Next Article in Journal
Two-Channel VO2 Memory Meta-Device for Terahertz Waves
Next Article in Special Issue
An Effective Photocatalytic Degradation of Industrial Pollutants through Converting Titanium Oxide to Magnetic Nanotubes and Hollow Nanorods by Kirkendall Effect
Previous Article in Journal
The Regulation of O2 Spin State and Direct Oxidation of CO at Room Temperature Using Triboelectric Plasma by Harvesting Mechanical Energy
Previous Article in Special Issue
Bactericidal Activity of Multilayered Hybrid Structures Comprising Titania Nanoparticles and CdSe Quantum Dots under Visible Light
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

The Influence of Nanoparticles on Fire Retardancy of Pedunculate Oak Wood

1
Faculty of Wood Sciences and Technology, Technical University in Zvolen, T.G. Masaryka 24, 96001 Zvolen, Slovakia
2
Faculty of Civil Engineering, Technical University of Košice, Vysokoškolská 4, 04200 Košice, Slovakia
3
Faculty of Forestry, Technical University in Zvolen, T.G. Masaryka 24, 96001 Zvolen, Slovakia
4
Faculty of Natural Sciences, Comenius University in Bratislava, Ilkovičova 6, 84215 Bratislava, Slovakia
*
Author to whom correspondence should be addressed.
Nanomaterials 2021, 11(12), 3405; https://doi.org/10.3390/nano11123405
Submission received: 15 November 2021 / Revised: 1 December 2021 / Accepted: 13 December 2021 / Published: 16 December 2021
(This article belongs to the Special Issue Nano-TiO2: Characterization and Application)

Abstract

:
Traditional flame retardants often contain halogens and produce toxic gases when burned. Hence, in this study, low-cost, environmentally friendly compounds that act as fire retardants are investigated. These materials often contain nanoparticles, from which TiO2 and SiO2 are the most promising. In this work, pedunculate oak wood specimens were modified with sodium silicate (Na2SiO3, i.e., water glass) and TiO2, SiO2, and ZnO nanoparticles using the vacuum-pressure technique. Changes in the samples and fire characteristics of modified wood were studied via thermal analysis (TA), infrared spectroscopy (FTIR), and scanning electron microscopy, coupled with energy-dispersive X-ray spectroscopy (SEM-EDX). The results of TA showed the most significant wood decomposition at a temperature of 350 °C, with a non-significant influence of the nanoparticles. A dominant effect of sodium silicate was observed in the main weight-loss step, resulting in a drop in decomposition temperature within the temperature range of 36–44 °C. More intensive decomposition of wood treated with water glass and nanoparticles led to a faster release of non-combustible gases, which slowed down the combustion process. The results demonstrated that wood modifications using sodium silicate and nanoparticle systems have potentially enhanced flame retardant properties.

1. Introduction

Wood is a natural material widely used for the construction of buildings and for the production of various building elements, furniture, and goods. It is also ecological, comfortable, and aesthetically pleasing; however, its use in building construction is limited and strictly regulated by fire and environmental safety rules because it is highly flammable. Therefore, to obtain fire-safe wood structures, fireproof materials need to be used. The fire protection agents applied to wooden structures contain ammonium phosphates or sulfates, chlorides, oxides, borates and other metal salts, boric acid, and halogen-containing flame retardants [1,2]. Many of these chemicals and their combustion products are highly toxic; therefore, new flame retardants are being sought out. Promising retarders from both environmental and economic points of view include nanomaterials [1,3].
Various nanomaterials with different application methods have been used to improve the fire resistance of wood. In the past, aluminium, boron, and halogens (e.g., bromine) and, more recently, phosphorus and nitrogen have been shown to be effective fire retardants in wood. Upon combustion, the halogenated compounds release toxic and/or highly corrosive gases, which are harmful to both humans and the environment. Halogenated flame retardants are therefore being phased out and replaced with halogen-free alternatives [4,5,6,7,8].
Nanocomposites constitute a new development in fire retardancy [9]. With the advent of nanotechnology in the past few decades, the prospects of nano scale fillers in polymer-based composites within flame retardancy applications have progressed rapidly. Although nanofillers do not inherently show excellent fire retardance, the incorporation of a low amount in polymer composites tends to provide drastic improvements in the thermal stability, smoke release amount, peak heat release rate, and speed at which flames spread throughout the nanocomposites. However, their efficiency is still insufficient in providing adequate fire retardancy when used alone in coatings. Therefore, their combination with other conventional fire retardant systems can give superior properties to the substrate. Nanocomposite-based coatings have been reported as a promising system that protects the oxidation of a char structure, once formed, and thus reinforces the fire-resistant properties of coatings [5,10,11,12]. Nanoparticles can form a coating on the surface of a material to inhibit the release of combustible gas and smoke, isolate oxygen from outside, and to prevent heat transfer. The combustible gas concentration can be diluted by the non-combustible gas produced by pyrolysis of the nanoparticles; the combustion chain reaction is suppressed by highly reactive free radicals produced by pyrolysis of the nanoparticles [13,14].
Zinc oxide nanoparticles have attracted the attention of researchers due to properties such as their biocompatibility, good chemical stability, and high absorption performance; in addition, they can serve as a heat-protective barrier for pine wood samples treated with the water suspension of ZnO [15]. Similarly, nano ZnO particles in the dispersion of potassium methyl siliconate shows a good level of fire retardancy comparable with a commercial flame retardant [16]. Rao et al. [17] found that a small amount of zinc oxide nanoparticles significantly increased the limited oxygen index values of the intumescent flame-retardant coating of plywood. The addition of ZnO nanoparticles could change the thermal degradation behaviours of coatings with increasing char residue percentages at high temperatures [17]. The impregnation of wood with Al2O3 and SiO2 leads to an increase in the char formation and lower thermal conductivity of the surface, as shown in the results from thermal analyses [18].
When comparing the effects of an aqueous dispersion of SiO2, TiO2, and ZrO2, SiO2 dispersion was the most effective in improving the fire properties of pine veneers [19]. In a study by Li et al. [20], titanium dioxide in association with a conventional intumescent flame retardant system that contains ammonium polyphosphate/pentaerythritol/melamine (APP–PER–MEL) was introduced to silicone-acrylate coatings. From the results obtained, its significant effect on char formation and the reduction in the spread of flames on a plywood plate were evident.
Garskaite et al. [21] reported the reinforcement of sapwood of Scots pine using aqueous formulations of sodium metasilicate and nano-TiO2 via a vacuum-pressure technique. Their results showed that the fixation of nano-TiO2 on the wood surface using an aqueous sodium silicate solution has potential in modeling low-cost and less fire-hazardous materials. Recently, Taghiyari et al. [22] found significant improvements in the fire properties of silver fir wood modified with nano-sepiolite. Erceg et al. [23] proposed two procedures for the synthesis of calcium phosphate composites with TiO2 nanoplates and nanowires with great potential for biomedical application; however, fire retardancy may be another area of their application. The hydrothermal method was used to modify a wood surface due to the deposition of TiO2/ZnO coatings at a relatively low temperature, and, as a result, the treated wood had an improved fire resistance, and the one-pot hydrothermal method was a feasible method used to fabricate non-flammable wood materials [24].
The growth of anatase TiO2 coating on a wood surface through the hydrolysis of tetrabutyl orthotitanate (TBOT) in different conditions, using a controlled hydrothermal method at low temperatures was reported [25]. The TiO2 coating effectively acts as a protective layer to prolong the duration of wood combustion and prevents harmful gases from spreading [26].
In recent years, nanoparticles such as titanium dioxide, silicon dioxide, and zinc oxide in intumescent flame retardant coatings have attracted much interest. Despite the progress made in research, many issues regarding the combination of retardants, the concentrations of substances used, methods of application, etc. are still unresolved. In addition, a comparison of the effect of these nanoparticles on wood flame retardancy under the same conditions is lacking.
The aim of this work was to assess three types of nanoparticles and sodium silicate to improve the fire resistance of oak wood.

