Abstract
This review article considers the formulation of a broad range of coatings designed for the protection and changing the appearance of wood surfaces. Findings from the literature are considered from the standpoint of the main chemical components, how they can be formulated into a spreadable product, the events leading to curing, and factors affecting the performance of the resulting coating layers on wood surfaces. A series of hypotheses are considered, relating to the mechanisms underlying wood coating products and their usage. Special attention is paid to the topics of adhesion at the coating-wood interface, the development of film strength and hardness, and challenges related to the past and continuing development of waterborne coating formulations. The modern technologist seeking to coat wood has many options to choose from, and there has been a need to make current knowledge related to the field more available to the wider scientific community.
Download PDF
Full Article
Chemical and Mechanistic Aspects of Wood Finishing: A Review Encompassing Paints and Clear Coats
Martin A. Hubbe ,* and Frederik Laleicke
This review article considers the formulation of a broad range of coatings designed for the protection and changing the appearance of wood surfaces. Findings from the literature are considered from the standpoint of the main chemical components, how they can be formulated into a spreadable product, the events leading to curing, and factors affecting the performance of the resulting coating layers on wood surfaces. A series of hypotheses are considered, relating to the mechanisms underlying wood coating products and their usage. Special attention is paid to the topics of adhesion at the coating-wood interface, the development of film strength and hardness, and challenges related to the past and continuing development of waterborne coating formulations. The modern technologist seeking to coat wood has many options to choose from, and there has been a need to make current knowledge related to the field more available to the wider scientific community.
DOI: 10.15376/biores.20.2.Hubbe
Keywords: Finishing of wood; Solvent-borne coatings; Waterborne coatings; Primers; Varnish; Paint; Wood protection; Alkyds; Latex; Lacquer
Contact information: Department of Forest Biomaterials, College of Natural Resources, North Carolina State University, Campus Box 8005, Raleigh, NC 27605-8005, USA;
* Corresponding author: hubbe@ncsu.edu
INTRODUCTION
The painting of wood can appear to be simple (Flexner 1994). Stir the paint, dip the brush, apply the paint, and wait for it to dry. The goal of this review article is to examine some of the underlying details. Though the application of a wood coating may seem easy to implement, the successful finishing of wood has required a great deal of innovation and optimization, which continues up to the present. This article attempts to answer questions such as, “How do these systems work?” Published literature is reviewed as a way to answer this and related questions.
Motivations to Apply Coatings to Wood
In a broad sense, a primary purpose of a layer of finish on a wood surface is to protect the wood (Williams et al. 1996; Allen 1984; Flexner 2005). This role is highlighted in Fig. 1, along with an additional role of affecting the visible appearance of the surface. Wood is an exemplary sustainable material, but without a protective finish or “coating,” it can suffer from early degradation. Factors responsible for such degradation have been reviewed (Feist 1982; Cogulet et al. 2018; Kropat et al. 2020). Modes of degradation of wood can be categorized as weathering (Grüll et al. 2014; Kropat et al. 2020), decay or rot (Clausen 1996; Goodell 2003; Brischke et al. 2006), insect damage (Stirling and Temiz 2014), surface damage, distortion of the shape (Flexner 1994), and mechanical breakage (Smith et al. 2007).
With the exception of the last listed category, all of these undesired degradation effects can be slowed down by application of a finish, which typically includes some form of polymeric resin, a liquid vehicle to either solubilize the resin or disperse emulsion droplets of the resin, and some other optional materials. These can include ultraviolet (UV) light-absorbing materials and various colorants. The finish is generally applied to wood as a liquid, after which it dries or cures to form either a stain effect or a solid film. The goal of this article is to explore published literature that explains the mechanisms by which wood coatings function.
Fig. 1. Figure highlighting some of the main functions of a coating layer on wood
Some further reasons to apply coatings to wood fall under the broad category of appearance (Marschner et al. 2005). Among the most basic of these is that finished wood can be a lot easier to clean (Flexner 2005). A wood surface lacking a finish layer will be susceptible to dirt, grime, and stains that can be difficult to remove. Categories of appearance to be considered later in this review include color, gloss, and the extent to which the finish has been designed to hide the appearance of what is below it.
Hypotheses
To help guide the discussion in this article, some general hypotheses can be offered, as listed below.
- Understanding wood’s degradation and the factors that affect it can provide a foundation for understanding the role of protective finishes.
- By blocking all or most rainwater from wetting the wood itself, a finish is able to decrease the swings of cyclical changes in moisture content that would have been experienced by unprotected wood exposed to outdoor weather.
- Due to its ability to absorb incident ultraviolet (UV) light, a suitably formulated finish can avoid or slow down the photodegradation of lignin near the surface of the wood, thus helping to maintain the integrity of the wood surface.
- By presenting a suitably hard and contiguous film at the wood surface, the finish can protect the wood from scratching and abrasion.
- By drying and curing in a way that develops smoothness, the finish can contribute to a glossy appearance, and the level of gloss can be adjusted by formulation of the finish mixture.
- Surface preparation steps and many other details are foundational to achieving film and reliable adhesion of the finish to the wood to achieve an expected service life of multiple years or even decades in some cases.
- To perform its various functions, an effective layer of finish will need to have undergone a curing process, which can involve evaporation of one or more carrier fluids (drying), oxidation and other chemical changes, and even covalent crosslinking in some cases.
- To be able to spread effectively onto either wood or a previous layer of finish, each successive layer of coating to be spread needs to be formulated with a suitable surface tension, as well as having resins able to spread and interact with the functional groups facing the surface.
- A final hypothesis to be considered in this article can be attributed to Flexner (1994, 2005). The background for this hypothesis involves the design of water-borne wood finishes. The general expectation of a water-borne finish is that it ought to perform its various roles equally well in comparison to traditional oil-borne formulations. The hypothesis states that a well-performing water-borne finish formulation will contain not only water, but also a minor amount of a slower-to-evaporate vehicle, which in addition needs to be a good solvent for the polymeric binder in the finish. Such a system can be expected to permit intermixing among polymer segments of adjacent macromolecules, thus resulting in a contiguous film.
Formulation principles can account for finish performance
As an approach to addressing the various hypotheses given above, it will be assumed that the ingredients and composition of a coating formulation can account for how well it performs. One of the most consequential changes in formulation, when comparing typical modern finishes with those used a century ago, has been the emergence of waterborne finishes (Flexner 1994, 2005; Cox 2003; Landry and Blanchet 2012; Poussard et al. 2016; Shukla et al. 2019). In comparison to the traditional furnishes, all of which have been formulated with more non-polar fluids, the waterborne formulations of finishes have a reputation of being more challenging to use, especially when facing issues related to the surface to be covered, weather conditions, and some aspects related to performance (Flexner 1994, 2005; Cox 2003). Despite such concerns, the late 1999s began a strong and continuing trend toward waterborne formulation of many finish products, and the main driver has been to reduce the emission of volatile organic solvents (Williams and Feist 1999). Since that time, waterborne formulations have undergone a maturation process, and users have gained experience in using such products.
