Next Article in Journal
Theoretical Model for Circular Concrete-Filled Steel Tubes Reinforced with Latticed Steel Angles Under Eccentric Loading
Next Article in Special Issue
Global Analysis of WELL Certification: Influence, Types of Spaces and Level Achieved
Previous Article in Journal
Low-Carbon Slag Concrete Design Optimization Method Considering the Coupled Effects of Formwork Stripping, Strength Progress, and Carbonation Durability
Previous Article in Special Issue
Effects of the Ductility Capacity on the Seismic Performance of Cross-Laminated Timber Structures Equipped with Frictional Isolators
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Enhancement of Thermal–Acoustic Properties of Pinus radiata by Impregnation of Bio-Phase-Change Materials

1
Millennium Institute on Green Ammonia, Pontificia Universidad Católica de Chile, Av. Vicuña Mackenna 4860, Macul, Santiago 7820436, Chile
2
Escuela de Construcción Civil, Facultad de Ingeniería, Pontificia Universidad Católica de Chile, Vicuña Mackenna 4860, Macul, Santiago 7820436, Chile
3
Centro Nacional de Excelencia para la Industria de la Madera (CENAMAD)—ANID Basal FB210015, Pontificia Universidad Católica de Chile, Av. Vicuña Mackenna 4860, Macul, Santiago 7810128, Chile
4
Bordeaux INP, Ecole Nationale Supérieure de Matériaux d’Agroalimentaire et de Chimie (ENSMAC), 16 Avenue Pey Berland, CEDEX, 33607 Pessac, France
5
Facultad de Química y Biología, Universidad de Santiago, Alameda 3363, Santiago 9170022, Chile
6
Departamento de Ingeniería Mecánica y Metalúrgica, Facultad de Ingeniería, Pontificia Universidad Católica de Chile, Vicuña Mackenna 4860, Santiago 7820436, Chile
7
Departamento de Ingeniería Mecánica, Facultad de Ingeniería, Universidad de Santiago, Alameda 3363, Estación Central, Santiago 9170022, Chile
*
Authors to whom correspondence should be addressed.
Buildings 2025, 15(8), 1320; https://doi.org/10.3390/buildings15081320
Submission received: 7 March 2025 / Revised: 9 April 2025 / Accepted: 11 April 2025 / Published: 16 April 2025
(This article belongs to the Special Issue Research on Timber and Timber–Concrete Buildings)

Abstract

Using fatty acids has generated significant interest in the building sector for improving energy storage in the form of latent heat. In this work, using vacuum impregnation, we analyzed the properties of a capric acid and myristic acid eutectic (83-17%) as a bio-based phase change material in Pinus radiata. The delignification of Pinus radiata samples facilitated the impregnation process, which was carried out using the Kraft pulping method. Morphological, chemical, mechanical, thermal, and acoustic impedance analyses were performed. The results revealed that impregnating PCM in Pinus radiata samples increases the thermal inertia of the impregnated samples, which is comparable to that of delignified samples. Additionally, the analyses showed no significant difference between natural and delignified samples after treatment with PCM.

1. Introduction

Wood has a good strength-to-weight ratio, is recyclable, and is relatively fast to process, all at a low cost [1]. However, its main limitation is its low thermal inertia. Its use induces large thermal fluctuations during the day and seasons, leading to discomfort and high energy consumption [2]. To optimize the energy performance of this porous material, phase-change materials (PCMs) have attracted particular interest for improving the energy efficiency of buildings, and these materials can be incorporated, for example, into wood [2,3]. They can store energy in the form of latent heat by exploiting their thermal properties. During the phase change of the material due to temperature variation, it stores (melts) or releases (crystallizes) energy in the form of latent heat [4,5,6]. The thermal properties of a PCM are related to its nature, i.e., its temperature and enthalpy of phase change, as well as its conductivity [6,7,8,9,10,11,12,13]. It is therefore of interest to develop a wood/PCM material with high thermal inertia, which implies that it would require a significant amount of energy to change its initial thermal condition [14], making it possible to improve the thermal efficiency of living spaces. Wood is not particularly effective for sound isolation; increasing its density by impregnating it with a PCM could enhance its acoustic insulating properties [15]. The effect of impregnation with PCMs has been studied in potential biomaterials for construction based on coffee waste [16], with an improvement observed in the acoustic absorption capacity, especially in the medium-frequency band, which, according to the authors, is similar to ambient noise.
PCMs are divided into three categories: organic, inorganic, and eutectic. The latter combines the first two families but is commonly regarded as a distinct group. In developing more environmentally friendly materials, particular interest is paid to organic bio-PCMs, such as fatty acids (FA), to replace paraffin derived from hydrocarbons [9,13,17]. These materials have several advantages, including their renewable nature, a remarkable enthalpy of phase change, improved thermal conductivity, and an attractive cost, making them suitable for thermal insulation in buildings. Among the most studied are capric acid (CA), lauric acid (LA), myristic acid (MA), palmitic acid (PA), and stearic acid (SA), which exhibit a temperature range of increasing state change, between 31.4 and 69 °C [13]. To obtain an adequate phase-change temperature, mixtures of these acids were studied at 20–25 °C. Eutectic compositions have enabled PCMs to achieve the desired thermal properties and have shown particular promise in applications in building materials. The study conducted by Duquesne et al. highlighted the remarkable properties of CA/MA and CA/PA eutectics [4,13]. In particular, the eutectic CA/MA with melting and crystallization temperatures of 24.14 and 21.37 °C, respectively, an enthalpy of fusion of 137.3 J/g, and a conductivity of 0.27 W/m/K is a promising PCM.
Several studies have already attempted to address this challenge by integrating this material into the wood, maintaining its thermal properties without impacting the mechanical properties of the matrix. The most common method is the vacuum impregnation of the PCM in the wood [18]. Many studies have been conducted using organic PCMs [14,19,20,21], particularly FA [21,22,23,24,25,26]. Some authors have studied delignification to increase the porosity of the wood and favor impregnation with PCM [14,26,27,28]. Lignin is the third most abundant family of compounds in wood, after cellulose and hemicellulose, and corresponds to a hydrophobic biopolymer comprising mainly p-hydroxyphenyl, guaiacyl, and syringyl [29]. It can account for up to 30% of the wood’s mass.
This work involved delignification and impregnation with a capric/myristic acid mixture in Pinus radiata samples. The delignification of Pinus radiata was studied to enhance the impregnation of the PCM, which was achieved through vacuum impregnation. Scanning electron microscopy, differential scanning calorimetry, thermal conductivity, and acoustic impedance were carried out to analyze the effect of the PCM on the wood samples.