2. Materials and Methods

2.1. Wood Treatment

Pedunculate oak (Quercus robur L.) specimens with dimensions of 10 × 40 × 50 mm (tangential (T) × radial (R) × longitudinal (L)) were cut from a stem harvested in central Slovakia. The specimens were conditioned in a climate chamber (20 ± 2 °C, 65 ± 3% relative humidity (RH)) for 21 days. The specimens were weighed before and after the conditioning, and the moisture content (MC) of these samples was estimated to be approximately 12% based on dry weight. The samples were divided into eight groups (untreated control; treated with 20% aqueous solution of sodium silicate (water glass, WG); treated with 3% dispersion of nanoparticles—TiO2, SiO2, and ZnO in water; and treated with 3% dispersion of nanoparticles—TiO2, SiO2, and ZnO in 20% aqueous solution of water glass (WG)). The nanoparticles were provided by Merck (Darmstadt, Germany): TiO2, purity ≥ 99.5%, size 21 nm; SiO2, purity ≥ 99.5%, size 5–15 nm; and ZnO, purity ≥ 97.0%, size < 100 nm.
Solutions of nanoparticles at a concentration of 3% were prepared for the impregnation procedure. For this purpose, the required weight of nanoparticles was dispersed in distilled water/20% solution of water glass using an ultrasonic dispenser. The wood samples were treated using the vacuum-pressure process; first, the specimens were kept in a vacuum (–5 kPa) for 2 h and then for 1 h at the pressure of 800 kPa. Afterwards, the treated samples were dried at room temperature to constant weight and then conditioned in a climate chamber (20 ± 2 °C, 65 ± 3% RH) for 7 days. The reference and modified samples were then mechanically disintegrated and milled to particle sizes of 200–300 μm using a POLYMIX PX-MFC 90D laboratory mill (Kinematica, Malters, Switzerland) and dried (4 h at 103 ± 2 °C).

2.2. Samples Analyses

2.2.1. Thermal Analysis

Thermal analysis (TG—thermogravimetry, DTG—differential thermogravimetry, and DSC—differential scanning calorimetry) was performed on the powder wood samples using a STA F3 Jupiter thermal analyzer (Netzsch, Selb, Germany) in the temperature interval from 25 to 600 °C. The measurements were carried out under nitrogen atmosphere for wood samples with weights of 12.5 ± 0.1 mg in corundum (Al2O3) crucibles. The heating rate applied was 10 °C/min for all measurements.

2.2.2. Infrared Spectroscopy

ATR-FTIR spectra of the homogeneously mixed wood powders of the untreated and modified samples were recorded using the Nicolet iS10 FT-IR spectrometer (Thermo Fisher Scientific, Waltham, MA, USA), equipped with Smart iTR using an attenuated total reflectance (ATR) sampling accessory attached to a diamond crystal. The spectra were collected in 32 scans at 4 cm−1 resolution over the range of 4000 to 650 cm−1. The data obtained were analyzed using OMNIC 9.0 software. Four replicates per sample were performed.

