

302

mechanical properties were successfully developed.

303

The research performed by Yun et al. (2020) showed that the addition of 0.1%-0.9%

304

building volume fraction of PCM with a phase change temperature of 20-28 oC could

305

save up to 3.19 kWh/m2 energy per year. Sharma and Rai (2020) concluded that

306

PCM-enhanced walls could reduce heat gain by 10.4%-26.6%, while also reducing

307

annual electricity consumption and greenhouse gas emissions. The phase-change hollow 15

308

wood composites with a PCM volume fraction of up to 13.7% in this study can

309

potentially be applied as non-structural thermal insulation and phase-change wall

310

materials. Future efforts will attempt to further reduce the density and increase the PCM

311

volume fraction and mechanical properties of PHWC.

312

4 Conclusions

313

Lightweight wood building composites filled with phase-change material were

314

successfully developed and

315

temperature fluctuations. The addition of hollow PVC tubes into LDF increased its

316

MOR and MOE, and the internal bond strength, 24 h thickness swelling, and 24 h water

317

absorption of HWC decreased with an increasing number of hollow PVC tubes. A

318

U-shaped vertical profile in the HWC was observed where the hollow PVC tubes were

319

present, and it fluctuated less in the PEG-filled tube. The thermal conductivities of LDF,

320

HWC, and PHWC ranged from 0.06-0.07 W/(m·K) at 33 oC, and the addition of hollow

321

PVC tubes and PEG filling only slightly affected the thermal conductivity. Combined

322

series and parallel model predicted the thermal conductivity of HWC and PHWC, and

323

their structures were very similar to the series structure. Both experimental data and

324

ANSYS simulation results showed that the addition of PEG decreased the top surface

325

temperature of PHWC when the bottom surface was continuously heated. Compared

326

with concrete, the use of PHWC as a non-structural wall material resulted in a much

327

lower indoor temperature, demonstrating its promising potential for use as a thermal

328

insulation and phase-change building material.

329

Acknowledgment

330

were able to store latent heat and reduce indoor

This work was supported by the Fundamental Research Funds for the Central

16

331

Universities (NO. 2016ZCQ01).

332

References

333

Abanda, F.H., Byers, L., 2016. An investigation of the impact of building orientation on

334

energy consumption in a domestic building using emerging BIM (Building Information

335

Modelling). Energy 97, 517-527.

336

Agegnehu, G., Srivastava, A.K., Bird, M.I., 2017. The role of biochar and

337

biochar-compost in improving soil quality and crop performance: A review. Applied soil

338

ecology 119, 156-170.

339

Akim, E.L., 2016. Biorefining of Wood. Fibre Chemistry 48(3), 181-190.

340

Ascione, F., Bianco, N., De Masi, R.F., de’ Rossi, F., Vanoli, G.P., 2014. Energy

341

refurbishment of existing buildings through the use of phase change materials: Energy

342

savings and indoor comfort in the cooling season. Applied Energy 113, 990-1007.

343

Berardi, U., 2017. A cross-country comparison of the building energy consumptions and

344

their trends. Resources, Conservation and Recycling 123, 230-241.

345

Binici, H., Gemci, R., Kucukonder, A., Solak, H.H., 2012. Investigating sound

346

insulation, thermal conductivity and radioactivity of chipboards produced with cotton

347

waste, fly ash and barite. Construction and Building Materials 30, 826-832.

348

Corporation, S.R., 2004. Toxicological Profile for Synthetic Vitreous Fibers. US

349

Department of Health and Human Services, Public Health Service, Agency for Toxic

350

Substances and Diseases Registry.

351

Dag, M., Akcay, N., Koten, H., Guner, K., 2019. Determination of Photovoltaic 17

352

Properties for Nanostructures. Journal of Electronic Materials 48(11), 6919-6931.

353

Guo, W., Lim, C.J., Bi, X., Sokhansanj, S., Melin, S., 2013. Determination of effective

354

thermal conductivity and specific heat capacity of wood pellets. Fuel 103, 347-355.

355

Hussain, M., Abbas, N., Zahra, N., Sajjad, U., Awan, M.B., 2019. Investigating the

356

performance of GFRP/wood-based honeycomb sandwich panels for sustainable prefab

357

building construction. SN Applied Sciences 1(8), 875.

358

Kawasaki, T., Zhang, M., Kawai, S., 1998. Manufacture and properties of

359

ultra-low-density fiberboard. Journal of Wood Science 44(5), 354-360.

360

Li, L., Hao, T., Chi, T., 2017. Evaluation on China's forestry resources efficiency based

361

on big data. Journal of Cleaner Production 142, 513-523.

362

Li, X., Chen, H., Li, H., Liu, L., Lu, Z., Zhang, T., Duan, W.H., 2015. Integration of

363

form-stable paraffin/nanosilica phase change material composites into vacuum

364

insulation panels for thermal energy storage. Applied Energy 159, 601-609.

365

Mardoyan, A., Braun, P., 2015. Analysis of Czech subsidies for solid biofuels.

366

International jounal of green energy 12(4), 405-408.

367

MarouÅ, J., Žák, J., 2015. Managing environmental innovation: Case study on

368

biorefineryconcept. Universidad del Zulia.

369

Maroušek, J., Strunecký, O., Stehel, V.J.C.T., Policy, E., 2019. Biochar farming:

370

defining economically perspective applications. Clean Technologies and Environmental

371

Policy, 1-7.