2. Materials and Methods

2.1. Materials

The specimens used in this study were made from Pinus radiata, sourced from the southern region of Chile, with planed edges and a uniform cross-section. The specimens used had dimensions of 7 × 22 × 100 mm. Pinus radiata was chosen for its excellent availability, versatility, and profitability, which means that it is widely used for structural applications involving bending loads. Industrial applications such as paper pulp, plywood, sawn timber, and chipboard production have also established it as one of the most adaptable and versatile tree species worldwide [30]. The main direction of orientation of the samples was parallel to the fibers in the longitudinal direction. Capric acid (CA), myristic acid (MA), ammonium persulfate, and sodium sulfite (Na2SO3) were ordered from Sigma-Aldrich. Sodium hydroxide and hydrogen peroxide (H2O2) were sourced from Merck.

2.2. Delignification of Pinus Radiata

The wood samples were dried in an oven at 103 °C for 24 h. Delignification was performed in an alkaline medium [26,27], using a solution of 2.5 M of NaOH (Merck) and 0.4 M of sodium sulfite (Sigma-Aldrich, St. Louis, MO, USA), which was boiled in a reflux reactor from 1 h to 6 h, to solubilize the lignin by the method of “kraft pulping” (a dominant alkaline pulping method due to its high-quality pulp production) [31]. Later, the samples were washed in distilled water and then boiled in 2.5 M H2O2 (Merck) to remove the last traces of lignin, until the samples turned white, indicating the process was complete. The samples were then rinsed with hot distilled water and immersed in three successive baths of acetone [32]. Finally, they were dried in the open air overnight and then in the oven at 103 °C for 24 h.

2.3. Preparation of a Capric/Myristic Acid Mixture as PCM

The preparation of the eutectic point of the capric acid (CA) and myristic acid (MA) mixture followed a precise procedure to ensure an accurate composition. According to the literature [13], the eutectic point of the capric acid (CA) and myristic acid (MA) mixture was 83 CA-17 MA wt.%, with a melting temperature of approximately 24.3 °C, as verified by DSC analysis using a DSC 4000 System (PerkinElmer, Waltham, MA, USA). With an analytical balance with a precision of ±0.1 mg, the acids were individually weighed and transferred into separate beakers. Each acid was then liquefied on a hot plate at a temperature above its melting point under an extractor hood. The liquid mixtures were combined in a single beaker and thoroughly homogenized to obtain a uniform mixture.

2.4. Impregnation of the Capric/Myristic Acid Mixture as a PCM in Pinus Radiata

The impregnation method was based on the Bethell process [33], which involves immersing the wood samples in a PCM mixture using a vessel placed in a vacuum oven (Memmert VO29) at 60 °C, which maintains the wood dilated and the PCM in a liquid state. During the impregnation process, a pressure of 15 kPa was applied for 20 min using a vacuum pump to extract air from the porous matrix of the wood. Later, the vacuum pressure was released, and the sample was maintained at atmospheric pressure for 45 min. Then, a vacuum step of 60 kPa was applied for 10 min to prevent the impregnant from recoiling and percolating through the wood surface, as shown in Figure 1.
Figure 1. Impregnation program [33].
Figure 1. Impregnation program [33].
Buildings 15 01320 g001
It is worth mentioning that the Bethell (full-cell) process begins with an initial vacuum of at least −85 kPa for 15 min or more, depending on the permeability and cross-sectional dimensions of the treated wood. The preservative solution is then introduced into the impregnation vessel and penetrates the wood’s pore system. Pressure is applied (usually between 800 and 1400 kPa, above atmospheric pressure) for 1–5 h, but occasionally many hours, depending on the permeability and cross-sectional dimensions. Pressure is maintained until the wood is fully impregnated or the rate of preservative uptake by the wood becomes negligible. The pressure is then released while the preservative solution drains from the vessel. The final step is to use the final vacuum (usually about −50 kPa for 5–10 min) to prevent the back effect and the bleeding of preservatives from the surface of the treated wood; this approach is commonly used for water-based preservatives.
Once impregnation was complete, the wood samples were removed and dried with paper towels, keeping the temperature of the wood samples below the melting point. The retention property of the PCM by the wood sample wDas calculated using the following formula:
P C M r e t e n t i o n = m i m p r e g n a t e d   P i n u s   s a m p l e m d r y   P i n u s   s a m p l e V d r y   P i n u s   s a m p l e
where the mimpregnated Pinus sample represents the mass of the impregnated Pinus radiata sample, the mdry Pinus sample is the mass of the unimpregnated sample, and the Vdry Pinus sample is the volume of the unimpregnated Pinus radiata sample. The dimensions of the samples are measured in each direction of space with a caliper at three points.

2.5. Attenuated Total Reflection Fourier Transform Infrared Characterization (ATR-FTIR)

The capric/myristic acid mixtures were chemically analyzed by attenuated total reflectance spectroscopy (Shimadzu IR Spirit FTIR) at 24 °C with 40% relative humidity, averaging at 10 scans with a resolution of 1 cm−1, which allows the achievement of high-resolution scans to obtain the unique absorption bands of functional groups in the sample [33].

2.6. Differential Scanning Calorimetry

The influence of impregnating the capric/myristic acid mixture as a PCM in Pinus radiata samples was studied using differential scanning calorimetry (DSC 4000 System, PerkinElmer), allowing us to estimate the eutectic composition of the acid mixture. DSC was conducted in a N2 atmosphere of 20 mL/min, starting at 10 °C, which was maintained for 5 min. Then, the samples were heated at 40 °C for 5 min, increasing 1 °C/min. Later, the samples were cooled to 10 °C at 1 °C/min and then maintained at 10 °C for 5 min.

2.7. Mechanical Properties: Excitation Vibration Pulse

The mechanical properties of the Pinus radiata specimens were measured using the Sonelastic® device, which uses a mechanical wave to calculate the modulus of elasticity of wood according to ASTM E1876-15 [34]. Pinus radiata samples were excited by a pulse generating an acoustic wave. Compression and tension responses were produced when the sample was subjected to bending excitation. The analysis of this wave and its characteristic frequencies enabled us to determine the longitudinal elastic, flexural, and shear (G) moduli [35]. Each measurement was repeated 5 times. The mechanical properties were measured by exciting the material in the direction of the fibers. The pulse was, therefore, generated at the first end of the material, and the response was recorded at the opposite end.

2.8. Morphological Characterization

Surface analysis of Pinus radiata specimens was conducted using a field-emission scanning electron microscope (FE-SEM, Zeiss GeminiSEM 360) operated at an accelerating voltage of 5 keV. This enabled relative chemical analysis using energy-dispersive X-ray spectroscopy (EDX, Oxford Max Ultra40).
Micro-computed tomography (micro-CT) analysis of Pinus radiata samples was also conducted using a high-resolution Bruker SkyScan 1272 microtomography system (Kontich, Belgium) with a rotation of 0.4° at 80 kV and 125 mA. The system was equipped with an aluminum filter of 0.25 mm thickness and a voxel size of 12 µm resolution. Three-dimensional scanned images were obtained using CTvox software (Version 3.3.1). The images were reoriented in space using Data Viewer software v1.5.1.9 to standardize sample positioning. Quantitative assessments were performed using micro-CT and software analysis in a volume of interest (VOI). Images of each sample were obtained using visualization software. Image processing and analysis were performed with CTAn software version 1.14.4.1. Images obtained from micro-CT were corrected by removing the background.

2.9. Thermal Analysis

Non-destructive analysis was performed using a Hot Disk TPS 1500 to determine the thermal properties of each specimen, according to ISO 22007-2 [36]. The tested specimens were cylindrical samples, previously dry at room temperature (20 °C), with a diameter of 40 mm and a thickness of 20 mm. For thermogravimetric analysis (TGA) using a PerkinElmer TGA 4000, cubic samples with a thickness of 1 mm were used. The Pinus radiata samples were placed inside a crucible and stabilized for 20 min. Thermogravimetric analysis was performed under a N2 atmosphere, using a flow rate of 20 mL/min with a two-step program: a 1 min holding time at 20 °C and a temperature scan from 20 °C to 800 °C, heated at a rate of 10 °C/min.

2.10. Acoustic Measurements

The sound absorption measurement tests of the Pinus radiata samples were conducted for cylindrical samples of 40 mm and 20 mm in thickness, using an impedance tube with an internal diameter of 40 mm, equipped with 2 microphones according to ISO 10534-2 requirements and ASTM E1050 standards [37,38]. Using a sonometer positioned 5 cm from the entrance of the tube, a white noise signal was applied and adjusted to 90 Db. Then, the sound spectrum was adjusted for 1/8 frequency bands between 100 Hz and 4000 Hz. The sound absorption test had to establish the wave number and consider the sample thickness. Then, the software calculated the surface impedance and sound absorption coefficient [39]. An initial test was conducted using an acoustic isolation foam sample with the exact dimensions of the samples studied. Samples were analyzed once the sound absorption measurements were similar to those reported in references.

3. Results and Discussion

3.1. Thermo-Chemical Analysis of Capric/Myristic Acid Mixtures

DSC analysis of the capric/myristic acid mixtures allowed us to determine the eutectic point between them, analyzing the melting temperature and the enthalpy of fusion of the mixture. Table 1 shows that the mixture of 83 wt% CA and 17 wt% MA is the eutectic mixture [4]. The lowest melting temperature is approximately 22 °C, with a melting enthalpy of around 146 J/g, corresponding to the energy the mixture can store as latent heat.
Figure 2 shows the attenuated total reflectance–Fourier transform infrared (ATR-FTIR) spectra of the eutectic capric/myristic acid mixture as a PCM, revealing peaks at 2800–3000 cm−1, which correspond to the stretching vibration of O-H groups [25], and the peak associated with the carbonyl group C=O present in the PCM molecules, observed at 1700–1715 cm−1 [40]. Additionally, the ATR-FTIR spectra exhibit 940 and 1460 cm−1 peaks, corresponding to stretching and bending vibrations of C-O and C=O groups, respectively [22]. The peaks at 1245 and 670 cm−1 are associated with O-H vibrations.

3.2. Delignification and Impregnation of Pinus Radiata Using PCM

Figure 3 illustrates the gradual development of a brown color during the delignification of Pinus radiata samples, a characteristic of lignosulfonate extraction. This indicates that the lignin concentration and the intensity of the brown color increase with the delignification time.
Table 2 shows the mass, volume, and density variations in Pinus radiata samples after delignification, indicating a reduction in these properties. This can be associated with the hydrophobic molecules that compose the lignin, which protect the wood against moisture. To avoid the deformation of the wood structure, a post-treatment with acetone was applied before drying [32].
Figure 4 shows cross-sectional digital images of Pinus radiata samples, delignified and impregnated with the eutectic capric/myristic acid mixture as a PCM, revealing the increase in porosity and cracks after delignification. These cracks increased with exposure time but decreased after the impregnation process. Increasing porosity could facilitate a higher PCM loading within the sample matrix than the non-delignified Pinus radiata.

3.3. Morphological Analysis

Figure 5 shows SEM images of Pinus radiata samples, revealing the effect of delignification and impregnation with the eutectic capric/myristic acid mixture as a PCM. As can be seen, the naturally impregnated samples display partial surface coverage with the PCM. In contrast, the delignified samples exhibit a surface completely covered with the PCM [14,26]. This suggests that the delignification process enhances PCM impregnation, likely due to increased sample porosity [41,42].
Additionally, the Pinus radiata samples were analyzed using micro-computed tomography (see Figure 6), revealing that the delignification process enhanced the sample porosity, as shown in Figure 4C.
The micro-CT analysis also revealed that impregnation with the eutectic CA/MA mixture as a PCM reduced the porosity of both the natural and delignified samples. The reduction in total porosity would affect the thermal and acoustic properties of the samples, as previously mentioned, because the air in the sample bulk would be replaced with a capric and myristic acid mixture [41,42,43], as shown in Table 3.

3.4. Mechanical Tests

Figure 7A shows the effect of the delignification and impregnation of the eutectic capric/myristic acid mixture as a PCM on the longitudinal elastic, flexural, and shear moduli of Pinus radiata samples, revealing a drastic decrease after 1 h of delignification, revealing that the delignification process did not significantly influence the mechanical moduli after a longer delignification time. However, adding a capric/myristic acid mixture as a PCM reduced the longitudinal elastic moduli. Additionally, the effect of the PCM on the mechanical properties of the samples was evaluated. Figure 7B reveals that impregnation with the capric and myristic PCM does not affect the flexural and shear moduli.

3.5. Thermal Properties

Thermal conductivity plays a crucial role in evaluating the heat transfer performance of the developed samples. This study analyzed the impact of delignification and impregnation with a capric/myristic acid mixture on the thermal conductivity of cylindrical Pinus radiata samples along the longitudinal axis at 20 °C, as shown in Table 4. The natural Pinus radiata samples exhibited a thermal conductivity (k) of 0.166 W/(m·K), which increased to 0.347 W/(m·K) after PCM impregnation—a 121.68% improvement. This enhancement is attributed to the infiltration of PCM into the porous structure of Pinus radiata. Naturally, the air trapped within the wood pores acts as an insulator, reducing thermal conductivity. However, upon PCM impregnation, the air in these cavities is replaced by the PCM, which has a higher thermal conductivity than air, as reported by Temiz et al. [44]. Also, Nazari et al. [45] found an increase in thermal conductivity in wood-based samples infused with a biobased PCM, highlighting the reduction in air pockets and the formation of continuous heat transfer pathways. Additionally, the thermal conductivity of delignified PCM-impregnated samples increased by 150.68% compared to the natural sample. This improvement is likely due to the delignification process, which enhances the matrix porosity (as confirmed by micro-CT analysis), allowing for better PCM incorporation. This, in turn, facilitates more continuous phonon pathways, improving heat conduction [46]. Table 4 shows the thermal conductivity of two common building materials, fiber cement and plaster, comparable to natural wood but nearly three times lower than that of PCM-infused composites.
Incorporating PCMs into the Pinus radiata samples enhances thermal conductivity and significantly increases thermal inertia, which is determined by a material’s mass and heat capacity. As shown in Table 4, the specific heat capacity of the PCM-infused composites increases, reflecting an increase in thermal mass. The natural PCM sample and the delignified PCM sample exhibit rises by 80.31% and 122.24%, respectively. This improvement is attributed to the PCM’s ability to absorb and release thermal energy during phase transitions, leveraging its latent heat storage capacity. Consequently, the overall heat capacity of PCM composites increases, approaching that of the reference materials, plasterboard and fiber cement board.
The influence of the PCM on the thermal inertia of Pinus radiata samples aligns with findings by Nazari et al. [47], who reported that a higher thermal mass results in greater thermal inertia, allowing for more efficient thermal management during temperature fluctuations. This enhanced thermal inertia makes the developed PCM composites suitable for energy-efficient building applications, helping maintain stable indoor temperatures [48]. Vasco et al. [21] also demonstrated a rapid increase in thermal conductivity and heat capacity after impregnating Pinus radiata with octadecane. Similarly, Feng Qiu et al. [49] investigated the thermophysical properties of a novel shape-stable phase-change material (SSPCM) based on fly ash modified with lauric acid, reporting an enhancement in thermal properties due to the incorporation of the PCM.
Table 4. Variation in the thermal properties of Pinus radiata samples.
Table 4. Variation in the thermal properties of Pinus radiata samples.
Pinus radiataκ (W/(mK))C (J/m3 K)I (J/m2 s1/2 K)
Natural1.7 × 10−1 ± 7 × 10−31.5 × 105 ± 9 × 1041.5 × 102 ± 6 × 101
Natural + PCM3.5 × 10−1 ± 7 × 10−33.4 × 105 ± 6 × 1043.4 × 102 ± 5 × 101
Delignified + PCM4.2 × 10−1 ± 2 × 10−23.5 × 105 ± 7 × 1043.9 × 102 ± 4 × 101
Fiber cement board [50]1.7 × 10−1 ± 2 × 10−36.9 × 105 ± 1 × 103
Plasterboard [50]1.4 × 10−1 ± 1 × 10−37.7 × 105 ± 1 × 103
The TGA of Pinus radiata samples was conducted to assess the effect of the eutectic capric/myristic acid mixture as a PCM, as shown in Figure 8. The natural sample (red curve) exhibited an initial weight loss below 100 °C due to moisture evaporation, followed by a significant mass reduction between 200 °C and 400 °C, corresponding to the degradation of hemicellulose, cellulose, and lignin, leaving behind a notable char residue [49]. The pure PCM (black curve) began degrading between 175 °C and 250 °C, indicating the volatilization and breakdown of fatty acids, consistent with the typical degradation behavior of fatty acid-based PCMs [51]. The delignified wood (purple curve) exhibited a two-stage, hemicellulose breakdown between 200 °C and 315 °C, followed by cellulose decomposition between 315 °C and 400 °C. Notably, the delignified sample exhibited a lower degradation temperature and a higher char residue. This reduction in thermal stability is attributed to the extraction steps involved in the delignification process [52]. However, despite this degradation, studies suggest that delignified wood forms a more stable carbonaceous residue due to crosslinking or dehydration reactions in cellulose, increasing char yield [53]. The PCM-impregnated Pinus radiata samples (green and blue curves) exhibited degradation onset between 175 °C and 250 °C, likely due to the decomposition of the PCM. Notably, the weight loss in the impregnated natural sample closely matched that in the untreated sample, suggesting that the PCM remains stable within the wood matrix and does not degrade it as in a pure PCM. Hanif et al. [51] observed that encapsulating a PCM within a porous wood structure improved its thermal stability. However, the delignified PCM-impregnated sample exhibited lower char formation than the natural impregnated sample, likely due to its increased porosity, facilitating oxygen penetration and accelerating degradation. Overall, the higher residual weight of the impregnated samples compared to a pure PCM suggests reduced volatilization and enhanced PCM retention within the wood structure. This stabilization of the PCM within the natural matrix highlights its potential for energy-efficient construction and insulation applications.

3.6. Acoustic Measurements

Figure 9 shows the sound absorption coefficient (SAC) variation of Pinus radiata samples resulting from delignification and impregnation with the eutectic capric/myristic acid mixture, as determined using an acoustic impedance analyzer. As can be seen, the sound absorption coefficient of the natural Pinus radiata samples varied with the frequency range, reaching a maximum SAC value at 2000 Hz, especially for the delignified sample impregnated with the PCM, which is higher compared to isolator materials used in construction applications, such as plywood panels.
It is worth mentioning that at every frequency, the delignified impregnated sample exhibits higher SAC values than the natural and natural impregnated samples, due to the higher porosity present in the delignified and impregnated sample, as shown in Table 5. This response can be attributed to the porosity absorbing the acoustic waves, transforming them into heat [54]. As a consequence, the sound is dispersed [39,55,56,57], which can enhance the sound absorption efficiency. Lashgari et al. [58] determined that the absorption coefficient increased at lower frequencies by enhancing the bulk density of wood chip samples. Beyond a certain density level, the porosity of the samples diminished, and as a consequence, the SAC decreased, explaining the acoustic behavior of the delignified sample. Table 5 also illustrates the impact of impregnation and delignification on the density of the Pinus radiata sample.
The SAC values for the three samples analyzed at 2000 Hz varied between 0.205 and 0.275, similar to those reported by Khrystoslavenko et al. [59] for different types of wood. The authors noted the average SAC values of common wood types, birch (B. pendula), pine (P. sylvestris), and oak (Q. robur), using thicknesses of 10 mm, 18 mm, 25 mm, and 30 mm. The SAC values were close to 0.16 at 2000 Hz for 30 mm birch, 0.13 at 2000 Hz for 18 mm birch, 0.15 at 4000 Hz for 25 mm pine, and 0.12 at 5000 Hz for 30 mm birch. It is worth mentioning that the SAC values not only depend on the dimensions of samples and density but also on the structural properties, such as cellulose crystallinity, as demonstrated by Kolya et al. [60].
It has been reported that incorporating a PCM can influence the sound absorption coefficient, depending on the analyzed frequency range and concentration. For instance, Choi et al. [16] analyzed the influence of an MPCM on the sound absorption coefficient of DCWB samples using an acoustic impedance tube test at low, medium, and high frequencies. At low frequencies, the sound absorption coefficient decreased with the addition of the MPCM. At the mid-frequency, the DCWB 3, 5, and 10 wt% samples exhibited a higher sound absorption coefficient than DCWB 0%. The authors reported that samples without the MPCM exhibited better sound absorption coefficients at high frequencies. Furthermore, the authors proposed that the MPCM favored a low sound absorption coefficient as the density increased.

4. Conclusions

This work involved a delignification process and impregnation with a capric/myristic acid mixture on Pinus radiata samples. The delignification of Pinus radiata was studied to improve the impregnation of the PCM, which was achieved by vacuum impregnation. Micro-CT analysis highlights the role of the delignification process before the impregnation due to the increase in porosity. Furthermore, the mechanical properties were not significantly modified by impregnation using capric acid/myristic acid as a phase-change material, regardless of whether the samples were delignified. Thermal analysis revealed that impregnating Pinus radiata with the capric acid/myristic acid phase-change material significantly enhanced the thermal inertia. Therefore, Pinus radiata can store more thermal energy using a mixture of capric acid and myristic acid as a phase-change material. The sound absorption coefficient of Pinus radiata increases when using capric acid and myristic acid as a phase-change material, as the density increases and the porosity diminishes after impregnation.

Author Contributions

Conceptualization, P.M., G.R.-G. and M.S.; methodology, P.M., G.S., D.C., I.E., I.A.U.-P., G.R.-G., C.S.-V., D.D.M. and M.S.; validation, P.M., D.A.V., G.R.-G. and M.S.; formal analysis, P.M., K.G., D.A.V. and M.S.; investigation, P.M. and M.S.; resources and data curation, P.M., G.S., D.C., I.E., I.A.U.-P. and K.G.; writing—original draft preparation, P.M., K.G., G.R.-G. and M.S.; writing—review and editing, P.M., K.G., D.A.V., G.R.-G. and M.S.; visualization, P.M., K.G. and M.S.; supervision, P.M., G.R.-G. and M.S.; project administration, P.M. and M.S.; funding acquisition, M.S. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by ANID BASAL FB210015 CENAMAD, and FONDECYT Regular 1201520.

Data Availability Statement

The authors confirm that the data supporting the findings of this study are available within the article.

Acknowledgments

This study was supported by ANID BASAL FB210015 CENAMAD, UC Grant Investigación Interdisciplinaria II190060, and FONDECYT REGULAR 1201520.

Conflicts of Interest

The authors declare no conflicts of interest.

Abbreviations

The following abbreviations are used in this manuscript:
PCMPhase-change material
CACapric acid
LALauric acid
MAMyristic acid
PAPalmitic acid
SAStearic acid
ATR-FTIRAttenuated total reflectance–Fourier transform infrared
SACSound absorption coefficient

References

  1. Hildebrandt, J.; Hagemann, N.; Thrän, D. The Contribution of Wood-Based Construction Materials for Leveraging a Low Carbon Building Sector in Europe. Sustain. Cities Soc. 2017, 34, 405–418. [Google Scholar] [CrossRef]
  2. Li, B.; Chhaybi, I.; Rodrigue, D.; Berg, S.; Sandberg, D.; Huber, J.; Li, D.; Wang, X. Phase Change Materials to Improve the Energy Savings of Wood Building Envelopes in Quebec. Wood Mater. Sci. Eng. 2024, 20, 217–231. [Google Scholar] [CrossRef]
  3. Gielen, D.; Boshell, F.; Saygin, D.; Bazilian, M.D.; Wagner, N.; Gorini, R. The Role of Renewable Energy in the Global Energy Transformation. Energy Strategy Rev. 2019, 24, 38–50. [Google Scholar] [CrossRef]
  4. Duquesne, M.; Mailhé, C.; Doppiu, S.; Dauvergne, J.L.; Santos-Moreno, S.; Godin, A.; Fleury, G.; Rouault, F.; Del Barrio, E.P. Characterization of Fatty Acids as Biobased Organic Materials for Latent Heat Storage. Materials 2021, 14, 4707. [Google Scholar] [CrossRef]
  5. Berardi, U.; Gallardo, A.A. Properties of Concretes Enhanced with Phase Change Materials for Building Applications. Energy Build. 2019, 199, 402–414. [Google Scholar] [CrossRef]
  6. Sari, A. Eutectic Mixtures of Some Fatty Acids for Low Temperature Solar Heating Applications: Thermal Properties and Thermal Reliability. Appl. Therm. Eng. 2005, 25, 2100–2107. [Google Scholar] [CrossRef]
  7. Gunasekara, S.N.; Martin, V.; Chiu, J.N. Phase Equilibrium in the Design of Phase Change Materials for Thermal Energy Storage: State-of-the-Art. Renew. Sustain. Energy Rev. 2017, 73, 558–581. [Google Scholar] [CrossRef]
  8. Kenisarin, M.M. Thermophysical Properties of Some Organic Phase Change Materials for Latent Heat Storage. A Review. Solar Energy 2014, 107, 553–575. [Google Scholar] [CrossRef]
  9. Soodoo, N.; Poopalam, K.D.; Bouzidi, L.; Narine, S.S. Fundamental Structure-Function Relationships in Vegetable Oil Based Phase Change Materials: A Critical Review. J. Energy Storage 2022, 51, 104355. [Google Scholar] [CrossRef]
  10. Hassan, F.; Jamil, F.; Hussain, A.; Ali, H.M.; Janjua, M.M.; Khushnood, S.; Farhan, M.; Altaf, K.; Said, Z.; Li, C. Recent advancements in latent heat phase change materials and their applications for thermal energy storage and buildings: A state of the art review. Sustain. Energy Technol. Assess. 2022, 49, 101646. [Google Scholar] [CrossRef]
  11. Jiang, Z.; Palacios, A.; Zou, B.; Zhao, Y.; Deng, W.; Zhang, X.; Ding, Y. A Review on the Fabrication Methods for Structurally Stabilised Composite Phase Change Materials and Their Impacts on the Properties of Materials. Renew. Sustain. Energy Rev. 2022, 159, 112134. [Google Scholar] [CrossRef]
  12. Yan, Q. The Thermal Properties of Shape-Stabilized Fatty Acid Mixtures Used for Wallboard. Int. J. Sustain. Energy 2011, 30, 47–54. [Google Scholar] [CrossRef]
  13. Duquesne, M.; Mailhé, C.; Ruiz-Onofre, K.; Achchaq, F. Biosourced Organic Materials for Latent Heat Storage: An Economic and Eco-Friendly Alternative. Energy 2019, 188, 116067. [Google Scholar] [CrossRef]
  14. Chen, H.; Xuan, J.; Deng, Q.; Gao, Y. WOOD/PCM Composite with Enhanced Energy Storage Density and Anisotropic Thermal Conductivity. Prog. Nat. Sci. Mater. Int. 2022, 32, 190–195. [Google Scholar] [CrossRef]
  15. Asdrubali, F.; Ferracuti, B.; Lombardi, L.; Guattari, C.; Evangelisti, L.; Grazieschi, G. A Review of Structural, Thermo-Physical, Acoustical, and Environmental Properties of Wooden Materials for Building Applications. Build. Environ. 2017, 114, 307–332. [Google Scholar] [CrossRef]
  16. Choi, J.Y.; Yun, B.Y.; Kim, Y.U.; Kang, Y.; Lee, S.C.; Kim, S. Evaluation of Thermal/Acoustic Performance to Confirm the Possibility of Coffee Waste in Building Materials in Using Bio-Based Microencapsulated PCM. Environ. Pollut. 2022, 294, 118616. [Google Scholar] [CrossRef]
  17. Mehrizi, A.A.; Karimi-Maleh, H.; Naddafi, M.; Karimi, F. Application of Bio-Based Phase Change Materials for Effective Heat Management. J. Energy Storage 2023, 61, 106859. [Google Scholar] [CrossRef]
  18. Li, D.; Zhuang, B.; Chen, Y.; Li, B.; Landry, V.; Kaboorani, A.; Wu, Z.; Wang, X.A. Incorporation Technology of Bio-Based Phase Change Materials for Building Envelope: A Review. Energy Build. 2022, 260, 111920. [Google Scholar] [CrossRef]
  19. Nam, J.; Yun, B.Y.; Choi, J.Y.; Kim, S. Potential of Wood as Thermal Energy Storage Materials: Different Characteristics Depending on the Hierarchical Structure and Components. Int. J. Energy Res. 2022, 46, 14926–14945. [Google Scholar] [CrossRef]
  20. Barreneche, C.; Vecstaudza, J.; Bajare, D.; Fernandez, A.I. PCM/Wood Composite to Store Thermal Energy in Passive Building Envelopes. In IOP Conference Series: Materials Science and Engineering; IOP Publishing: Bristol, UK, 2017; Volume 251. [Google Scholar] [CrossRef]
  21. Vasco, D.A.; Salinas-Lira, C.; Barra-Reyes, I.; Elustondo, D.M. Kinematic Characterization of the Pressure-Dependent PCM Impregnation Process for Radiata Pine Wood Samples. Eur. J. Wood Wood Prod. 2018, 76, 1461–1469. [Google Scholar] [CrossRef]
  22. Sarı, A.; Hekimoğlu, G.; Tyagi, V.V. Low Cost and Eco-Friendly Wood Fiber-Based Composite Phase Change Material: Development, Characterization and Lab-Scale Thermoregulation Performance for Thermal Energy Storage. Energy 2020, 195, 116983. [Google Scholar] [CrossRef]
  23. Ma, L.; Guo, C.; Ou, R.; Sun, L.; Wang, Q.; Li, L. Preparation and Characterization of Modified Porous Wood Flour/Lauric-Myristic Acid Eutectic Mixture as a Form-Stable Phase Change Material. Energy Fuels 2018, 32, 5453–5461. [Google Scholar] [CrossRef]
  24. Hussain, S.I.; Kalaiselvam, S. Nanoencapsulation of Oleic Acid Phase Change Material with Ag2O Nanoparticles-Based Urea Formaldehyde Shell for Building Thermal Energy Storage. J. Therm. Anal. Calorim. 2020, 140, 133–147. [Google Scholar] [CrossRef]
  25. Sarı, A.; Bicer, A.; Alkan, C.; Özcan, A.N. Thermal Energy Storage Characteristics of Myristic Acid-Palmitic Eutectic Mixtures Encapsulated in PMMA Shell. Sol. Energy Mater. Sol. Cells 2019, 193, 1–6. [Google Scholar] [CrossRef]
  26. Ma, L.; Wang, Q.; Li, L. Delignified Wood/Capric Acid-Palmitic Acid Mixture Stable-Form Phase Change Material for Thermal Storage. Sol. Energy Mater. Sol. Cells 2019, 194, 215–221. [Google Scholar] [CrossRef]
  27. Kumar, A.; Jyske, T.; Petrič, M. Delignified Wood from Understanding the Hierarchically Aligned Cellulosic Structures to Creating Novel Functional Materials: A Review. Adv. Sustain. Syst. 2021, 5, 2000251. [Google Scholar] [CrossRef]
  28. Zhu, M.; Song, J.; Li, T.; Gong, A.; Wang, Y.; Dai, J.; Yao, Y.; Luo, W.; Henderson, D.; Hu, L. Highly Anisotropic, Highly Transparent Wood Composites. Adv. Mater. 2016, 28, 5181–5187. [Google Scholar] [CrossRef]
  29. Vanholme, R.; De Meester, B.; Ralph, J.; Boerjan, W. Lignin Biosynthesis and Its Integration into Metabolism. Curr. Opin. Biotechnol. 2019, 56, 230–239. [Google Scholar] [CrossRef]
  30. Véliz-Fadic, F.I.; Rodríguez-Grau, G.; Marín-Uribe, C.R.; García-Giraldo, J.M.; González-Palacio, L.; Araya-Letelier, G. Flexural Performance Assessment of the Effect of the Splice Length of the Jupiter Ray Type Made of Radiata Pine Using Computer-Aided Design and Computer-Assisted Manufacturing. Constr. Build. Mater. 2024, 449, 138272. [Google Scholar] [CrossRef]
  31. Fearon, O.; Kuitunen, S.; Ruuttunen, K.; Alopaeus, V.; Vuorinen, T. Detailed Modeling of Kraft Pulping Chemistry. Delignification. Ind. Eng. Chem. Res. 2020, 59, 12977–12985. [Google Scholar] [CrossRef]
  32. Han, X.; Ye, Y.; Lam, F.; Pu, J.; Jiang, F. Hydrogen-Bonding-Induced Assembly of Aligned Cellulose Nanofibers into Ultrastrong and Tough Bulk Materials. J. Mater. Chem. 2019, 7, 27023–27031. [Google Scholar] [CrossRef]
  33. Fuentes-Sepúlveda, R.; García-Herrera, C.; Vasco, D.A.; Salinas-Lira, C.; Ananías, R.A. Thermal Characterization of Pinus Radiata Wood Vacuum-Impregnated with Octadecane. Energies 2020, 13, 942. [Google Scholar] [CrossRef]
  34. Otani, L.B.; de Alcântara Segundinho, P.G.; Morales, E.A.M.; Pereira, A.H.A. Elastic Moduli Characterization of Wood and Wood Products Using the Impulse Excitation Technique. ATCP Phys. Eng. 2015, 1, 33. [Google Scholar] [CrossRef]
  35. Baradit, E.; Fuentealba, C.; Yáñez, M. Elastic constants of Chilean Pinus radiata using ultrasound. Maderas. Cienc. Tecnol. 2021, 23, 1–10. [Google Scholar] [CrossRef]
  36. ISO 22007-2:2008(E); Plastics-Determination of Thermal Conductivity and Thermal Diffusivity-Part 2:Transient Plane Heat Source (Hot Disc) Method. The International Organization for Standardization: Geneva, Switzerland, 2008.
  37. ASTM E1050-19; Standard Test Method for Impedance and Absorption of Acoustical Materials Using a Tube, Two Microphones and a Digital Frequency Analysis System. ASTM International: West Conshohocken, PA, USA, 2005. [CrossRef]
  38. Tao, J.; Wang, P.; Qiu, X.; Pan, J. Static Flow Resistivity Measurements Based on the ISO 10534.2 Standard Impedance Tube. Build. Environ. 2015, 94, 853–858. [Google Scholar] [CrossRef]
  39. Mohammadi, M.; Taban, E.; Tan, W.H.; Din, N.B.C.; Putra, A.; Berardi, U. Recent Progress in Natural Fiber Reinforced Composite as Sound Absorber Material. J. Build. Eng. 2024, 84, 108514. [Google Scholar] [CrossRef]
  40. Steinhaus, T. Evaluation of the Thermophysical Properties of Poly(MethylMethacrylate): A Reference Material for the Development of a Flammability Test for Micro-Gravity Environments; The University of Maryland: College Park, MD, USA, 2009. [Google Scholar]
  41. Hung Anh, L.D.; Pásztory, Z. An Overview of Factors Influencing Thermal Conductivity of Building Insulation Materials. J. Build. Eng. 2021, 44, 102604. [Google Scholar] [CrossRef]
  42. Cao, L.; Fu, Q.; Si, Y.; Ding, B.; Yu, J. Porous Materials for Sound Absorption. Compos. Commun. 2018, 10, 25–35. [Google Scholar] [CrossRef]
  43. Peng, L. Sound Absorption and Insulation Functional Composites. In Advanced High Strength Natural Fibre Composites in Construction; Woodhead Publishing: Sawston, UK, 2017; pp. 333–373. [Google Scholar] [CrossRef]
  44. Temiz, A.; Hekimoğlu, G.; Köse Demirel, G.; Sarı, A.; Mohamad Amini, M.H. Phase Change Material Impregnated Wood for Passive Thermal Management of Timber Buildings. Int. J. Energy Res. 2020, 44, 10495–10505. [Google Scholar] [CrossRef]
  45. Nazari, M.; Jebrane, M.; Terziev, N. Solid Wood Impregnated with a Bio-Based Phase Change Material for Low Temperature Energy Storage in Building Application. J. Therm. Anal. Calorim. 2022, 147, 10677–10692. [Google Scholar] [CrossRef]
  46. Khakalo, A.; Tanaka, A.; Korpela, A.; Orelma, H. Delignification and Ionic Liquid Treatment of Wood toward Multifunctional High-Performance Structural Materials. ACS Appl. Mater. Interfaces 2020, 12, 23532–23542. [Google Scholar] [CrossRef] [PubMed]
  47. Nazari, M.; Jebrane, M.; Gao, J.; Terziev, N. Thermal Performance and Mold Discoloration of Thermally Modified Wood Containing Bio-Based Phase Change Material for Heat Storage. Energy Storage 2022, 4, e340. [Google Scholar] [CrossRef]
  48. Soret, G.M.; Vacca, P.; Tignard, J.; Hidalgo, J.P.; Maluk, C.; Aitchison, M.; Torero, J.L. Thermal Inertia as an Integrative Parameter for Building Performance. J. Build. Eng. 2021, 33, 101623. [Google Scholar] [CrossRef]
  49. Chen, Y.; Tshabalala, M.A.; Gao, J.; Stark, N.M.; Fan, Y.; Ibach, R.E. Thermal Behavior of Extracted and Delignified Pine Wood Flour. Thermochim. Acta 2014, 591, 40–44. [Google Scholar] [CrossRef]
  50. Vasco, D.A.; Muñoz-Mejías, M.; Pino-Sepúlveda, R.; Ortega-Aguilera, R.; García-Herrera, C. Thermal Simulation of a Social Dwelling in Chile: Effect of the Thermal Zone and the Temperature-Dependant Thermophysical Properties of Light Envelope Materials. Appl. Therm. Eng. 2017, 112, 771–783. [Google Scholar] [CrossRef]
  51. Hanif, F.; Imran, M.; Zhang, Y.; Jia, Z.; Lu, X.; Lu, R.; Tang, B. Form-Stable Phase Change Material with Wood-Based Materials as Support. Polymers 2023, 15, 942. [Google Scholar] [CrossRef]
  52. Chen, Y.; Tshabalala, M.A.; Stark, N.M.; Gao, J.; Fan, Y.; Ibach, R.E. Thermal Properties of Extracted and Delignified Wood Flour for Use in Wood-Plastic Composites. In Proceedings of the Advancements in Fiber-Polymer Composites: Wood Fiber, Natural Fibers and Nanocellulose, Milwaukee, WI, USA, 6–7 May 2013. [Google Scholar]
  53. Li, H.; Guo, X.; He, Y.; Zheng, R. A Green Steam-Modified Delignification Method to Prepare Low-Lignin Delignified Wood for Thick, Large Highly Transparent Wood Composites. J. Mater. Res. 2019, 34, 932–940. [Google Scholar] [CrossRef]
  54. Amares, S.; Sujatmika, E.; Hong, T.W.; Durairaj, R.; Hamid, H.S.H.B. A Review: Characteristics of Noise Absorption Material. J. Phys. Conf. Ser. 2017, 908, 012005. [Google Scholar] [CrossRef]
  55. Asdrubali, F. Survey on the Acoustical Properties of New Sustainable Materials for Noise Control. In Proceedings of Euronoise; Acoustical Society of Finland: Tampere, Finland, 2006; Volume 30, pp. 1–10. [Google Scholar]
  56. Iannace, G.; Ciaburro, G. Modelling Sound Absorption Properties for Recycled Polyethylene Terephthalate-Based Material Using Gaussian Regression. Build. Acoust. 2021, 28, 185–196. [Google Scholar] [CrossRef]
  57. Ciaburro, G.; Iannace, G.; Ricciotti, L.; Apicella, A.; Perrotta, V.; Aversa, R. Acoustic Applications of a Foamed Geopolymeric-Architected Metamaterial. Appl. Sci. 2024, 14, 1207. [Google Scholar] [CrossRef]
  58. Lashgari, M.; Taban, E.; SheikhMozafar, M.J.; Soltan, P.; Attenborough, K.; Khavanin, A. Wood Chip Sound Absorbers: Measurements and Models. Appl. Acoust. 2024, 220, 109963. [Google Scholar] [CrossRef]
  59. Khrystoslavenko, O.; Astrauskas, T.; Grubliauskas, R. Sound Absorption Properties of Charcoal Made from Wood Waste. Sustainability 2023, 15, 8196. [Google Scholar] [CrossRef]
  60. Kolya, H.; Kang, C.W. High Acoustic Absorption Properties of Hackberry Compared to Nine Different Hardwood Species: A Novel Finding for Acoustical Engineers. Appl. Acoust. 2020, 169, 107475. [Google Scholar] [CrossRef]
  61. Amran, M.; Fediuk, R.; Murali, G.; Vatin, N.; Al-Fakih, A. Sound-Absorbing Acoustic Concretes: A Review. Sustainability 2021, 13, 10712. [Google Scholar] [CrossRef]
Figure 2. ATR-FTIR spectra of capric/myristic acid mixture as PCM.
Figure 2. ATR-FTIR spectra of capric/myristic acid mixture as PCM.
Buildings 15 01320 g002
Figure 3. Evolution of the concentration of lignin in solution during delignification.
Figure 3. Evolution of the concentration of lignin in solution during delignification.
Buildings 15 01320 g003
Figure 4. Digital images of Pinus radiata samples: (A) natural, (B) natural and impregnated with PCM, (C) delignified, and (D) delignified and impregnated with PCM.
Figure 4. Digital images of Pinus radiata samples: (A) natural, (B) natural and impregnated with PCM, (C) delignified, and (D) delignified and impregnated with PCM.
Buildings 15 01320 g004
Figure 5. SEM images of Pinus radiata samples: (A,D) natural, (B,E) natural and impregnated with PCM, and (C,F) delignified and impregnated with PCM.
Figure 5. SEM images of Pinus radiata samples: (A,D) natural, (B,E) natural and impregnated with PCM, and (C,F) delignified and impregnated with PCM.
Buildings 15 01320 g005
Figure 6. Micro-CT images of Pinus radiata samples: (A) natural, (B) delignified, (C,D) natural and delignified, impregnated with PCM.
Figure 6. Micro-CT images of Pinus radiata samples: (A) natural, (B) delignified, (C,D) natural and delignified, impregnated with PCM.
Buildings 15 01320 g006
Figure 7. Evolution of the mechanical properties of Pinus radiata samples. (A) Effect of delignification time on (blue) longitudinal elastic modulus, (green) flexural modulus, and (orange) shear modulus. (B) Influence of impregnation with PCM on the (red) shear and (black) elastic moduli.
Figure 7. Evolution of the mechanical properties of Pinus radiata samples. (A) Effect of delignification time on (blue) longitudinal elastic modulus, (green) flexural modulus, and (orange) shear modulus. (B) Influence of impregnation with PCM on the (red) shear and (black) elastic moduli.
Buildings 15 01320 g007
Figure 8. TGA images of Pinus radiata samples. () Natural, () PCM, () delignified Natural, () natural impregnated with PCM, and () Delignified impregnated with PCM.
Figure 8. TGA images of Pinus radiata samples. () Natural, () PCM, () delignified Natural, () natural impregnated with PCM, and () Delignified impregnated with PCM.
Buildings 15 01320 g008
Figure 9. Sound absorption coefficients for the Pinus radiata samples.
Figure 9. Sound absorption coefficients for the Pinus radiata samples.
Buildings 15 01320 g009
Table 1. Thermal properties of some fatty acids and their eutectics.
Table 1. Thermal properties of some fatty acids and their eutectics.
Acid MixturePeak Temperature (°C)Melting Temperature (°C)Melting Enthalpy (J/g)
MA55.5753.86151
CA32.9531.55133
82%MA-18%CA25.0422.36139
83%MA-17%CA 25.1822.09146
84%MA-16%CA25.2722.39113
Table 2. Variation in physical parameters of Pinus radiata samples as a function of the delignification time.
Table 2. Variation in physical parameters of Pinus radiata samples as a function of the delignification time.
Delignification Time (h)Mass Variation (%)Volume Variation (%)Density Variation (%)
1−10.98 ± 0.3−9.01 ± 2.1−2.3 ± 2.5
2−13.98 ± 0.7−15.03 ± 1.5+1.35 ± 1.9
3−14.62 ± 1.4−16.95 ± 1.8+2.83 ± 2.8
6−26.21 ± 0.5−17.53 ± 1.5−10.51 ± 1.6
Table 3. Variation in porosity of Pinus radiata samples.
Table 3. Variation in porosity of Pinus radiata samples.
SampleTotal Porosity (%)
Natural62.37 ± 2.24
Delignified71.75 ± 0.43
Natural + PCM48.67 ± 5.31
Delignified + PCM65.53 ± 2.37
Table 5. Variation in the density of Pinus radiata samples.
Table 5. Variation in the density of Pinus radiata samples.
Pinus radiata SampleSound Absorption Coefficient at 2000 HzDensity
(kg/m3)
Natural0.2053.81 × 10+2 ± 1 × 10+1
Natural impregnated0.2157.19 × 10+2 ± 5 × 100
Delignified and impregnated0.2759.18 × 10+2 ± 1 × 10+1
B. pendula (30 mm thick) [59]0.16~2.34 × 10+2
P. sylvestris (30 mm thick) [59]<0.15~2.63 × 10+2
Hard maple [60]0.2594.52 × 10+2
Silver poplar [60]0.4324.76 × 10+2
Plywood panel [54]0.1-
Fiberglass board [54]0.95-
Painted concrete [61]0.020-
Styrofoam 100 kg/m2 [61]0.22-
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

Molina, P.; Sancy, M.; Sève, G.; Córdova, D.; Erazo, I.; Sepúlveda-Vásquez, C.; Di Mambro, D.; George, K.; Urzúa-Parra, I.A.; Vasco, D.A.; et al. Enhancement of Thermal–Acoustic Properties of Pinus radiata by Impregnation of Bio-Phase-Change Materials. Buildings 2025, 15, 1320. https://doi.org/10.3390/buildings15081320

AMA Style

Molina P, Sancy M, Sève G, Córdova D, Erazo I, Sepúlveda-Vásquez C, Di Mambro D, George K, Urzúa-Parra IA, Vasco DA, et al. Enhancement of Thermal–Acoustic Properties of Pinus radiata by Impregnation of Bio-Phase-Change Materials. Buildings. 2025; 15(8):1320. https://doi.org/10.3390/buildings15081320

Chicago/Turabian Style

Molina, Paulo, Mamié Sancy, Gabrielle Sève, Deborah Córdova, Ignacio Erazo, Carlos Sepúlveda-Vásquez, David Di Mambro, Kesiya George, Ignacio A. Urzúa-Parra, Diego A. Vasco, and et al. 2025. "Enhancement of Thermal–Acoustic Properties of Pinus radiata by Impregnation of Bio-Phase-Change Materials" Buildings 15, no. 8: 1320. https://doi.org/10.3390/buildings15081320

APA Style

Molina, P., Sancy, M., Sève, G., Córdova, D., Erazo, I., Sepúlveda-Vásquez, C., Di Mambro, D., George, K., Urzúa-Parra, I. A., Vasco, D. A., & Rodríguez-Grau, G. (2025). Enhancement of Thermal–Acoustic Properties of Pinus radiata by Impregnation of Bio-Phase-Change Materials. Buildings, 15(8), 1320. https://doi.org/10.3390/buildings15081320

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