Published Reviews Related to Wood Finishing
Some topics related to the chemical and mechanistic wood finishing already have been covered in review articles and books. Some notable examples are listed in Table 1.
Table 1. Review Articles and Books Covering Some Chemical or Mechanistic Aspects of Wood Finishing
CHEMICAL ASPECTS OF WOOD FINISH TYPES
A wide variety of chemical ingredients are used in coating formulations for wood, and the commercial products can be categorized in different ways. For instance, one can differentiate between the major classes of oil-based finishes, waterborne finishes, reactive coatings, and powder coatings. Alternatively, one can focus on the intended end-usage, such as interior finish products and exterior finish products. Another option would be to differentiate between clear coatings, stained or wash coatings, vs. opaque finishes. In this section, the main focus will be placed on the chemical composition of the resin, i.e. the ingredient that will be providing the main adhesion within the coating layer.
Types of Finish, with Emphasis on the Binder Component
Because of this article’s focus on chemical and mechanistic aspects, the cited examples are not limited to commercially available products. For example, the history of shellac finish goes back about 5000 years (Ahuja 2024), and shellac continues to be used in antique restoration and in specialized applications (Waring 1963; Williams et al. 1996; Flexner 2005). However, shellac has largely been displaced in the market by products that offer greater water-resistance and other advantages. At the other end of the spectrum, recent university research often focusses on concepts and formulations that lie beyond what is presently commercially available. Though the results of such studies have potential to contribute to a better understanding of how coating formulations and applied coatings behave, it is important to keep in mind that many of the systems described in scientific articles will not ever become available for purchase. Reasons can include high costs of materials, inherent difficulties of displacing mature technologies that already are on the market, and practical disadvantages not yet revealed by the initial research.
Table 2 highlights key attributes of some of the major types of finish uses for a variety of applications. For readers who need even more details, perhaps when paying attention to narrowing down the best selection for a specific application, more extensive tables are provided by Flexner (2005).
Table 2. Finish Formulation Options, with Emphasis on the Binder
General Formulation Principles
This section considers some basic requirements of coating formulations, which can be expected to be valid regardless of their contents and fluid medium. This includes such issues as initial uniformity, continuing uniformity over time, and whether some of the components are dissolved or dispersed as colloidal particles.
Uniformity of the mixture
The human eye is very good at finding imperfections, especially when viewing surfaces that are intended all to have the same color or gloss. Any clumping of the formulation can be expected to show up as tell-tale bumps in the coating, which will be regarded by viewers as evidence of poor workmanship. The first step in achieving an even mixture is likely to involve some form of milling or stirring. For instance, Landry et al. (2008b, 2010) compared high-speed mixing vs. several kinds of milling relative to the even dispersal of a nano-sized clay component. A strong correlation was found between the quality of dispersion of the particles in the formulation and the resulting optical and mechanical properties of the coatings. In follow-up work, the same research team found that sub-optimal coating results were obtained in a case where nano-sized solid particles had formed agglomerates in a UV-curable waterborne formulation (Sow et al. 2011). The findings of that study suggest that some of the problems were due to incompatibility of overly hydrophobic (silane-treated) mineral surfaces when attempting to prepare a waterborne formulation. More promising results were obtained by Vardanyan et al. (2014), who were using cellulose nanocrystals (CNC) as a reinforcing agent in another UV-curing waterborne formulation. The authors noted increases in hardness and scratch resistance, while not seeing any deterioration in appearance, all of which provide evidence that the starting formulation must have had a uniform distribution.
As illustrated in Fig. 2, the mixing of solids particles, especially the resins and any pigments, into the coating formulation typically involves impeller stirring. The hydrodynamic shear forces are expected to help break apart connections between individual particles such as to achieve a uniform dispersion, which can be regarded as a necessary condition in order to achieve uniformity in the coated layer.
Fig. 2. The essential role of impeller stirring in the preparation of typical wood finish formulations
Colloidal stability
Effective mixing, usually brought about by application of hydrodynamic shear, is not the only issue faced by formulators aiming to achieve a uniform suspension, especially in cases where the formulation needs to contain emulsion droplets or solid particles. In order for the formulation to remain stable over the expected length of time (e.g., the listed shelf life), there will need to be something to prevent the particles from coming together and forming agglomerates. In other words, there need to be colloidal-range repulsive forces between the particles or droplets in the mixture.
Colloidal instability problems often can be overcome by adding a small amount of a well-chosen surface-active agent (Baumstark and Tiarks 2002; Godnjavec et al. 2012; Acarali and Demir 2021; Bansal et al. 2022; Pan and Yan 2023). The idea is that one end of the surfactant is attracted to the surface of the entity to be dispersed, and the other is compatible with the suspending medium. In the formulations of wood finishes, such agents are typically nonionic.
Another possibility is to employ particles as emulsion stabilizers, i.e. the formation of Pickering emulsions (Chevalier and Bolzinger 2013). A recent trend has been to study the use of nanocellulose particles for formation of wood finishes. Thus, Grüneberger et al. (2015) found that addition of nanofibrillated cellulose (NFC) aided in the stabilization of UV-absorbing ZnO nanoparticles in a waterborne finish. The approach worked well in formulations containing acrylic resin, but aggregation phenomena could not be overcome in formulations containing alkyd resins (Grüneberger et al. 2014). Mikkonen et al. (2019) found that hemicelluloses from wood could be used as an effective stabilizer for alkyd resins.
Since hemicelluloses are hydrophilic, their adsorption onto the alkyd resins would be expected to provide a steric stabilization effect (Napper 1977) that can be helpful in formulating a waterborne emulsion. The concept is that the hydrophilic material at the surfaces prevents their close approach, thus tending to maintain colloidal stability. Poaty et al. (2014) employed a contrasting approach for preparing a waterborne finish formulation. Their first step was to treat the surface of cellulose nanocrystals with acryloyl chloride or alkyl quaternary ammonium bromides, thus making the surfaces hydrophobic. The acrylic resins, after stabilization by the hydrophobized nanocellulose, showed good stability in the aqueous system, but the main benefit was improved scratch resistance of the resulting finish. The effects were attributed to better compatibility between the coated nanocellulose and the resin. Related work was reported by Veigel et al. (2017). A linseed oil-based coating was improved by adding NFC that had been derivatized with alkyl groups, to render it hydrophobic. Yoo and Youngblood (2017) surface-modified cellulose nanocrystals so that they could be more compatible with a tung oil finish system. The distribution of the CNC was improved, as was the scratch resistance of the coatings. Salleh et al. (2013) noted a positive effect of modifying a silica pigment with a hydrophobic silane additive before its usage in a finish formulation with a radiation curable acrylate resin. Related work was reported by Sow et al. (2011). Zhu et al. (2016) showed that it is possible to reduce the volatile organic carbon (VOC) content of a finish formulation by the usage of nanoparticles acting as a stabilizer of waterborne polyurethane varnish.
Solubility principles
Highly noticeable blemishes in coatings, known as fisheyes, appear to be associated with phase separation phenomena (Allen 1984). A frequent cause of such blemishes is the movement of wood resins into the wet layer of the finish. But because of differences in the solubility behavior between the resins and the fluid medium, the phases remain separate. Similar effects may be seen when highly hydrophobic silicon oils or waxes, possibly coming from a previous layer of finish, fail to be well mixed in the next finish formulation (Cox 2003; Flexner 2005; Huang et al. 2023). Such problems often can be predicted and addressed by the application of solubility principles (Hansen 2007). As illustrated in Fig. 3, the Hansen system is based on determination of three characteristic properties for each component in the formulation, namely the Hildebrand parameter (related to the cohesive energy density), the hydrogen bonding ability, and the polarity. Because the two materials represented in Fig. 3 show an overlap of the solubility regions (represented as spherical regions), one predicts that a uniformity mixture can be achieved. Though it will not be possible to avoid incompatibilities between ingredients of a coating formulation and various contaminants, the solubility calculations can be helpful for identifying additives that can serve as compatibilizers between components having contrasting solubility.
Fig. 3. Representation of the three main parameters considered in the Hansen system to predict phased compatibility of mixtures
Pigments or Filler Particles
This section considers some different ways in which solid particles present within a coating formulation can affect not only the final properties of the coating, but also how it behaves during its application to the wood. They can scatter light, thus contributing to hiding power. In addition, they can provide reinforcement of the film layer, they can affect the permeability of a coating layer, and they can affect the flow characteristics.
To avoid confusion, it is important to note that the word “filler” can mean two different things in the context of coatings for wood. A “paste filler” is a viscous mixture that can be used to fill large pores in wood surfaces, as in the end grain of oak (Allen 1984; Flexner 2005). This section will consider a different optional meaning of the term, referring to solid particles and their effects on appearance and mechanical properties.
Hiding power
Solid particles, especially titanium dioxide, are widely used in finish formulations when the goal is to achieve opacity. Additional reasons to add particles include absorption of UV light (to protect the wood), and to act as a reinforcing agent. Thus, one can envision a pigmented coating for wood as a kind of composite, in which the particles can play a reinforcing role, and the resin may be envisioned as playing the role of the matrix material.
The ability of white particles, added to a paint formulation, to prevent light from passing through the coating depends on the random scattering of the light. Each time the light encounters a localized abrupt change in refractive index, the light energy changes its direction. A contrasting effect happens in cases where the surface is smooth, at the scale of the light waves; in such cases there will be some mirror-like reflection (gloss), which will be discussed later. Effects of light scattering, resulting from particles or bubbles within a dense layer can be estimated by applying equations developed for that purpose by Kubelka and Munk (1931), whose results have been subsequently made more precise (Yang and Kruse 2004). Kubelka and Munk derived relationships that envision the one-directional progress of a single beam, being subject to both scattering (random change in direction) and absorption (gradual decrease in intensity) during its progress. The superior performance of titanium dioxide, and to a lesser extent ZnO and PbO, can be attributed to their much higher index of refraction than a typical resin, making up most of the remaining content of the coating. A higher contrast in refractive index gives rise to a greater light scattering effect, and hence greater hiding power.
Some key aspects associated with light interactions, including the Kubelka-Munk analysis, are illustrated in Fig. 4. Notice that the incident light is envisioned as coming from the upper level. To the extent that the surface is smooth enough, within a length scale associated with the wavelengths of light, some of the light may be reflected in mirror-like fashion. But more likely, the rough nature of a realistic coating layer will mean that much of the light is diffusely reflected as it encounters a change in refractive index as it passes from the air into the coating material. Further scattering of light, leading to more diffuse reflection, can be expected when the light encounters particles, especially if they have contrasting values of refractive index. The transmitted light beam is depicted as becoming gradually thinner as it passes through the coating layer. This is due to an expectation that some of the light energy will become absorbed by chromophores in the coating.
Fig. 4. Some ways in which light can interact with a coating layer
Reinforcement effects
It is well known that the incorporation of solid particles, especially if they have a high aspect ratio and strong interfacial adhesion with the matrix material, will give rise to an increased elastic modulus of the material (Hubbe and Grigsby 2020). Such a system is illustrated schematically in Fig. 5.
Fig. 5. Representation of a coating film that has been reinforced by elongated articles, such as fibers
As shown by Huang et al. (2022), particles present in a wood coating can provide additional benefits, such as scratch resistance. Increased hardness was observed when lignin-containing nanofibrillated cellulose was added to the formulation for a waterborne coating (Huang et al. 2019). However, all of the potential benefits just mentioned can be achieved only if there is molecular contact over at least a large proportion of the interfacial area. Poor compatibility between the matrix and reinforcing particles can be expected to result in strength properties similar to or lower than that of the matrix by itself (Hubbe and Grigsby 2020).
Studies have shown that in addition to increasing the Young’s modulus of a formulated system, the presence of stiff reinforcing particles generally will decrease the elongation ability before breakage (Hubbe and Grigsby 2020). This effect is illustrated in Fig. 6.
Fig. 6. Expected effect of elongated reinforcing particles on the elastic modulus and elongation to breakage of a plastic material
A further, sometimes unrecognized effect of small solid particles present in the formulation of a wood coating may involve the nucleation of interactions occurring at the interface between polymer components of the mixture. Thus, Fallah et al. (2017) attributed an observed increase in the glass transition temperature, Young’s modulus, and hardness of a coating to the repression of polymer segment movements by the added nano-silica particles.
Permeability effects
Though crystalline mineral particles can be regarded as being totally impermeable to the passage of gaseous or liquid molecules, such permeants still will be able to diffuse around the outsides of such particles. The term “tortuosity” is used to quantify the average increase in diffusion path length when a coating contains impervious particles having various shapes (Wolf and Strieder 1990; Ghanbarian et al. 2013; Simões da Silva et al. 2022). In principle, the greatest contribution to tortuosity will be observed for thin, platy particles. Such a system is illustrated in Fig. 7. Wolf and Strieder (1990) compared the calculated tortuosity values obtained from a large number of empirical studies. Only in the case of platy particles, such as clays, mica, etc., were the benefits in terms of slowing diffusion clearly evident, relative to other potential effects, such as the possible creating of air gaps by the incorporation of particles in a film. Another potentially confusing effect has been demonstrated for a case in which an interfacial layer surrounding a mineral particle (glass beads in the cited work) may provide enhanced permeation, compared to diffusion through the matrix phase (Donkers et al. 2013). In light of the relatively low cost of platy mineral particles, together with the importance of being able to control the rate of water vapor permeation, it would make sense to use such an approach to control water vapor permeation in future formulations of wood coatings. There appears to be a need for research focused on this topic. However, depending on the details of the situation, an excessive barrier against vapor permeation may trap moisture on the inside of the coating, thus promoting rot formation (van Meel et al. 2011).
Fig. 7. Representation of highly platy and somewhat aligned particles in a coating layer, giving rise to enhanced resistance to diffusion of molecules through the layer
Rheological effects
The flow characteristics of a coating formulation are expected to be important, especially during storage, application, and drying or curing phases. During storage, it may be advantageous for the mixture to adopt a transient “house of cards” type of structuring, thereby hindering any tendency for settling of dense particles to the bottoms of containers. Upon mixing, it can be advantageous in some cases for there to be a shear-thinning effect, thus favoring spreading of the mixture as a thin film. In general, the rheological behavior of wood coating formulations can be described as viscoelastic (Mihaila et al. 2022). A higher viscosity also has been shown to slow the permeation of the coating formulation into the pores of wood (Nussbaum et al. 1998; de Meijer et al. 2001a,b; Van den Bulcke et al. 2008; Nejad and Cooper 2010, 2011; Tamantini et al. 2023). When high-aspect particles, such as nanofibrillated cellulose, are used in formulations, one can expect a large contribution to viscosity coefficients, in addition to strong shear-thinning behavior (Hubbe et al. 2017; Song et al. 2023).
Fig. 8. Example of a shear-thinning effect attributable to the progressive alignment of elongated particles in suspension with increasing shear rate (selected data replotted from Nazari et al. (2016).
As illustrated in Fig. 8, the coefficient of viscosity in a suspension of particles tends to decrease as a function of increasing shear rate. The data represented in Fig. 8 are replotted from Nazari et al. (2016), who studied suspensions of nanofibrillated cellulose particles, which have a very high aspect ratio. The downward slope represented by the data indicates a strong shear-thinning effect, which can be attributed to a progressive alignment of the particles with increasing shear rate.
Spreading and Immobilization
The next logical step after preparing the components of the formulation and mixing them together is applying them to the wood (or previous coating layer) surface. Two aspects of this process will be considered here, the wetting of the surface and the flow phenomena. The most common modes of application are by brush, roller, spray, and dipping (Williams et al. 1996). A typical paint brush incorporates some features intended to facilitate holding a large amount of the coating formulation, as well as the accurate placement onto the surface. The ends of the individual bristles of a paint brush are typically flagged (separated into narrower filaments) as the end of the brush is approached, thus providing more surface to accommodate the wet material (Flexner 1994, 2005). In addition, the tip is often cut as a chisel shape. When large amounts of finish need to be transferred, as in the case of some stain applications, a foam pad is recommended (Allen 1984).
Regardless of how the finish is applied, it can be a challenge to evaluate the thickness and other aspects of coating topography. Kibleur et al. (2022) describe a dual-energy computed tomography method, which is based on X-ray imaging. This approach made it possible to differentiate between the coating and the underlying wood fiber material, for the creation of 3D images representing the coating.
Wetting
The contact angle of a droplet of coating formulation on the surface that needs to be covered provides key evidence regarding whether or not adequate spreading of the formulation will occur, resulting in molecular contact between the two phases (Hubbe et al. 2015a). Phenomena such as rough “alligator” surfaces can result when at least some parts of the substrate give rise to contact angles that are near to 90 degrees or higher, thus representing nonwettable parts of the surface. The relatively high interfacial tension of pure water implies that more difficult wetting may be expected with waterborne coating formulations onto certain wood surfaces. Addition of small amounts of surfactant to such formulations will help to overcome such effects (Flexner 1994). A particular challenge is faced when attempting to apply waterborne coatings to surfaces that are both hydrophobic and covered with large pores. In such cases, the finish may fail to penetrate into the pores, as can be predicted by the equation derived by Cassie and Baxter (1944).
Rheological factors and immobilization
The shear-thinning behavior mentioned earlier can become important when the formulation is being applied, especially in the case of spray applications, which can involve relatively high levels of hydrodynamic shear. Once the wet coating formulation has been applied to a surface, its viscosity will affect the degree to which it penetrates into pores in the wood, including fiber lumens and vessel openings. When there is an intention to limit the amount of penetration, it may be favorable if the stationary material gradually develops a higher viscosity, possibly as a means to control or limit the degree of penetration into a porous wood surface. Corcione and Frigione (2012) showed evidence confirming the increased viscosity and shear-thinning behavior of a UV-curable waterborne coating formulation, upon the addition of platy boemite nanoparticles. Likewise, de Meijer et al. (2001b) found that increasing the solids content in a wood coating formulation tended to increase the observed coefficient of viscosity. The coefficient of viscosity can be expected to rise rapidly during the drying and curing of a coating (Fumero and García 2005). Factors that are expected to contribute to an increase of viscosity with the passage of time can include the randomizing of alignment of particles within the formulation, the establishment of structures among polymers or particles (Kulichikhin and Malkin 2022), any crosslinking reactions or effects (Romani et al. 2002), and the evaporation of solvent (Caicedo-Casso et al. 2019).
The formation of a “skin” at the outer surface of a freshly applied coating on wood can be regarded as a specific aspect of immobilization of the coating (Fumero and García 2005). The immediate cause of the phenomena is likely to be the evaporation of the most volatile component(s) in the formulation. The positive aspect of skin formation is that it may shorten the time that has to pass before the object can be touched gently without damaging it. A potential negative aspect is that after skin formation, any moisture that has been trapped under the coating may promote blister formation, especially if heat is being applied (Williams et al. 1996).
Curing Issues
Different coating formulations can cure by different mechanisms, sometimes with more than one process happening at once. The term curing will be used here to denote the transformation from a fresh, wet layer to a layer that is not only dry but approaching its ultimate strength. As noted by Carera et al. (2022), it is generally desirable that a coating should dry relatively quickly under a broad range of environmental conditions. For instance, the coating should be able to withstand a rain event that occurs sooner than expected after its application. On the other hand, it ought not to set so quickly as to make the application difficult or lead to defects. In addition, it has been proposed that the rate of curing of a wood coating needs to be fast enough to prevent excessive penetration into the wood (Rawat et al. 2019). There is a critical need to evaluate the required curing time in the case of systems that are cured by the application of ultraviolet light (Wenning 2002; Karder et al. 2014; Hermann et al. 2021).
It can be argued that the actual curing of a finish must involve interactions among the resin component elements, which may be dissolved polymers or latex particles, etc. Thus, it has been stated that some penetrating finishes, containing only oils, can be regarded as non-curing (Flexner 1994). Such a characterization can be disputed, however, since (as will be discussed below), drying oils such as tung oil and linseed oil are expected to gradually cure by oxidation. According to Flexner (1994), the main curing mechanisms involves with various kinds of coatings are evaporation, reaction, and coalescence.
As suggested by Fig. 9, curing can involve at least three different zones of activity, starting with the evaporation of water, organic solvents, or both. In addition, the initial immobilization of the coating onto the surface can be assisted by capillary wicking and or diffusion of the vehicle into the porous structure of the wool. Certain kinds of coating are able to undergo a chemical reaction (see later section on epoxy resins, for instance). The scheme for the oxidation reaction, which pertains to an alkyd varnish system, is based on work by Miccichè et al. (2006). This is an example showing how oxidation of a highly unsaturated fatty acid species resulted in dimerization. Note that the “R” group shown in the figure represents another similar monomer that also has at least two unsaturated groups along its alkyl chain. By the same mechanism, or by other reaction paths, such oxidation reactions and their subsequent reactions contribute to insolubilization of the coating.
Fig. 9. Three important locations affecting the curing of various wood coatings, namely the adjacent air (evaporation of vehicle), the coating itself (optional curing reactions), and the underlying wood (absorption of vehicle)
Evaporation
Especially in the case of shellac and lacquer finishes (Waring 1963), it is clear that evaporation is the main mechanism by which drying occurs. Mihaila et al. (2019a,b) used a thermogravimetric analysis (TGA) device to track the drying of various commercial varnish films applied to fir wood as a function of time. Models of the drying process, fitted to the data, were judged to be consistent with a diffusion-limited process of drying. Two periods of drying were shown, a rapid drying period (about ten minutes) followed by a decreasing speed of drying period. Song et al. (2023), who used hot-air drying of a waterborne formulation, showed that the drying rate was affected by both the temperature and the humidity.
Figure 10 shows both the lac bug, which produces the crude lacquer compound, and its chemical structure (Ahuja 2024). Using the resins of a host tree, for instance the kusum tree (genus Schleichera), the bug excretes a material that is readily soluble in ethyl alcohol and thus cures by solvent evaporation.
Fig. 10. Lac bug and its excreted product, lacquer, a traditional wood finish that is soluble in ethyl alcohol
Oxidation
Varnishes, which generally are formulated as resins dissolved or colloidally suspended in drying oils, are well known to be slow-curing (Flexner 1994, 2005). A possible explanation is that such coatings require substantial time to complete a chemical reaction. The reaction, involving oxidation, appears to bring about polymerization of the chemically unsaturated oils (Stenberg et al. 2005; Arminger et al. 2020). Stenberg et al. (2005) showed that the results of such curing depended on the detailed structures of the unsaturated oils that were used. Metal salt additives, known as dryers, are often added as a way to accelerate the curing (Flexner 1994; Arminger et al. 2020). Flexner (2005) lists the salts of calcium, manganese, and zinc as some of the most common dryers.
Figure 11 shows a representative structure for linseed oil (Orlova et al. 2021), which is one of the most widely used drying oils for preparation of varnishes. Linseed oil stands out from other vegetable oils has having an especially high level of the most highly unsaturated component, the linolenic acid. This high degree of unsaturation renders it susceptible to a series of curing reactions, which include oxidation. As was shown already in Fig. 10, oxidation can be followed by additional reactions, further oxidation, and radical combination, which contribute to curing of resins that contain linseed oil. Another reaction path starts with hydroperoxide formation (Orlova et al. 2021).
Fig. 11. Representative structure for linseed oil, showing the relative frequencies of three of the major components of the natural product (figure redrawn from an original by Orlova et al. 2021).
Interdiffusion
Cabrera et al. (2022) envision a basic four-step process by which a conventional waterborne latex coating formulation will cure. Initially the wet form on the wood surface will contain a continuous water phase, within which the latex particles are still separate from each other. In the next stage, sufficient evaporation will have taken place so that the latex particles come into contact. Further evaporation causes the initially spherical latex particles to adopt deformed shapes, resembling polygons, in order to fit more closely together. In the last stage, diffusion occurs among polymer segments at the latex particle surfaces, thereby creating a strongly bonded condition.
Another aspect that may hold the key to effective formulation and curing of latex paints and clear finishes is the presence of various slow-to-evaporate solvents, in addition to water. As mentioned earlier, once a substantial amount of water has evaporated, solvents such as glycol ethers or propylene glycol ethers will become enriched in the mixture. The changed solutions conditions can be fine-tuned in a formulation so as to achieve greater solubilization, or at least tackiness of latex (Flexner 1994, 2005). In principle, this mechanism would be expected to facilitate polymer segment interactions among adjacent latex particles, bringing about an interdiffusion of chain segments. This type of process is shown schematically in Fig. 12.
Fig. 12. Schematic diagram of a proposed mechanism of curing of a waterborne latex coating formulation
Time to cure
Cabrera et al. (2022) presented evidence suggesting that colloidal destabilization plays a key role in coating formulations having a quick-drying capability. These formulations were found to rely upon rapid formation of a surface film as a main mechanism. Such a mechanism was found not to require major evaporation of the formulation’s fluid phase. One can envision a process by which relatively rapid evaporation of some of the local solvent (which may be water), leads to coalescence of resin entities present near to the air interface. A question that does not seem to have been considered is whether resin particles, which would tend to be more hydrophobic than the water medium, might be enriched at those surfaces even before evaporation of a freshly applied waterborne coating on wood.
Reactive Curing Systems
A contrasting group of curing mechanisms to be considered here is specific to certain pairs of chemical compounds that are able to react and form covalent bonds with each other in the course of drying of a coating layer. The prime example is epoxy resins. The UV-curing systems also fall into this category, as well are some coating formulations that include cross-linking agents.
Epoxy resins
Epoxy resins are favored for applications in which the coating needs to be able to resist intensive usage, as in the case of a bar surface or tabletop (Allen 1984). As illustrated in Fig. 13, a typical epoxy resin is prepared by first reacting bis-phenol A with epichlorohydrin to form bisphenol A diglycidyl ether (BADGE). The next step is to react the BADGE with a selected multifunctional amine, which undergoes an addition reaction and results in polymerization. Conditions need to be optimized to achieve a prepolymer molecular mass that is still low enough to allow easy pouring and a suitable shelf-life, but high enough so that the addition of a hardener will be able to bring about suitably rapid crosslinking. Hardeners can include phenalkamines, polythiophenes, and other multifunctional amines. The amine groups react with available epoxy rings on the resin. In addition, there can be further reaction between epoxy groups and hydroxyl groups. This chemistry is illustrated in Fig. 13.
Fig. 13. Reaction of bisphenol A diglycidyl ether with a diamine to achieve curing of an epoxy resin
The chemistry of a typical epoxy system is based on combining two liquids, each of which has the potential to be prepared “solvent-free” or mostly just active ingredients. An inherent advantage is that low levels of volatile organic solvent (VOC) emissions can be achieved (Dvorchak 1997; Fortmann et al. 1998; de Meijer 2001). Though epoxy systems can achieve very high film strength and hardness (Motawie and Sadek 1999), they sometimes provide insufficient adhesion at the wood interface. Flexner (2005) recommends using a relatively coarse sanding pretreatment to overcome such deficiencies.
It is also possible to formulate an epoxy system by preparing emulsions, allowing for application in a manner resembling a waterborne latex system (Eyann et al. 2023). The cited article pertains to a mixed system with acrylate chemistry in addition to the epoxy system.
UV-cured finishes
Based on the number of recent publications, there is intense interest in UV-curable formulations. The explanation may lie in some practical advantages. Unlike the epoxy systems, the whole formulation can be prepared as a single batch, and the batch has a relatively long shelf life. In addition, by being less dependent on evaporation to cure the coating, formulators have been able to reduce organic solvents (Agnol et al. 2021; Gustafsson and Börjesson 2007). Landry et al. (2015) demonstrated that the UV light can be applied by means of light-emitting diodes. The function of the UV light, in typical cases, is to catalyze the addition polymerization of an acrylic-type prepolymer or an epoxy acrylate oligomer (Ali et al. 1997). Photoinitiators are used to convert the UV light to odd-electron species, i.e. free radicals (Allen 1996). An example is shown in Fig. 14 (Dossin et al. 2022). The resulting coatings can be optimized to develop moderate durability relative to weathering in varnish formulations (Ayata 2019; Cavus 2021; Gurleyen 2021). Dixit (2021) showed that a successful coating formulation of this type can be prepared with bio-based citric acid, together with other ingredients. Zhang et al. (2022) tested a variety of treatments with silane coupling agents and showed that they generally improved the adhesion performance of waterborne paint. Zhang et al. (2023) found that increasing levels of sol-gel treatment tended to decrease the level of cracks in layer of waterborne polyurethane-modified acrylic acid varnish.
Fig. 14. Example of a photoinitiator, the generation of free-radical species, and the establishment of crosslinking within an acrylic-type resin
Sol-gel systems for wood coatings
The curing process for a typical sol-gel-type coating formulation involves a reaction between a trialkoxysilane reagent and water, followed by further reaction with the -OH groups at a wood surface. Figure 15 provides a general scheme for this kind of reaction system (Hubbe et al. 2015b). At the upper left of the figure, the starting material is assumed to be various alkyl-triethoxysilanes or alkyl-trimethoxysilanes, which are commonly available reagents falling into the category of alkyl-trialkoxysilanes. As a first step, these are added to water, rendering unstable intermediate species. The lower part of the figure indicates successive condensation reactions in which water is lost, leading eventually to three-dimensional cross-linked clusters of increasing size.
Fig. 15. Schematic diagram of a sol-gel condensation of silica species. Figure adapted from Hubbe et al. 2015b, copyright owner
Gholamiyan et al. (2016) studied the feasibility of carrying out such a process as a pre-coat, with the aim of establishing a better adhesive foothold on the wood surface. Indeed, better adhesion was achieved with the subsequent coating layer, compared to its direct application to the wood. Yona et al. (2021a) achieved mixed results, finding better effects of the sol-gel treatment when the surface already had been treated with mineral. Further work (Yona et al. 2021b) showed better performance when using a hydrophobically substituted silane reagent.
Zheng et al. (2015) reported the preparation of tough, abrasion-resistant coatings containing titanium dioxide pigment along with an alkoxysilane to develop a sol-gel curing system. The cross-link density was improved and excellent adhesion to the wood was observed. However, the coating was found to gradually lose its hydrophobic character over time. Members of the same research group (Zheng et al. 2020) showed that a silane coupling agent could be added along with the application of waterborne acrylic resin. That approach was reported to improve both the strength of the coating and its bonding to the surface.
ADHESION TO THE WOOD
The purpose of this section is to look more closely at the topic of adhesion, taking advantage of some of the sub-processes and mechanisms already considered in this review. It can be hypothesized that successful adhesion – especially the adhesion of a coating to the wood, will depend on such factors as the chemistry of the resin, effects due to drying oils, uniformity and colloidal stability issues, wettability issues, as well as the state of curing. As shown in Fig. 16, some of the key contributions to adhesion, at the molecular level, can include covalent bonding, hydrogen bonding, London dispersion forces, and mechanical interlocking between the coating structure and the pores of the wood material. Inadequate adhesion of finish to wood can have serious consequences, including premature peeling and cracking, as well as failure to protect the wood. Before considering several kinds of factors that may contribute to better adhesion of coatings layers, some methods of testing the adhesion will be reviewed.
Fig. 16. Key contributions to the adhesion of a wood coating, on a molecular scale
Testing of Wood Coating Adhesion
Several approaches to assessment of finish layer adhesion have been described in recent publications, and many of these can be described as pull-off methods or assessments based on microscopic observations (Budakçi and Sönmez 2010). Ahola (1991) described “torque tests” in which the shear stress required to remove finish levels was quantified. It was found that pre-weathering, before application of the coating to the wood, had a negative effect on the strength of adhesion, after exposure of the coated surface to further weathering. Follow-up work showed that such adverse effects of pre-weathering could be reduced by prompt application of a pigment stain to the wood, thus allowing the film-forming coating layers to be applied somewhat later (Ahola 1995). Van den Bulcke et al. (2006) used pull-off tests to further study the effects of weathering on coating adhesion. Dilik et al. (2015) used a pull-off test to assess the forces needed to pull polyurethane-based paint layers from the coated particleboard or medium density fiberboard surfaces. Vitosyte et al. (2013) used a pull-off test to evaluate effects of surface roughness on adhesion to an ash wood surface. Karaman et al. (2023) used computer modeling to predict levels of adhesion of waterborne polyurethane finish to oak, chestnut, and pine wood surfaces, as assessed by ASTM D4541.
Knaebe and Williams (1993) describe an improved test based on fracture toughness, which was said to give more convenient and reliable results in comparison to available shear block, lap shear, and tensile test procedures. de Meijer and Militz (2000a) describe a “new method” involving an adhesive tape to allow measurement of the force needed to peel layers of finish from wood. In general, coating formulations achieving better wetting of the wood substrate, as well as greater penetration into the pore structure, were found to give higher resistance to peeling.
Further information regarding adhesion can be obtained by microscopy and related methods. Rijckaert et al. (2001) used fluorescence microscopy to evaluate the penetration of various adhesives into the pores of wood surfaces (Van den Bulcke et al. 2003). Kanbayashi et al. (2019) describe an application of Raman spectrometry, by means of confocal microscopy, to reveal the precise location of resins and pigments from coatings within the microstructure of wood. In this way, it is possible to answer questions related to whether the coating has entered the pore structure.
In view of the diversity of test methods that have been used to quantify the adhesion of coating layers to wood surfaces, no one test method stands out. Choice of an appropriate method can be expected to depend on the specific interests of the investigators. While various pull-off tests tend to be quick and easy to evaluate, they generally do not mimic the conditions likely to result in flaking of wood coatings in actual applications. Conditioning steps before testing, such as artificial weathering (Kropat et al. 2020), may be just as important as the specific test method that is used for quantification of resistance to detachment.
Initial Wetting of Wood by Finish
Wood wettability as necessary but not sufficient
The ability of a film of coating to adhere to a wood surface depends upon both chemistry and structure. One can envision a series of steps, starting with spreading of the liquid formulation, by means of a brush, roller, or spraying, etc. Wetting of the wood by the formulation can be regarded as a necessary but not sufficient criterion for achieving strong adhesion. If there are air gaps, i.e. bubbles or lack of molecular contact, they can provide starting points for peeling or crack development in the coating. Features of roughness, including pores, can provide a mechanical interlocking effect with the applied finish. But all of these aspects that can contribute to good adhesion can be defeated, at least in part, if the wood surface is covered by a sufficient amount of waxy, monomeric extractives, such as triglyceride fats, fatty acids, and other wood resins. These may serve as a weak boundary layer between the finish and the underlying wood.
As illustrated in Fig. 17, a first step in figuring out whether wettability may be playing a determinative role with respect to adhesion of the coating can involve the measurement of contact angle of the fresh formulation. Such measurements have been carried out in various cases (Krystofiak et al. 2016; Miklecic et al. 2022). The work of Miklecic et al. (2022) helps to address concerns of an expected greater difficulty of spreading waterborne formulations onto wood, especially in cases where the wood has a hydrophobic surface. Indeed, the contact angles measured for waterborne formulations were reported to be about three times greater than the corresponding measurements for solvent-based finishes.
Fig. 17. General expectation of favorable wetting of wood surfaces by solvent-based formulations, in contrast to uncertainties about wetting by waterborne formulations, especially in the absence of surfactants
In principle, the presence of a film of water on a wood surface might be expected to block interaction with oil-based formulations. Such an effect is consistent with a recommendation that a wood surface to be painted ought not to have a moisture content higher than 12% (Williams et al. 1996). Rather, even if the surface has been prewetted to address concerns related to undesired swelling of the wood by waterborne coatings, the wood should be dried (and possibly resanded) before application of the finish. An exception to this rule is in cases of waterborne finishes and aqueous staining of some wood surfaces, where sometimes it can be an advantage to work with a water-dampened wood surface (Flexner 1994).
Although contact information, obtained with fresh formulations of coatings, can be used to judge initial wettability, the results of such tests will not necessarily be indicative of success. For instance, in the work of Krystofiak et al. (2016), the wettability information was used to draw conclusions regarding the work of adhesion and other parameters in the final coating; it can be argued that such estimates could be inaccurate due to the fact that the most of the liquid medium, which plays a big role during the initial spreading of the formulation, will have evaporated by the time the coating has solidified. Thus, even though the necessary condition of achieving good contact with the surface can be supported by contact angle information, there are further aspects of the full process of curing that cannot be accurately revealed, in typical cases, by such tests.
Thermal modification and wettability
Due to the increasing popularity of thermally modified wood, studies have been carried out regarding such treatment on the wettability of such products by coating formulations. Such work is highlighted in Table 3. As can be seen, although most of the listed studies showed decreased wettability and/or decreased coating adhesion strength on the heat-treated woods, some of the results were the opposite. Figure 18 illustrates the most common observations. Though at relatively low temperatures, heating may cause hydrophobic extractives to come to the surface (Klinc et al. 2017), higher temperatures can be expected to thermally degrade both the extractives and the hemicellulose (Pratiwi et al. 2019). The wettability of wood heated in the presence of steam increased (Quintos et al. 2023). By contrast, decreased wettability and adhesion can be expected in cases where the heat treatment causes a net accumulation of hydrophobic monomers to the wood surface (Chang et al. 2019).
Table 3. Reported Effects of Thermal Treatment of Wood on its Wettability and Adhesion Characteristics
In light of the effects of heat treatment, as just shown, it is reasonable to ask whether related effects might be obtained just by waiting a long time. Observations related to that concept were reported by Kiliç and Sögütlü (2022) for a large number of adhesives on oak, chestnut, and pine. No significant differences were found with respect to aging time. Another reasonable question to ask is how heat treatment following wood coating will affect coating adhesion. Such a question was addressed by Atar et al. (2015), who evaluated the pull-off strength of various varnishes from beech, oak, poplar, and pine woods. Heat treatment after coating significantly decreased the pull-off strength.
Fig. 18. Typical effects of thermal treatment of wood on its wettability by waterborne finish formulations
Plasma after heat treatment
It is well known that plasma treatment, depending on the gas components, energy level, and various other factors, can greatly affect wettability conditions of surfaces. In particular, corona treatment, involving relatively low-energy plasma with air, is known to oxidize various surfaces and increase their wettability (Zigon et al. 2018). Table 4 lists various studies evaluating the effects of various plasma treatments of wood surfaces on the subsequent wettability or adhesion with coating formulations. The main expected effect of corona treatment is illustrated in Fig. 19, where the contact angle decreases after treatment. The effect can be attributed to the increased number of hydrophilic functional groups, such as -OH and -COO. The high affinity of water for such groups means that waterborne formulations will spread better on the corona-treated wood. Waterborne formulations often do not spread well and fail to penetrate pores on hydrophobic surfaces, such as waxy wood, thereby leading to poor adhesion.