2.2.3. Scanning Electron Microscopy–X-ray Spectroscopy Observations

The wood sections were mounted on specimen stubs, sputter-coated with gold (layer thickness of 150 nm) in the Sputter Coater K650X (Quorum Technologies, Ashford, UK) in an argon atmosphere, and examined using high-vacuum scanning electron microscopy coupled with energy-dispersive X-ray spectroscopy (SEM-EDX). SEM-EDX observations of the earlywood vessels were performed using a JEOL JSM-6390LV instrument (JEOL, Tokyo, Japan), operating at 20 kV and a working distance of 15 mm, equipped with an EDX spectroscope INCAx-act (Oxford Instruments, Abingdon, UK). The elemental composition of the nanoparticles and water glass was assessed on the radial and tangential cell wall surfaces of three specimens per treatment, with 3–6 measurements per specimen.

3. Results and Discussion

3.1. Thermal Analysis

Thermal analysis proved that the weight loss (5%, TG curve, not shown) of unmodified oak wood starts at temperatures below 100 °C (max. at 54 °C, DTG curve) due to the removal of absorbed water. The second weight-loss step was detected at a temperature of 290 °C (shoulder on TG curve). The most significant wood decomposition was observed at 350 °C (Figure 1).
According to Sebio-Puñal et al. [27], the peak at 350 °C can be easily assigned to cellulose and the shoulder at 290 °C can be assigned to holocellulose. The weight loss with a maximum decrease at 290 °C is in good agreement with the non-glucosic saccharide content in oak wood [28]; therefore, this peak can be attributed primarily to hemicellulose decomposition. The decompositions of hemicellulose, cellulose, and lignin take place in a relatively narrow range of temperatures, partially overlapping. The complex structure of lignin leads to degradation in a wide temperature range, which overlaps those of hemicellulose and cellulose; specifically, oak lignin degrades over a broad temperature scale [27,29,30]. Whole wood starts to thermally degrade at about 250 °C, caused by various reactions (dehydration, decarboxylation, decarbonylation, depolymerization, etc.) to produce carbon dioxide, carbon monoxide, water acetaldehyde, propenal, methanol, acetic acid, other volatile compounds, and tar and char residues [21,31]. To improve the flammability properties of oak wood, the environmentally friendly modifier sodium silicate (water glass) in combination with various nanoparticles were used. Figure 2 presents the comparison of the main decomposition peaks of the samples treated with 3% dispersion of nanoparticles in a 20% aqueous solution of WG.
Figure 1 and Figure 2 show that nanoparticles have only negligible effects, and the changes in thermal behaviour were influenced mainly by WG (Table 1). A similar observation was reported by Garskaite et al. [21] in Scots pine wood treated with sodium silicate and TiO2 nanoparticles. The dominant effect of WG was observed in the main weight-loss step. The drop in the temperature decomposition was approx. 36–44 °C compared with samples treated without WG, with a non-significant impact of the nanoparticles (Figure 1 and Figure 2, Table 1). Furthermore, the residues at a temperature of 600 °C were higher by approx. 8–12% for the samples treated with WG solution + nanoparticles compared with those treated with water + nanoparticles (Table 1). Alkali solutions strongly change the structures of wood components and accelerate their decomposition at elevated temperatures [32,33]. Thermogravimetry (TG) coupled with the Fourier transform infrared spectroscopy (FTIR) showed that the presence of WG increases the ratio of CO2 band to the carboxylic (C=O) band. The reduced intensities of the carboxylic groups in the treated wood spectra indicated possible extraction of the aliphatic compound during the alkali treatment. Some chemical reactions between TiO2 and the amorphous Na-O-Si gel upon heating were observed [21]. More intensive decomposition of wood treated with WG, and nanoparticles leads to a faster release of non-combustible gases, which slows down the combustion process. From this point of view, ZnO has the greatest influence among the nanoparticles investigated. This effect was also indicated by a reduction in enthalpy (Table 2). The impact of sodium silicate and various compounds for flame retardancy was investigated to find low-cost and less toxic flame retardants. Their retarding effect was evaluated [21,34,35,36] also by measuring the limiting oxygen index (LOI), and a good correlation between thermal analyses and LOI values was found.