372

Matalkah, F., Bharadwaj, H., Soroushian, P., Wu, W., Almalkawi, A., Balachandra, A.M.,

18

373

Peyvandi, A., 2017. Development of sandwich composites for building construction

374

with locally available materials. Construction and Building Materials 147, 380-387.

375

Mi, X., Liu, R., Cui, H., Memon, S.A., Xing, F., Lo, Y., 2016. Energy and economic

376

analysis of building integrated with PCM in different cities of China. Applied Energy

377

175, 324-336.

378

Monteiro, S., Martins, J., Magalhães, F.D., Carvalho, L., 2019. Low Density Wood

379

Particleboards Bonded with Starch Foam—Study of Production Process Conditions.

380

Materials 12(12), 1975.

381

Panyakaew, S., Fotios, S., 2011. New thermal insulation boards made from coconut

382

husk and bagasse. Energy and Buildings 43(7), 1732-1739.

383

Pfundstein, M., Gellert, R., Spitzner, M., Rudolphi, A., 2012. Insulating materials:

384

principles, materials, applications. Walter de Gruyter.

385

Pradhan, P., Mahajani, S.M., Arora, A., 2018. Production and utilization of fuel pellets

386

from biomass: A review. Fuel Processing Technology 181, 215-232.

387

Sánchez, L., Sánchez, P., de Lucas, A., Carmona, M., Rodríguez, J.F.J.C., Science, P.,

388

2007. Microencapsulation of PCMs with a polystyrene shell. Colloid and Polymer

389

Science 285(12), 1377-1385.

390

Shafie-khah, M., Kheradmand, M., Javadi, S., Azenha, M., de Aguiar, J.L.B.,

391

Castro-Gomes, J., Siano, P., Catalão, J.P.S., 2016. Optimal behavior of responsive

392

residential demand considering hybrid phase change materials. Applied Energy 163,

393

81-92.

19

394

Sharma, V., Rai, A.C., 2020. Performance assessment of residential building envelopes

395

enhanced with phase change materials. Energy and Buildings 208, 109664.

396

Voth, C., White, N., Yadama, V., Cofer, W., 2015. Design and Evaluation of

397

Thin-Walled Hollow-Core Wood-Strand Sandwich Panels. Journal of Renewable

398

Materials 3(3), 234-243.

399

Wu, S., Yan, T., Kuai, Z., Pan, W., 2019. Thermal conductivity enhancement on phase

400

change materials for thermal energy storage: A review. Energy Storage Materials,

401

online.

402

Xie, Y., Tong, Q., Chen, Y., Liu, J., Lin, M., 2011. Manufctre and properties of ultra-low

403

density fiberboard from wood fiber. BioResources 6(4), 4055-4066.

404

Yun, B.Y., Park, J.H., Yang, S., Wi, S., Kim, S., 2020. Integrated analysis of the energy

405

and economic efficiency of PCM as an indoor decoration element: Application to an

406

apartment building. Solar Energy 196, 437-447.

407

Zhang, H., Shi, G., Shang, H., 2015. Study on Thermal Conductivity of PVC Plastic

408

Mortar. Materials Review(16), 142-146.

409

Zhao, W., Yang, H., Chen, M., Yang, P., Yang, J., 2011. Measurement of specific

410

thermal/thermal conductivity coefficient of porous expanded pearlite concrete and

411

evaluation of heat preservation property. New Building Material 38(1), 78-80.

20

412

Figure Captions

413 414

Figure 1 (a) Hollow PVC tubes orientation clamp, (b) hollow wood composites, and (c)

415

hollow wood composites filled with polyethylene glycol

416

Figure 2 Schematic diagram of heat transfer testing

417

Figure 3 Physical and mechanical properties of HWC, (a) MOE, and MOR, (b) internal

418

bond strength, (c) thickness swelling, and (d) water absorption

419

Figure 4 (a) Typical vertical density profiles of LDF, HWC, and PHWC, and (b) density

420

distribution of a hollow PVC tube

421

Figure 5 (a) Thermal conductivity and resistance of LDF, HWC, and PHWC, and

422

thermal conductivity comparison of predicted values with experimental data of (b)

423

HWC and (c) PHWC

424

Figure 6 (a) Heat flow of PEG, top surface temperature of (b) HWC, (c) PHWC, and (d)

425

comparison of LDF, HWC, and PHWC with a bottom hot plate temperature of 50 oC.

426

Figure 7 (a) Heat flux and temperature cloud field of HWC and PHWC, (b) indoor

427

surface temperature simulation results, and (c) temperature distribution simulation

428

results of a 240 mm wall made of PHWC and concrete

21

Figure 1 a

c

b

22

Figure 2

23

Figure 3 b

a

d

c

24

Figure 4 a

b

25

Figure 5 a

b

c

26

Figure 6 a

b

c

d

27

Figure 7 b

a

c

28

1

Highlights:

2

Lightweight hollow wood composites with a density of 0.3 g/cm3 were fabricated

3

The thermal conductivity of lightweight hollow wood composites is 0.06~0.07 W/(m·K)

4

Polyethylene glycol as a phase change material was added into hollow wood composites

5

Phase-change lightweight wood composites were better insulators than concrete

Declaration of interests ☒ The authors declare that they have no known competing financial interests or personal relationships that could have appeared to influence the work reported in this paper. ☐The authors declare the following financial interests/personal relationships which may be considered as potential competing interests: