Next Article in Journal
Distribution of Gonipterus Species and Their Egg Parasitoids in Australia: Implications for Biological Control
Next Article in Special Issue
Best Practices for Quasistatic Berkovich Nanoindentation of Wood Cell Walls
Previous Article in Journal
Climate Response in Tree-Rings of Sawara Cypress [Chamaecyparis pisifera (Siebold & Zucc.) Endl.] in Poland
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Temperature-Dependent Creep Behavior and Quasi-Static Mechanical Properties of Heat-Treated Wood

1
College of Materials Science and Art Design, Inner Mongolia Agriculture University, Hohhot 010018, China
2
College of Materials Science and Engineering, Nanjing Forestry University, Nanjing 210037, China
3
Center for Renewable Carbon, University of Tennessee, Knoxville, TN 37996, USA
*
Author to whom correspondence should be addressed.
Submission received: 21 June 2021 / Revised: 19 July 2021 / Accepted: 19 July 2021 / Published: 22 July 2021
(This article belongs to the Special Issue Nanoindentation in Wood)

Abstract

:
In this paper, Berkovich depth-sensing indentation has been used to study the effects of the temperature-dependent quasi-static mechanical properties and creep deformation of heat-treated wood at temperatures from 20 °C to 180 °C. The characteristics of the load–depth curve, creep strain rate, creep compliance, and creep stress exponent of heat-treated wood are evaluated. The results showed that high temperature heat treatment improved the hardness of wood cell walls and reduced the creep rate of wood cell walls. This is mainly due to the improvement of the crystallinity of the cellulose, and the recondensation and crosslinking reaction of the lignocellulose structure. The Burgers model is well fitted to study the creep behavior of heat-treated wood cell walls under different temperatures.

1. Introduction

As a kind of complex natural polymer material, wood exhibits obvious viscoelastic properties, such as creep and stress relaxation [1,2,3]. The viscoelasticity of wood is very sensitive to changes in service temperature, moisture content and wood species [4,5,6]. Wood heat treatment can effectively improve the defects of cracking, warping deformation and easy decay of wood material [7,8]. Therefore, heat-treated wood with elegant color and dimensional stability is widely used for furniture, flooring, cladding and outdoor utilities [9], while the mechanical properties of the heat-treated wood are critical for its modification and application [10]. Most previous studies were focused on the changes of micromechanical properties of cell walls with different heat treatments that use nanoindentation [11,12]. Nanoindentation is widely used to characterize the quasi-static mechanical properties of metals and polymers at the micro scale [13,14,15]. The quasi-static mechanical properties of samples are measured by the stress–strain curves under constant load [15]. Nanoindentation is also applicable to the study of plastic deformation and dislocations of materials [16,17,18]. The creep of materials such as polymers and soft metals has been studied using data collected during the pressing process [19,20]. Nanoindentation is also applied to characterize the viscoelastic properties of biological tissue structure by extending the load-holding time [21]. In previous studies, the creep model has provided an interesting explanation for the rheological behavior of chemical-modified materials [22,23,24]. Creep is an important form of viscoelastic behavior of wood, which includes stress relaxation and dynamic viscoelasticity [25]. As a complex organic polymer, wood exhibits viscoelastic behavior due to the movement of molecular chain segments, flow within molecules, slip between molecules, crystallization and molecular orientation [26,27].
There is no published study on the micro-mechanical creep deformation of heat-treated wood under different service temperatures. This study of the viscoelasticity of heat-treated wood provides the fundamental theoretical basis for heat treatment. The purpose of this paper is to test the hypothesis that heat treatment can reduce the creep ratio of wood at high temperatures and to provide a theoretical basis for the use of heat-treated wood as wood flooring. The temperature-dependent creep characteristics of heat-treated wood subjected to sustained constant stress were measured.

2. Materials and Methods

2.1. Materials

Larch (Larix gmelinii (Rupr.) Kuzen) is a rapid planting species in artificial forests, which is widely distributed in Northern and eastern China due to its excellent adaptability. The dimensions of the sample are 10 mm, 8 mm and 8 mm (axial, radial and tangential), respectively. The bottom and top of the samples are parallel and perpendicular to the axial section of wood cell wall. Samples were precisely prepared from the same growth ring and the late wood location of the growth ring was marked. The samples were placed in an oven for 48 h conventional drying at 103 °C. All samples were then placed in moisture-proof plastic bags for high temperature heat treatment.

2.2. Heat Treatment

After a conventional drying process, the wood specimens were treated by 1′′ tube furnance (Thermo Scientific, Asheville, NC, USA) at 180 ± 1 °C and 210 ± 1 °C for 6 h. Nitrogen was used as the shielding and heat conduction gas for heat treatment and the gas flow rate was 20 mL/min. All specimens were then deposited in an oven at 103 °C to await nanoindentation testing. The specific parameters of the heat treatment process were consistent with previous studies [28]. H0 refers to the oven-dry sample, and H1 and H2 are the heat-treated samples at 180 °C and 210 °C.
Quasi-static nanoindentation:
The wood samples were fixed onto a sample holder. The samples were well prepared into a pyramid shape. Then, the tips of samples were smoothed using an ultramicrotome equipped with glass and diamond knifes. Preparation of the samples followed the steps represented by Meng [5].
A TriboIndenter (Hysitron Inc., Minneapolis, MN, USA), integrated with a Berkovich diamond tip and operated in open-loop control mode, was applied to measure the wood S2 layer cell wall laminae.
On the basis of nanoindentation, the micro-mechanical properties (hardness and elastic modulus) were obtained from the load–displacement curves:
Hardness:
H = P m a x A
Pmax is the peak load force and A is acquired by the projected contact area.
The sample’s reduced elastic modulus (Er) was obtained as follows:
E r = d P d h × 1 2 × π A
where dP/dh (stiffness) is the slope of the line of the unloading process.
The thermal stage is used to achieve a temperature control stage for different service temperatures by means of the nanoindentation test. The three-segment load ramp of the indentation experiment was performed, including: load time for 5 s, load-holding process for 30 s and unloading process for 5 s, as seen in Figure 1. The peak load force of nanoindentation was 400 μN for all tests. An indentation depth of 150~200 nm was achieved on the wood cell wall, as seen in Figure 2. Figure 2a shows the morphology of the cell wall under scanning electron microscope, and Figure 2b–d show the SPM scanning images before and after indentation to verify the validity of the indentation data. The samples were indentation tested at 20 °C, 60 °C, 100 °C, 140 °C and 180 °C for wood specimens. The nanoindentations were performed under ambient relative humidity, which was 45 ± 2% during the experimental procedure.

2.3. Viscoelastic Properties-Creep Behavior Nanoindentation Test

The short-term viscoelastic behavior of heat-treated wood cell walls was investigated under constant load. Nanoindentation equipped with a Berkovich tip was used to study the effect of the temperature-dependent hardness, modulus and short-term mechanical creep behavior from room temperature to 180 °C. For nanoindentation, the creep strain rate was obtained through the following equations. The stress can be calculated as follows.
d ε d t = 1 h   d h d t
σ = P A = P 24.5 h 2  
where d ε /dt is the creep strain rate, h is the indent depth, t is the creep time, σ is the stress, P is the constant peak loading force and A is the indent contact area. A is the projected contact area given in our present work.
Based on the curves of indent depth versus time, the creep strain rate can be obtained. The creep strain rates of samples are strongly dependent on the stress. The creep data of wood cell wall can be obtained by nanoindentation test. The strain and creep compliance of wood cell wall can be calculated as follows.
ε ( t ) = σ 0   J ( t )
Since
ε ( t ) = h ( t ) h i n
and
σ 0 = P 0 A 0
Then, Equation (3) can be rewritten as
J ( t ) = A 0 P 0 h i n   h ( t )  
Before nanoindentations, the wood samples were held at the testing temperature for 10 min to attain thermal equilibrium. The thermal stage was also added to measure the creep behavior of samples in different temperature conditions (ambient condition, 60 °C 100 °C, 140 °C and 180 °C). At least 10 measurements were repeated and the average value, that could reduce the noise of the creep curves, was calculated.
Burgers model
The Burgers model was selected to simulate the experimental data and to investigate the heat-treated and untreated wood at different temperatures, respectively. The schematic image of the Burgers model is represented in Figure 3.
By applying the Burgers model, the creep compliance can be rewritten as follows:
J ( t ) = J 0 + J 1 t + J 2 [ 1 exp ( t τ 0 B ) ]
where J0 = 1/ E e B , J1 = 1/ η 1 B , J2 = 1/ E d B and τ 0 B = η 2 B / E d B . E e B and E d B are the spring constants, while η 1 B and η 2 B are the viscosity parameters of pure viscous damper shown in Figure 3.

3. Results and Discussion

3.1. Thermal Drift

The temperature of the hot stage was monitored by a thermocouple placed on the sample surface to detect the real time surface temperature. Figure 4 shows the thermal drift rate during the nanoindentation test. The thermal drift rate of each sample gradually increases with the increase in the test temperature. After proper temperature balance is attained, all the drift rates are controlled and kept at less than 1.2 nm/s. Therefore, the thermal drift has little effect on the test results [29].

3.2. Temperature on Quasi-Static Mechanical Properties

Figure 5a,b show the reduced elastic modulus and hardness of the wood cell wall of heat-treated and untreated wood at different temperatures, where the error line represents the standard deviation of the samples. The elastic modulus and hardness of untreated samples increased with the increase in ambient temperature from 20 °C to 100 °C. The increasing trend is mainly caused by the reduction in the moisture content of the samples during the temperature increase [28]. The reduction in moisture content of wood caused the decrease in the mobility of cellulose molecular chains and chain segments, which contributes to the increase in elastic modulus and hardness [18]. The moisture content of the heat-treated wood is lower than that of the untreated wood, so the hardness and elastic modulus of the heat-treated wood are less affected by the ambient temperature (20 °C to 100 °C). The hardness and elastic modulus of H0 decreased noticeably with the service temperature being raised above 140 °C, which included the softening process of wood and the mild pyrolysis of wood components (especially the hemicellulose degradation) [7]. The gas environment of the sample in the nanoindentation test process caused the oxidation reaction of the extract and the pyrolysis of the wood’s main components [8]. Large deviation values of H0 hardness and elastic modulus at 140 °C and 180 °C also indicate that wood cell walls undergo complex chemical reactions at this time. In other words, the nanoindentation test at 180 °C can also be viewed as an air heat treatment, while the hardness and elastic modulus of H1 and H2 samples showed small fluctuations, indicating that the high temperature (140 °C~180 °C) environment has less effect on the heat-treated samples. The heat-treated wood has better high temperature resistance, which is caused by the decrease in the hemicellulose content, cellulose recondensation and cross-linking reactions of the lignin structure [5].

3.3. Temperature on Creep Behaviors

Generally, wood creep occurs under compressive stress, which is related to the variation of wood species, mechanical properties, stress level, temperature, moisture content and other factors [1,4,28]. The parameters of the creep behavior of the cell walls were obtained from the data recorded in the nanoindentation peak force-holding process. The displacement–time curve of untreated wood at different service temperatures is shown in Figure 6. The displacement increased rapidly at the beginning of loading and then slowed down gradually. This creep behavior consists of instantaneous elastic deformation, viscoelastic deformation and viscoelastic degeneration [19,20]. The maximum creep displacement of H0 occurs at 20 °C. For the H0 nanoindentation test at 180 °C, the creep displacement increased significantly, and the curve fluctuated noticeably; this is mainly caused by the degradation of hemicellulose and the softening of lignin [16]. For the high temperature nanoindentation test, the movement of the cellulose molecular chain was activated, the free volume gradually increased, and the intermolecular chain began to slip [30]. Hydrogen bonds between hemicellulose and cellulose in the amorphous region were broken, which also promoted the creep displacement [16,28].

3.4. Heat Treatment on Creep Behaviors

According to the microstructure of wood, the cell wall can be seen as complex lignocellulosic composites [13]. The nanoindentation is mainly applied to the S2 layer of the cell wall, which is the thickest part of the cell wall [1,5]. Figure 7 shows the creep compliance for different test temperature conditions (20 °C, 60 °C, 100 °C, 140 °C and 180 °C) of H0, H1 and H2. Comparing the H0, H1 and H2 curves at the same test temperature, it is clear that the heat treatment completely changes the wood cell wall creep behavior. In Figure 7, the creep compliance of the cell wall is decreased after heat treatment at room temperature (20 °C). In other words, heat treatment can effectively reduce the creep behavior of the wood cell walls at 20 °C. H2 exhibits minimal creep compliance from room temperature to 100 °C, while it represents the maximum creep compliance at 180 °C. The creep of H2 sample increases rapidly when ambient temperature exceeds 100 °C. In addition, the H0 and H1 curves overlap when the ambient temperature reaches 140 °C, which is most likely caused by hemicellulose Softening. As a complex physical modification process, heat treatment includes the degradation and modification of hemicellulose [30], the degradation and crystallization of amorphous cellulose and the condensation polymerization of lignin [31,32]. After heat treatment, the reinforced wood cell walls showed less creep, which was closely correlated with the tighter bond of the wood cell wall.

3.5. Burgers Model

Nanoindentation is widely applied in the microscopic creep deformation of materials [33,34]. This study shows that nanoindentation technology can effectively test the heat treatment of wood cell wall creep behavior. The experimental data are simulated using the four component Burgers model. Figure 8 and Figure 9 show the experimental data for wood cell wall creep compliance and the red line is creep compliance simulated by the Burgers model. The creep simulated curves are in good agreement with the test data, with a goodness of fit of 0.99. The Burgers model is suitable for forecasting the creep behavior of wood cell walls. According to nanoindentation test data, the creep parameters of wood cell walls are obtained based on Burgers model fitting. The parameters of the Burgers model include E1, E2, η1, η2 and τ0: elastic modulus, viscoelastic modulus, plasticity coefficient, viscoelasticity and relaxation time. These parameters are listed in Table 1.
The creep deformation of wood is greatly affected by heat treatment, environmental temperature and humidity [5,35]. At room temperature, the Burgers models E1, E2, η1 and η2 of the wood cell walls increased gradually with the increase in heat treatment intensity, as seen in Table 1. This demonstrates that heat treatment can reduce wood cell wall creep behavior at room temperature, which is consistent with results obtained using the quasi-static nanoindentation test [18,28]. When the service temperature increased (60~140 °C), the elastic modulus, viscoelastic modulus, plastic coefficient and viscoelasticity of the untreated specimen showed a significant increase, which is related to the softening of the wood cell wall. The creep parameters of the Burgers model of the untreated sample were significantly reduced and there was a significant fluctuation in the creep curve at 180 °C, which was caused by real-time softening and pyrolysis of the wood components during the 180 °C nanoindentation test. Compared with the untreated sample, the creep curve of H2 is relatively stable at 180 °C. The results showed that the Burgers model was well fitted to the creep curves of untreated and heat-treated Larix cell wall. The difference between the fitting data of the Burgers model and the experimental data is accurately controlled.

4. Conclusions

After reasonable temperature balance treatment, the nanoindentation overcomes the limitations of the inherent room temperature test and can be used to test the microscopic creep behavior under high temperature conditions. The hardness of the wood cell walls significantly increased and the creep rate decreased noticeably after heat treatment. The improvement in the micromechanical properties of wood after heat treatment is mainly due to the recondensation and cross-linking reaction of the lignocellulosic structure and the increase in the crystallinity of cellulose.
It has been verified that the parameters of creep behavior fitted by the Burgers model are sensitive to changes in ambient temperature. The Burgers model can adequately fit the creep of wood cell walls under temperature response.
The nanoindentation test during the peak stress holding stage also demonstrated the softening of wood components at high temperature. It was also proven that the Burgers model is appropriate to study the viscoelastic creep behavior of wood cell walls.

Author Contributions

Conceptualization, D.X., X.W. and S.W.; methodology, D.X.; software, D.X.; validation, D.X. and S.W.; formal analysis, D.X., X.W. and X.W.; investigation, D.X.; resources, D.X.; data curation, D.X. and S.W.; writing—original draft preparation, D.X.; writing—review and editing, D.X., X.W. and S.W.; visualization, D.X.; supervision, S.W.; project administration, D.X.; funding acquisition, D.X. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by the National Natural Science Foundation of China (Grant No. 31700484) and the Research Start-Up Project of Inner Mongolia Agricultural University (NDYB 2017-4).

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Wang, D.; Lin, L.; Fu, F. The difference of creep compliance for wood cell wall CML and secondary S2 layer by nanoindentation. Mech. Time-Depend. Mater. 2019, 9, 1–12. [Google Scholar] [CrossRef]
  2. Passard, J.; Perré, P. Viscoelastic behaviour of green wood across the grain. Part II. A temperature dependent constitutive model defined by inverse method. Ann. For. Sci. 2005, 62, 823–830. [Google Scholar] [CrossRef]
  3. Zhang, T.; Bai, S.L.; Zhang, Y.F.; Thibaut, B. Viscoelastic properties of wood materials characterized by nanoindentation experiments. Wood Sci. Technol. 2012, 46, 1003–1016. [Google Scholar] [CrossRef] [Green Version]
  4. Báder, M.; Németh, R.; Konnerth, J. Micromechanical properties of longitudinally compressed wood. Eur. J. Wood Wood Prod. 2019, 77, 341–351. [Google Scholar] [CrossRef] [Green Version]
  5. Meng, Y.; Xia, Y.; Young, T.M.; Cai, Z.; Wang, S. Viscoelasticity of wood cell walls with different moisture content as measured by nanoindentation. RSC Adv. 2015, 5, 47538–47547. [Google Scholar] [CrossRef]
  6. Kaufman, J.D.; Miller, G.J.; Morgan, E.F.; Klapperich, C.M. Time-dependent mechanical characterization of poly(2-hydroxyethyl methacrylate) hydrogels using nanoindentation and unconfined compression. J. Mater. Res. 2008, 23, 1472–1481. [Google Scholar] [CrossRef] [Green Version]
  7. Esteves, B.; Cruz-Lopes, L.S.; Figueirinha, A.; Lemons, L.T.; Ferreira, J.; Pereira, H.; Domingos, I. Heat-treated wood as chromium adsorption material. Eur. J. Wood Wood Prod. 2017, 75, 903–909. [Google Scholar] [CrossRef]
  8. Kufre, E.O.; Fengcai, L.; Yandan, C.; Biao, H. Effect of silicone oil heat treatment on the chemical composition, cellulose crystalline structure and contact angle of Chinese parasol wood. Carbohydr. Polym. 2017, 164, 179–185. [Google Scholar]
  9. Nguyen, T.T.; Nguyen, T.H.V.; Ji, X.D.; Yuan, B.N.; Trinh, H.M.; Do, K.T.L.; Guo, M.H. Prediction of the color change of heat-treated wood during artificial weathering by artificial neural network. Eur. J. Wood Wood Prod. 2019, 77, 1107–1116. [Google Scholar] [CrossRef]
  10. Chang, C.W.; Kuo, W.L.; Lu, K.T. On the Effect of Heat Treatments on the Adhesion, Finishing and Decay Resistance of Japanese cedar (Cryptomeria japonica D. Don) and Formosa acacia (Acacia confuse Merr. (Leguminosae)). Forests 2019, 10, 586. [Google Scholar] [CrossRef] [Green Version]
  11. Yuan, H.; Tang, S.; Luo, Q.; Xiao, T.; Wu, Y. Micro-FTIR spectroscopy and partial least-squares regression for rapid determination of moisture content of nanogram-scaled heat-treated wood. J. Wood Sci. 2020, 66, 1–7. [Google Scholar] [CrossRef]
  12. Delucis, R.; Beltrame, R.; Gatto, D.A. Discolouration of heat-treated fast-growing eucalyptus wood exposed to natural weathering. Cell Chem. Technol. 2019, 53, 635–641. [Google Scholar] [CrossRef]
  13. Rindler, A.; Hansmann, C.; Konnerth, J. The effect of moisture on the mechanical response of wood, adhesive and their interphase by means of nanoindentation. Wood Sci. Technol. 2019, 53, 729–746. [Google Scholar] [CrossRef] [Green Version]
  14. Liang, R.; Zhu, Y.H.; Yang, X.; Gao, J.S.; Cai, L.P. Study on the ultrastructure and properties of gelatinous layer in poplar. J. Mater. Sci. 2021, 56, 415–427. [Google Scholar] [CrossRef]
  15. Lucca, D.A.; Herrmann, K.; Klopfstein, M.J. Nanoindentation: Measuring methods and applications. CIRP Ann. 2010, 59, 803–819. [Google Scholar] [CrossRef]
  16. Li, Y.; Huang, C.; Wang, L.; Wang, S.; Wang, X. The effects of thermal treatment on the nanomechanical behavior of bamboo (Phyllostachys pubescens Mazel ex H. De Lehaie) cell walls observed by nanoindentation, XRD, and wet chemistry. Holzforschung 2016, 71, 129–135. [Google Scholar] [CrossRef]
  17. Xing, D.; Li, J.; Wang, S. Comparison of the chemical and micromechanical properties of Larix spp. after eco-friendly heat treatments measured by in situ nanoindentation. Sci. Rep. 2020, 10, 4358. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  18. Wang, X.; Zhao, L.; Deng, Y.; Li, Y.; Wang, S. Effect of the penetration of isocyanates (pMDI) on the nanomechanics of wood cell wall evaluated by AFM-IR and nanoindentation (NI). Holzforschung 2017, 72, 301–309. [Google Scholar] [CrossRef]
  19. Chen, Y.P.; Guo, J.P.; Cai, Y.Y. An anisotropic distribution dislocation loop model for simulation of nanoindentation of single crystals. Mech. Mater. 2017, 108, 1–10. [Google Scholar] [CrossRef]
  20. Han, R.; Wang, X.; Zhu, G.; Han, N.; Xing, F. Investigation on viscoelastic properties of urea-formaldehyde microcapsules by using nanoindentation. Polym. Test. 2019, 80, 106146. [Google Scholar] [CrossRef]
  21. Shepherd, T.N.; Zhang, J.; Ovaert, T.C.; Roeder, R.K.; Niebur, G.L. Direct comparison of nanoindentation and macroscopic measurements of bone viscoelasticity. J. Mech. Behav. Biomed. 2011, 4, 2055–2062. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  22. Hadjab, I.; Farlay, D.; Crozier, P.; Douillard, T.; Follet, H. Intrinsic Properties of Osteomalacia Bone evaluated by Nanoindentation and FTIRM Analysis. J. Biomech. 2021, 117, 110247. [Google Scholar] [CrossRef] [PubMed]
  23. Zhang, Z.N.; Tian, Z.Q.; Yuan, Z.Q.; Li, J.J.; Wu, W.P.; Xia, R. Temperature and loading sensitivity investigation of nanoindentation short-term creep behavior in Nafion (R). Mater. Res. Express 2019, 6, 055304. [Google Scholar] [CrossRef]
  24. Liu, K.; Ostadhassan, M.; Bubach, B. Application of nanoindentation to characterize creep behavior of oil shales. J. Pet. Sci. Eng. 2018, 167, 729–736. [Google Scholar] [CrossRef]
  25. Gibson, L.J. The hierarchical structure and mechanics of plant materials. J. R. Soc. Interface 2012, 9, 2749–2766. [Google Scholar] [CrossRef]
  26. Barthelat, F.; Zhen, Y.; Buehler, M.J. Structure and mechanics of interfaces in biological materials. Nat. Rev. Mater. 2016, 1, 16007. [Google Scholar] [CrossRef]
  27. Grishkewich, N.; Mohammed, N.; Tang, J.; Tam, K.C. Recent advances in the application of cellulose nanocrystals. Curr. Opin. Colloid 2017, 29, 32–45. [Google Scholar] [CrossRef]
  28. Xing, D.; Li, J.; Wang, X.; Wang, S.Q. In situ measurement of heat-treated wood cell wall at elevated temperature by nanoindentation. Ind. Crop. Prod. 2016, 87, 142–149. [Google Scholar] [CrossRef]
  29. Ben, D.B.; James, F.S. High-temperature nanoindentation testing of fused silica and other materials. Philos. Mag. A 2002, 82, 2179–2186. [Google Scholar]
  30. Stanzl-Tschegg, S.; Beikircher, W.; Loidl, D. Comparison of mechanical properties of thermally modified wood at growth ring and cell wall level by means of instrumented indentation tests. Holzforschung 2009, 63, 443–448. [Google Scholar] [CrossRef]
  31. Marcio, R.S.; José, O.B.; José, S.G.; Gilmara, O.M.; Carlito, C.J.; André, L.C.; Francisco, A.R.L. Chemical and Mechanical Properties Changes in Corymbia Citriodora Wood Submitted to Heat Treatment. Int. J. Mater. Eng. 2015, 5, 98–104. [Google Scholar]
  32. Zickler, G.A.; Schöberl, T.; Paris, O. Mechanical properties of pyrolysed wood: A nanoindentation study. Philos. Mag. 2006, 86, 1373–1386. [Google Scholar] [CrossRef]
  33. Xu, F.; Long, Z.L.; Deng, X.H.; Zhang, P. Loading rate sensitivity of nanoindentation creep behavior in a Fe-based bulk metallic glass. Trans. Nonferrous Met. Soc. China 2013, 23, 1646–1651. [Google Scholar] [CrossRef]
  34. Chen, Y.H.; Huang, J.C.; Wang, L.; Nieh, T.G. Effect of residual stresses on nanoindentation creep behavior of Zr-based bulk metallic glasses. Intermetallics 2013, 41, 58–62. [Google Scholar] [CrossRef]
  35. Mukudai, J.; Yata, S. Modeling and simulation of viscoelastic behavior (tensile strain) of wood under moisture change. Wood Sci. Technol. 1986, 20, 335–348. [Google Scholar] [CrossRef]
Figure 1. The load force–displacement curve of nanoindentation and the schematic diagram of the nanoindentation system [28].
Figure 1. The load force–displacement curve of nanoindentation and the schematic diagram of the nanoindentation system [28].
Forests 12 00968 g001
Figure 2. The image of SEM/SPM image before and after nanoindentation and the samples before and after heat treatment: (a) SEM image; (b) wood cell wall before indentation; (c) the marks of indentation; (d) after indentation; (e) wood samples before and after heat treatment.
Figure 2. The image of SEM/SPM image before and after nanoindentation and the samples before and after heat treatment: (a) SEM image; (b) wood cell wall before indentation; (c) the marks of indentation; (d) after indentation; (e) wood samples before and after heat treatment.
Forests 12 00968 g002
Figure 3. Schematic image of the Burgers model.
Figure 3. Schematic image of the Burgers model.
Forests 12 00968 g003
Figure 4. Thermal drift rate with different service temperatures.
Figure 4. Thermal drift rate with different service temperatures.
Forests 12 00968 g004
Figure 5. (a) The reduced modulus of heat-treated and untreated wood cell wall; (b) the hardness of heat-treated and untreated wood cell wall.
Figure 5. (a) The reduced modulus of heat-treated and untreated wood cell wall; (b) the hardness of heat-treated and untreated wood cell wall.
Forests 12 00968 g005
Figure 6. The creep displacement of untreated wood cell wall.
Figure 6. The creep displacement of untreated wood cell wall.
Forests 12 00968 g006
Figure 7. The effect of heat treatment on the creep compliance of the wood cell wall.
Figure 7. The effect of heat treatment on the creep compliance of the wood cell wall.
Forests 12 00968 g007
Figure 8. The test data and simulated curves of untreated wood cell wall.
Figure 8. The test data and simulated curves of untreated wood cell wall.
Forests 12 00968 g008
Figure 9. The test data and simulated curves of H2 heat-treated wood cell wall.
Figure 9. The test data and simulated curves of H2 heat-treated wood cell wall.
Forests 12 00968 g009
Table 1. Parameters of Burgers model.
Table 1. Parameters of Burgers model.
Temp (°C)No.E1
(GPa)
E2
(GPa)
η1
(GPa·S)
η2
(GPa·S)
τ0
(S)
R2
20H04.5715.33769.2347.523.100.997
H14.7127.991086.1974.202.650.996
H24.9333.371282.6088.482.650.995
60H05.6235.141848.01125.663.580.996
H15.4231.40719.4267.522.150.998
H26.9648.502403.71135.462.790.992
100H05.8642.071788.11138.733.300.996
H15.2929.881097.2481.922.740.997
H26.8347.823034.18123.772.590.992
140H06.1349.951443.00105.252.110.992
H18.1349.021315.31155.383.170.998
H26.0948.361710.5998.172.030.997
180H05.3538.65591.7270.491.820.981
H16.2544.701736.7055.561.240.990
H24.1528.561349.0172.752.550.992
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Xing, D.; Wang, X.; Wang, S. Temperature-Dependent Creep Behavior and Quasi-Static Mechanical Properties of Heat-Treated Wood. Forests 2021, 12, 968. https://0-doi-org.brum.beds.ac.uk/10.3390/f12080968

AMA Style

Xing D, Wang X, Wang S. Temperature-Dependent Creep Behavior and Quasi-Static Mechanical Properties of Heat-Treated Wood. Forests. 2021; 12(8):968. https://0-doi-org.brum.beds.ac.uk/10.3390/f12080968

Chicago/Turabian Style

Xing, Dong, Xinzhou Wang, and Siqun Wang. 2021. "Temperature-Dependent Creep Behavior and Quasi-Static Mechanical Properties of Heat-Treated Wood" Forests 12, no. 8: 968. https://0-doi-org.brum.beds.ac.uk/10.3390/f12080968

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