3.2. Infrared Spectroscopy

The FTIR spectra were measured on the wood surface before (denoted as “Oak wood”) and after the application of an aqueous solution of sodium silicate (WG) mixed with three types of nanoparticles (TiO2, SiO2, and ZnO, denoted as “Oak wood + WG + nanoparticles”). Given that the changes in the spectra were almost the same for each of the mixtures used, we decided to evaluate them together. As can be seen (Figure 3, Figure 4 and Figure 5), band changes in the range of 3100 to 3600 cm−1 are negligible. This wide band is assigned to O−H vibrations in cellulose, hemicellulose, and lignin structures [37]. An exception is the surface treated with an aqueous solution of ZnO + WG. The growth in the band with a peak at 3380 cm−1 can, in this case, be influenced by the presence of a wide band characteristic of ZnO, as well as the possible presence of water molecules absorbed on the surface of ZnO nanocrystalline powder [21,38]. The interval between 2950 and 2850 cm−1 (symmetric and asymmetric C−H stretching vibrations in aliphatic compounds) [39] shows similar behaviors. In the area from the 1800 to 800 cm−1 bands assigned to stretching and deformation vibrations of all wood components (fingerprint region), more significant changes in absorbances were recorded.
A decrease in band absorbance on 1735 cm−1 (C=O stretching in unconjugated carbonyl groups) of more than 35% (Figure 3, Figure 4 and Figure 5) was observed on the treated samples, indicating changes in several functional groups in lignin and hemicelluloses (carbonyls, aldehydes, ketones, and carboxylic acids) [40,41]. This decrease may be due to changes in the polysaccharides due to the action of an alkaline Na2SiO3 solution. Alkaline treatment of wood promotes the deacetylation of hemicelluloses and has an effect on the degradation of xylans [42]. A separation of xylans and glucomannans is also found in wood [21]. The band around 1600 cm−1 (C=C stretching conjugated with an aromatic ring in lignin) is practically unchanged. An exception is a sample treated with a solution containing ZnO nanoparticles, where an increase in absorbance of about 25% was observed. This may be due to overlap with the relatively wide band present in the ZnO spectrum (Zn−O stretching vibrations in crystalic structures) [43,44]. The band absorbance near 1500 cm−1 (C=C stretching conjugated with aromatic ring in lignin) decreased by around 20%. This mainly indicates a decrease in the number of methoxyl groups, confirming the decrease in lignin content [45,46].
The bands also decreased at 1460 cm−1 (asymmetric CH3 bending in methoxyl groups in lignin), 1370 cm−1 (symmetric and asymmetric CH3 bending), 1320 cm−1 (C−O vibration in syringyl derivatives), and 1235 cm−1 (C−O stretching vibration in xylan and syringyl ring), which are associated with lignin and hemicelluloses [47,48,49,50]. Their decrease supports the assumption that lignin degradation is caused by the presence of an alkaline environment (the pH of the water glass solution used in the experiment was more than 10). In contrast to previous trends, a slight increase in the 1030 cm−1 band (C−O deformation vibrations in cellulose) was observed (Figure 3, Figure 4, and Figure 5). Since this band does not occur in the spectra of the nanoparticles used, the increase is probably supported by a partial overlap and superposition with the band characteristic of water glass (Si−O stretching vibrations) [51]. The band at 897 cm−1 is associated with C−O−C stretching vibrations at glycosidic linkage in cellulose [52]. Absorbance on this band shows a permanent decrease, which confirms the degradation of cellulose.

3.3. Scanning Electron Microscopy–X-ray Spectroscopy Observations

In untreated control wood samples, the vessel walls were found to be free of any deposits and cell wall debris. The vessel-ray pits showed both large and small apertures (Figure 6A). On the contrary, massive layers were observed on vessel wall surfaces when wood was treated with the WG solution. Non-continuous, deeply cracked deposits containing sodium silicate partially covered the cell wall surfaces and pit apertures (Figure 6B).
When TiO2 nanoparticles were applied alone, they formed very thin, non-continuous layers of apparently small particles. The particles penetrated into the pits and were deposited on the pit borders and inside the pit apertures (Figure 7A). The addition of WG in combination with TiO2 nanoparticles caused the formation of extremely large aggregates attached to the vessel wall surface (Figure 7B).
SiO2 nanoparticles did not form a continuous layer but rather clumps consisting of relatively larger particles. Smaller particles again penetrated into the pit apertures, but due to the large size of some clumps, some pits were fully covered with SiO2 nanoparticles (Figure 8A). When the WG solution was applied in combination with SiO2 nanoparticles, mostly roundish aggregate deposits were observed (Figure 8B). ZnO nanoparticles were deposited on the vessel wall surface as a relatively thin continuous layer. Except for some small cracks, a smooth layer covered the cell wall surface and pits (Figure 9A). On the contrary, the addition of WG in combination with ZnO nanoparticles caused the formation of very large aggregates consisting of small nanoparticles covering the surface of vessel-ray pits (Figure 9B). The elemental composition of both the examined nanoparticles and the WG solution was confirmed by the EDX analysis, as shown in the images at the bottom of Figure 6, Figure 7, Figure 8 and Figure 9.

4. Conclusions

Three types of nanoparticles (TiO2, SiO2, and ZnO) and sodium silicate were investigated to improve the fire resistance of oak wood. Thermal analyses show the most considerable wood decomposition at a temperature of 350 °C with a non-significant influence of the nanoparticles. On the other hand, the presence of sodium silicate resulted in a rapid drop in the decomposition temperature. Faster decomposition of wood treated with sodium silicate and nanoparticles led to a faster release of non-combustible gases, which slowed down the combustion process. From this point of view, ZnO has the greatest influence among the nanoparticles investigated. This effect was also indicated by a reduction in enthalpy. The results demonstrated that wood modifications using sodium silicate and nanoparticle systems have potentially enhanced flame retardant properties.

Author Contributions

Conceptualization and methodology, and editing D.K.; FTIR analyses and evaluation I.K., thermal analyses, and evaluation A.E., scanning electron microscopy–X-ray spectroscopy J.Ď. and J.K., data analysis E.K., writing—original draft preparation and supervision F.K. All authors have read and agreed to the published version of the manuscript.

Funding

This work was supported by the Slovak Research and Development Agency under contracts APVV-17-0005 (25%) and APVV-16-0326 (25%), and by the VEGA Agency of Ministry of Education, Science, Research, and Sport of the Slovak Republic 1/0387/18 (25%) and 1/0450/19 (25%).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

The data are available within the article.

Conflicts of Interest

The authors declare no conflict of interest. The funders had no role in the design of the study; in the collection, analyses, or interpretation of data; in the writing of the manuscript; or in the decision to publish the results.

References

  1. Seo, H.J.; Hwang, W.; Lee, M.C. Fire properties of Pinus densiflora utilizing fire-retardant chemicals based on borated and phosphorus (I)—Combustion characteristics. Bioresources 2017, 12, 5417–5427. [Google Scholar] [CrossRef] [Green Version]
  2. Vakhitova, L.N. Fire retardant nanocoating for wood protection. In Nanotechnology in Eco-Efficient Construction; Woodhead Publishing: Cambridge, UK, 2019. [Google Scholar]
  3. Papadopoulos, A.N.; Bikiaris, D.N.; Mitropoulos, A.C.; Kyzas, G.Z. Nanomaterials and chemical modifications for enhanced key wood properties: A review. Nanomaterials 2019, 9, 607. [Google Scholar] [CrossRef] [Green Version]
  4. Luyt, A.S.; Malik, S.S.; Gasmi, S.A.; Porfyris, A.; Andronopoulou, A.; Korres, D.; Vouyiouka, S.; Grosshauser, M.; Pfaendner, R.; Brüll, R.; et al. Halogen-Free Flame-Retardant Compounds. Thermal Decomposition and Flammability Behavior for Alternative Polyethylene Grades. Polymers 2019, 11, 1479. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  5. Popescu, C.M.; Pfriem, A. Treatments and modification to improve the reaction to fire of wood and wood based products—An overview. Fire Mater. 2020, 44, 100–111. [Google Scholar] [CrossRef] [Green Version]
  6. Sauerbier, P.; Mayer, A.K.; Emmerich, L.; Militz, H. Fire retardant treatment of wood—State of the art and future perspectives. In Wood & Fire Safety; Makovicka Osvaldova, L., Markert, F., Zelinka, S., Eds.; Springer: Cham, Switzerland, 2020. [Google Scholar] [CrossRef]
  7. Xia, W.; Fan, S.; Xu, T. Inhibitory action of halogen-free fire retardants on combustion and volatile emission of bituminous components. Sci. Prog. 2021, 104, 00368504211035215. [Google Scholar] [CrossRef]
  8. Xia, W.; Wang, S.; Xu, T.; Jin, G. Flame retarding and smoke suppressing mechanisms of nano composite flame retardants on bitumen and bituminous mixture. Constr. Build. Mater. 2021, 266, 121203. [Google Scholar] [CrossRef]
  9. Giudice, C.A.; Pereyra, A.M. Silica nanoparticles in high silica/alkali molar ratio solutions as fire-retardant impregnants for woods. Fire Mater. 2010, 34, 177–187. [Google Scholar] [CrossRef]
  10. Wang, Z.; Han, E.; Liu, F.; Ke, W. Thermal behavior of nano-TiO2 in fire-resistant coating. J. Mater. Sci. Technol. 2007, 23, 547–550. [Google Scholar]
  11. Spear, M.J.; Curling, S.F.; Dimitriou, A.; Ormondroyd, G.A. Review of Functional Treatments for Modified Wood. Coatings 2021, 11, 327. [Google Scholar] [CrossRef]
  12. Vahidi, G.; Bajwa, D.S.; Shojaeiarani, J.; Stark, N.; Darabi, A. Advancements in traditional and nanosized flame retardants for polymers—A review. J. Appl. Polym. Sci. 2021, 138, 50050. [Google Scholar] [CrossRef]
  13. Fu, B.; Li, X.; Yuan, G.; Chen, W.; Pan, Y. Preparation and flame retardant and smoke suppression properties of bamboo-wood hybrid scrimber filled with calcium and magnesium nanoparticles. J. Nanomater. 2014, 2014, 3. [Google Scholar] [CrossRef]
  14. Zhao, X.; Babu, H.; Llorca, J.; Wang, D. Impact of halogen-free flame retardant with varied phosphorus chemical surrounding on the properties of diglycidyl ether of bisphenol-A type epoxy resin: Synthesis, fire behaviour, flame-retardant mechanism and mechanical properties. RSC Adv. 2016, 6, 59226–59236. [Google Scholar] [CrossRef] [Green Version]
  15. Favarim, H.R.; Leite, L.O. Performance of ZnO nanoparticles for fire retardant and UV protection of pine wood. BioResources 2018, 13, 6963–6969. [Google Scholar] [CrossRef]
  16. Samanta, A.K.; Bhattacharyya, R.; Jose, S.; Basu, G.; Chowdhury, R. Fire retardant finish of jute fabric with nano zinc oxide. Cellulose 2017, 24, 1143–1157. [Google Scholar] [CrossRef]
  17. Nageswara Rao, T.; Naidu, T.M.; Kim, M.S.; Parvatamma, B.; Prashanthi, Y.; Heun Koo, B. Influence of zinc oxide nanoparticles and char forming agent polymer on flame retardancy of intumescent flame retardant coatings. Nanomaterials 2020, 10, 42. [Google Scholar] [CrossRef] [Green Version]
  18. Deveci, I.; Sacli, C.; Turkoglu, T.; Baysal, E.; Toker, H.; Peker, H. Effect of SiO2 and Al2O3 nanoparticles treatment on thermal behavior of oriental beech wood. Wood Res. Slovak. 2018, 63, 573–581. [Google Scholar]
  19. Bueno, A.B.F.; Banonn, M.V.N.; Morentin, L.; Garcia, M.J.M. Treatment of natural woodveneers with nanooxides to improve their fire behaviour. In Proceedings of the 2nd International Conference on Structural Nano Composites, Madrid, Spain, 20–21 May 2014. [Google Scholar] [CrossRef] [Green Version]
  20. Li, H.F.; Hu, Z.W.; Zhang, S.; Gu, X.Y.; Wang, H.J.; Jiang, P.; Zhao, Q. Effects of titanium dioxide on the flammability and char formation of water-based coatings containing intumescent flame retardants. Prog. Org. Coat. 2015, 78, 318–324. [Google Scholar] [CrossRef]
  21. Garskaite, E.; Karlsson, O.; Stankeviciute, Z.; Kareiva, A.; Jones, D.; Sandberg, D. Surface hardness and flammability of Na2SiO3 and nano-TiO2 reinforced wood composites. RSC Adv. 2019, 9, 27973–27986. [Google Scholar] [CrossRef] [Green Version]
  22. Taghiyari, H.R.; Tajvidi, Μ.; Soltani, A.; Esmailpour, A.; Khodadoosti, G.; Jafarzadeh, H.; Militz, H.; Papadopoulos, A.N. Improving fire retardancy of unheated and heat-treated fir wood by nano-sepiolite. Eur. J. Wood Prod. 2021, 79, 841–849. [Google Scholar] [CrossRef]
  23. Erceg, I.; Selmani, A.; Gajović, A.; Radatović, B.; Šegota, S.; Ćurlin, M.; Strasser, V.; Kontrec, J.; Kralj, D.; Maltar-Strmečki, N.; et al. Precipitation at Room Temperature as a Fast and Versatile Method for Calcium Phosphate/TiO2 Nanocomposites Synthesis. Nanomaterials 2021, 11, 1523. [Google Scholar] [CrossRef]
  24. Sun, Q.F.; Lu, Y.; Xia, Y.Z.; Yang, D.J.; Li, J.; Liu, Y.X. Flame retardancy of wood treated by TiO2/ZnO coating. Surf. Eng. 2012, 28, 555–559. [Google Scholar] [CrossRef] [Green Version]
  25. Li, J.; Yu, H.; Sun, Q.; Liu, Y.; Cui, Y.; Lu, Y. Growth of TiO2 coating on wood surface using controlled hydrothermal method at low temperatures. Appl. Surf. Sci. 2010, 256, 5046–5050. [Google Scholar] [CrossRef]
  26. Sun, Q.; Yu, H.; Liu, Y.; Li, J.; Cui, Y.; Lu, Y. Prolonging the combustion duration of wood by TiO2 coating synthesized using cosolvent-controlled hydrothermal method. J. Mater. Sci. 2010, 45, 6661–6667. [Google Scholar] [CrossRef]
  27. Sebio-Puñal, T.; Naya, S.; López-Beceiro, J.; Tarrío-Saavedra, J.; Artiaga, R. Thermogravimetric analysis of wood, holocellulose, and lignin from five wood species. J. Therm. Anal. Calorim. 2012, 109, 1163–1167. [Google Scholar] [CrossRef]
  28. Čabalová, I.; Kačík, F.; Lagaňa, R.; Výbohová, E.; Bubeníková, T.; Čaňová, I.; Ďurkovič, J. Effect of thermal treatment on the chemical, physical, and mechanical properties of pedunculate oak (Quercus robur L.) wood. BioResources 2018, 13, 157–170. [Google Scholar] [CrossRef]
  29. González-Díaz, E.; Alonso-López, J.M. Characterization by thermogravimetric analysis of the wood used in Canary architectural heritage. J. Cult. Herit. 2017, 23, 111–118. [Google Scholar] [CrossRef]
  30. Lühr, C.; Pecenka, R. Development of a model for the fast analysis of polymer mixtures based on cellulose, hemicellulose (xylan), lignin using thermogravimetric analysis and application of the model to poplar wood. Fuel 2020, 277, 118169. [Google Scholar] [CrossRef]
  31. Rowell, R.M.; LeVan-Green, S.L. Thermal properties. In Handbook of Wood Chemistry and Wood Composites; CRC Press: Boca Raton, FL, USA, 2005; pp. 121–138. [Google Scholar]
  32. Xu, E.; Wang, D.; Lin, L. Chemical structure and mechanical properties of wood cell walls treated with acid and alkali solution. Forests 2020, 11, 87. [Google Scholar] [CrossRef] [Green Version]
  33. Sundberg, K.E.; Holmbom, B.R.; Pranovich, A.V. Chemical changes in thermomechanical pulp at alkaline conditions. J. Wood Chem. Technol. 2003, 23, 89–112. [Google Scholar] [CrossRef]
  34. Shabir Mahr, M.; Hübert, T.; Schartel, B.; Bahr, H.; Sabel, M.; Militz, H. Fire retardancy effects in single and double layered sol-gel derived TiO2 and SiO2-wood composites. J. Sol. Gel. Sci. Technol. 2012, 64, 452–464. [Google Scholar] [CrossRef]
  35. Li, P.; Zhang, Y.; Zuo, Y.; Wu, Y.; Yuan, G.; Lu, J. Comparison of silicate impregnation methods to reinforce Chinese fir wood. Holzforschung 2021, 75, 126–137. [Google Scholar] [CrossRef]
  36. Sun, Z.; Lv, J.; Wang, Z.; Wu, Y.; Yuan, G.; Zou, Y. Sodium silicate/waterborne epoxy resin hybrid-modified Chinese fir wood. Wood Sci. Technol. 2021, 55, 837–855. [Google Scholar] [CrossRef]
  37. Kubovský, I.; Kačíková, D.; Kačík, F. Structural Changes of Oak Wood Main Components Caused by Thermal Modification. Polymers 2020, 12, 485. [Google Scholar] [CrossRef] [Green Version]
  38. Bobrowski, A.; Stypuła, B.; Hutera, B.; Kmita, A.; Drożyński, D.; Starowicz, M. FTIR spectroscopy of water glass—The binder moulding modified by ZnO nanoparticles. Metalurgija 2012, 51, 477–480. [Google Scholar]
  39. Vartanian, E.; Barres, O.; Roque, C. FTIR spectroscopy of woods: A new approach to study the weathering of the carving face of a sculpture. Spectrochim. Acta A 2015, 136, 1255–1259. [Google Scholar] [CrossRef]
  40. Colom, X.; Carrillo, F.; Nogués, F.; Garriga, P. Structural analysis of photodegraded wood by means of FTIR spectroscopy. Polym. Degrad. Stabil. 2003, 80, 543–549. [Google Scholar] [CrossRef]
  41. Popescu, M.C.; Popescu, C.M.; Lisa, G.; Sakata, Y. Evaluation of morphological and chemical aspects of different wood species by spectroscopy and thermal methods. J. Mol. Struct. 2011, 988, 65–72. [Google Scholar] [CrossRef]
  42. Dibdiaková, J.; Geffertová, J.; Rázgová, Z. Alkali and alkali/oxidation treatment of poplar wood (Populus Nigra)—Influence on the kraft pulp properties. Acta Facultatis Xylologiae Zvolen 2010, 52, 53–62. [Google Scholar]
  43. Jayarambabu, N.; Kumari, B.S.; Rao, K.V.; Prabhu, Y.T. Germination and growth characteristics of mungbean seeds (Vigna radiata L.) affected by synthesized zinc oxide nanoparticles. Int. J. Curr. Eng. Technol. 2014, 4, 2347–5161. [Google Scholar]
  44. Etcheverry, L.P.; Flores, W.H.; Silva, D.L.; Moreira, E.C. Annealing effects on the structural and optical properties of ZnO nanostructures. Mater. Res. 2018, 21, 1–7. [Google Scholar] [CrossRef]
  45. Guo, J.; Song, K.; Salmén, L.; Yin, Y. Changes of wood cell walls in response to hygro-mechanical steam treatment. Carbohydr. Polym. 2015, 115, 207–214. [Google Scholar] [CrossRef] [PubMed]
  46. Lourenço, A.; Pereira, H. Compositional Variability of Lignin in Biomass. Intech Open 2018, 2, 64–98. [Google Scholar] [CrossRef] [Green Version]
  47. Faix, O.; Lin, S.Y.; Dence, C.W. Fourier transform infrared spectroscopy. In Methods in Lignin Chemistry; Lin, S.Y., Dence, C.W., Eds.; Springer: Berlin, Germany, 1992; pp. 83–109. [Google Scholar]
  48. Müller, G.; Schöpper, C.; Vos, H.; Kharazipour, A.; Polle, A. FTIR-ATR spectroscopic analysis of changes in wood properties during particle and fibreboard production of hard and softwood trees. BioResources 2009, 4, 49–71. [Google Scholar] [CrossRef]
  49. Lionetto, F.; Del Sole, R.; Cannoletta, D.; Vasapollo, G.; Maffezzoli, A. Monitoring wood degradation during weathering by cellulose crystallinity. Materials 2012, 5, 1910–1922. [Google Scholar] [CrossRef] [Green Version]
  50. Kubovský, I.; Oberhofnerová, E.; Kačík, F.; Pánek, M. Surface changes of selected hardwoods due to weather conditions. Forests 2018, 9, 557. [Google Scholar] [CrossRef] [Green Version]
  51. Pfeffer, A.; Mai, C.; Militz, H. Weathering characteristics of wood treated with water glass, siloxane or DMDHEU. Eur. J. Wood Prod. 2012, 70, 165–176. [Google Scholar] [CrossRef] [Green Version]
  52. Ciolacu, D.; Ciolacu, F.; Popa, V. Amorphous cellulose-structure and characterization. Cell. Chem. Technol. 2011, 45, 13–21. [Google Scholar]
Figure 1. Differential thermal gravimetry (DTG) curves of the samples treated with 3% dispersion of nanoparticles (1—Oak wood, green; 2—SiO2, red; 3—TiO2, blue; 4—ZnO, black).
Figure 1. Differential thermal gravimetry (DTG) curves of the samples treated with 3% dispersion of nanoparticles (1—Oak wood, green; 2—SiO2, red; 3—TiO2, blue; 4—ZnO, black).
Nanomaterials 11 03405 g001
Figure 2. Comparison of DTG curves of the samples treated with 3% dispersion of nanoparticles in 20% aqueous solution of water glass (WG) (5—SiO2 + WG, red; 6—TiO2 + WG, blue; 7—WG, green; 8—ZnO + WG, black).
Figure 2. Comparison of DTG curves of the samples treated with 3% dispersion of nanoparticles in 20% aqueous solution of water glass (WG) (5—SiO2 + WG, red; 6—TiO2 + WG, blue; 7—WG, green; 8—ZnO + WG, black).
Nanomaterials 11 03405 g002
Figure 3. FTIR spectrum of oak wood coated with water glass and TiO2 nanoparticles.
Figure 3. FTIR spectrum of oak wood coated with water glass and TiO2 nanoparticles.
Nanomaterials 11 03405 g003
Figure 4. FTIR spectrum of oak wood coated with water glass and SiO2 nanoparticles.
Figure 4. FTIR spectrum of oak wood coated with water glass and SiO2 nanoparticles.
Nanomaterials 11 03405 g004
Figure 5. FTIR spectrum of oak wood coated with water glass and ZnO nanoparticles.
Figure 5. FTIR spectrum of oak wood coated with water glass and ZnO nanoparticles.
Nanomaterials 11 03405 g005
Figure 6. Scanning electron microscopy images of earlywood vessel wall surfaces (upper images) accompanied with the representative EDX spectra showing the elemental composition (bottom images). (A) Untreated control oak wood, radial section, scale bar = 50 μm. (B) Treatment with the aqueous solution of water glass, tangential section, scale bar = 50 μm.
Figure 6. Scanning electron microscopy images of earlywood vessel wall surfaces (upper images) accompanied with the representative EDX spectra showing the elemental composition (bottom images). (A) Untreated control oak wood, radial section, scale bar = 50 μm. (B) Treatment with the aqueous solution of water glass, tangential section, scale bar = 50 μm.
Nanomaterials 11 03405 g006
Figure 7. Scanning electron microscopy images of earlywood vessel wall surfaces (upper images) accompanied with the representative EDX spectra showing the elemental composition (bottom images). (A) TiO2 nanoparticles, radial section, scale bar = 5 μm. (B) Aqueous solution of water glass in combination with TiO2 nanoparticles, radial section, scale bar = 10 μm.
Figure 7. Scanning electron microscopy images of earlywood vessel wall surfaces (upper images) accompanied with the representative EDX spectra showing the elemental composition (bottom images). (A) TiO2 nanoparticles, radial section, scale bar = 5 μm. (B) Aqueous solution of water glass in combination with TiO2 nanoparticles, radial section, scale bar = 10 μm.
Nanomaterials 11 03405 g007
Figure 8. Scanning electron microscopy images of earlywood vessel wall surfaces (upper images) accompanied with the representative EDX spectra showing the elemental composition (bottom images). (A) SiO2 nanoparticles, radial section, scale bar = 5 μm. (B) Aqueous solution of water glass in combination with SiO2 nanoparticles, radial section, scale bar = 10 μm.
Figure 8. Scanning electron microscopy images of earlywood vessel wall surfaces (upper images) accompanied with the representative EDX spectra showing the elemental composition (bottom images). (A) SiO2 nanoparticles, radial section, scale bar = 5 μm. (B) Aqueous solution of water glass in combination with SiO2 nanoparticles, radial section, scale bar = 10 μm.
Nanomaterials 11 03405 g008
Figure 9. Scanning electron microscopy images of earlywood vessel wall surfaces (upper images) accompanied with the representative EDX spectra showing the elemental composition (bottom images). (A) ZnO nanoparticles, radial section, scale bar = 2 μm. (B) Aqueous solution of water glass in combination with ZnO nanoparticles, radial section, scale bar = 5 μm.
Figure 9. Scanning electron microscopy images of earlywood vessel wall surfaces (upper images) accompanied with the representative EDX spectra showing the elemental composition (bottom images). (A) ZnO nanoparticles, radial section, scale bar = 2 μm. (B) Aqueous solution of water glass in combination with ZnO nanoparticles, radial section, scale bar = 5 μm.
Nanomaterials 11 03405 g009
Table 1. Thermal degradation temperatures and residue at 600 °C from TG and DTG analyses.
Table 1. Thermal degradation temperatures and residue at 600 °C from TG and DTG analyses.
SampleT1 (°C)T2 (°C)Residue at 600 °C (%)
Oak wood290.0350.418.3
Oak wood + WG306.525.8
Oak wood + TiO2294.0348.517.5
Oak wood + WG + TiO2262.1310.726.0
Oak wood + SiO2294.0354.516.2
Oak wood + WG + SiO2318.225.3
Oak wood + ZnO293.0348.617.2
Oak wood + WG + ZnO219.0305.328.4
Table 2. Peak temperatures and enthalpy obtained from DSC curves.
Table 2. Peak temperatures and enthalpy obtained from DSC curves.
SampleT1 (°C)ΔH (J/g)T2 (°C)ΔH (J/g)
Oak wood352.560.1440.357.9
Oak wood + WG406.913.7
Oak wood + TiO2351.559.6450.878.1
Oak wood + WG + TiO2364.07.8
Oak wood + SiO2358.453.1409.459.8
Oak wood + WG + SiO2374.77.1442.610.3
Oak wood + ZnO351.169.2434.552.7
Oak wood + WG + ZnO357.6–1.3
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Kačíková, D.; Kubovský, I.; Eštoková, A.; Kačík, F.; Kmeťová, E.; Kováč, J.; Ďurkovič, J. The Influence of Nanoparticles on Fire Retardancy of Pedunculate Oak Wood. Nanomaterials 2021, 11, 3405. https://doi.org/10.3390/nano11123405

AMA Style

Kačíková D, Kubovský I, Eštoková A, Kačík F, Kmeťová E, Kováč J, Ďurkovič J. The Influence of Nanoparticles on Fire Retardancy of Pedunculate Oak Wood. Nanomaterials. 2021; 11(12):3405. https://doi.org/10.3390/nano11123405

Chicago/Turabian Style

Kačíková, Danica, Ivan Kubovský, Adriana Eštoková, František Kačík, Elena Kmeťová, Ján Kováč, and Jaroslav Ďurkovič. 2021. "The Influence of Nanoparticles on Fire Retardancy of Pedunculate Oak Wood" Nanomaterials 11, no. 12: 3405. https://doi.org/10.3390/nano11123405

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