Fig. 19. Expected effect of corona treatment (relatively mild plasma treatment in air) on the wettability of wood by aqueous formulations
Table 4. Reported Effects Plasma Treatment of Wood before Finish Application on its Wettability and Adhesion Characteristics
Wood weathering and wettability
Exposure to weather likewise has been found to decrease the adhesion of various coatings, but the reasons appear to be different. Studies reporting such results are listed in Table 5. As illustrated in Fig. 20, it has been shown that weathering generally removes the lignin from the outermost millimeter of the wood surface (Kropat et al. 2020), thus tending to render the surface more hydrophilic, but weaker.
Fig. 20. Summary of the effects of natural weathering (with UV light and intermittent rain) on the surface of exposed wood. Figure modified from an original copyrighted by Kropat et al. (2020)
In most of the cases listed in Table 5 the authors did not conduct wettability tests. However, the findings of Basri et al. (2022) are consistent with the higher polarity and hydrogen bonding ability of cellulose, compared to the lost lignin.
Table 5. Reported Effects Weathering of Wood before Finish Application on its Wettability and Adhesion Characteristics
Yet another strategy to modify the wetting and adhesion characteristics of wood surfaces involves the addition of resin. Thus, Kielmann and Mai (2016a,b) found that treatment with phenol-formaldehyde resin increased the adhesion of coating. Less blistering, flaking, and cracking were observed during subsequent weathering.
Even if a coating formulation adequately wets of wood surface, an additional requirement is that the macromolecules in the facing surfaces have sufficient mutual solubility so that they can become sufficiently integrated in a three-dimensional manner at the interface (Hansen 2007). Without such three-dimensional integration of polymer segments from the two materials, weak adhesion can be expected. In the case of weathered wood, one can expect a strong contrast between the hydrophilic nature of the cellulose-enriched material at the wood surface and the much less polar resins of typical wood coating formulations. By contrast, lignin is much less hydrophilic and contains aromatic groups that can be expected to match much better to coating resins in terms of the Hansen solubility parameters.
Chemical modification and wettability
Chemical modification of a wood surface, such as acetylation, has the potential to more profoundly affect wettability. As illustrated in Fig. 21, the abundance of -OH groups on typical wood surfaces provides a suitable location for esterification reactions. For example, Pandey and Srinivas (2015) applied acetylation and benzoylation to rubberwood surfaces. After coating with a polyurethane-based finish, there was improved resistance to weathering. The effect was attributed to a high resistance to moisture of the esterified wood surfaces. Alade et al. (2024) observed improved adhesion of pine wood to oil-based coating following its esterification with acetic anhydride, etherification with 1,3-dimethylol-4,5-dihydroxyethyleneurea, or esterification with a sorbitol/citric acid formulation. However, a waterborne coating formulation showed improved adhesion only with the sorbitol/citric acid formulation. Though such results may point in the direction of various useful strategies, one needs to bear in mind that chemical derivatization of wood surfaces is inherently more challenging and expensive than other wood treatments considered in this article.
Fig. 21. Illustration of chemical derivatizations with either acetic acid or long-chain fatty acids to provide esterified wood surfaces having much higher hydrophobic character
Surfactants and wettability
In principle, it would make sense to add wetting agents to decrease the contact angle of waterborne coating formulations when they are applied to wood surfaces (Flexner 1994). A wetting agent generally can be described as a surface-active agent that is insoluble in water but having a sufficient content of hydrophilic groups to interact strongly with water at an interface. Ekstedt (2003) evaluated the effect of using such additives in the case of alkyd emulsion paints. Excessive levels of such surfactants resulted in high water absorption by the applied paint layers. Thus, it will be important to keep the amounts of such wetting agents at a relatively low level, as needed to promote the initial wetting of the wood.
Flow into Pores
Evidence of flow into pores in the wood surface
In order for the resins in a coating formulation to be able to establish a firm connection to wood (especially in the case of a primer coating), it is important that the mixture wets not only the outer surface, but also the surfaces of pores, such as those of vessels, tracheids, and spaces created by the microscopic-scale tearing action of milling or sanding. In other words, adequate adhesion of the first coating layer to the wood may require an optimized degree of penetration into the wood.
Van den Bulke et al. (2003) used laser confocal microscopy to quantify penetration of waterborne coatings into the pores of wood. A range of different penetration depths were observed for a series of different solvent-borne and waterborne formulations, with no clear differences based on which class of coating was being applied. Though the authors drew a curve through their data, purporting to show a very strong positive relationship between penetration depth and maximum adhesive force before failure, a linear regression would have shown a weak relationship. Rijckaert et al. (2001) used confocal microscopy with fluorescent material to compare the penetration of several different coatings. Notably, certain oil-based formulations penetrated to a much greater degree in comparison to the waterborne coating formulations that were evaluated. Such a difference is predicted by the lower surface tension of most nonpolar liquids, which generally results in lower contact angles on solid surfaces, as well as contributing to better wettability (Hubbe et al. 2015a). Related work has been reported by Kanbayashi et al. (2019), who combined confocal microscopy with Raman spectrometry to be able to identify chemical components in the images.
Erich et al. (2006) and Bessières et al. (2013) used X-ray tomography to assess the degree to which different solvent-borne, waterborne, or powder coatings penetrated into the pore structure of wood surfaces. The ability of the different resins to flow into pores were found to be governed by resin type, the solids content, and the methods used for drying. de Meijer et al. (1998, 2001) similarly concluded that flow into pores was affected by the resin type, pigmentation, solids content, and drying speed. Passages allowing inward flow of the formulation included tracheids and ray cells of softwoods. Tyloses in the wood also were found to block the penetration of freshly applied coating. Czarniak et al. (2022) used scanning electron microscopy (SEM) to evaluate paint penetration. They observed no improvement in penetration following plasma treatment of the surface, and they used the SEM evidence to help explain their results. It seems possible that the plasma-induced chemical changes may have been mostly on the outer surfaces and not affecting the wettability of pore surfaces as much. Singh and Dawson (2006) and Singh et al. (2007) showed that the resolution of SEM images showing that such features can be enhanced by use of osmium tetroxide (OsO4), which strongly blocks electrons. They showed that wood coatings can penetrate even into nano-size pores in the wood.
Kibleur et al. (2022) showed that a method called dual-energy computed tomography (DECT) can be used to obtain accurate 3D images of coatings or resins on wood surfaces. This X-ray method was shown to be able to differentiate between the resin, the pigment, and the wood in the obtained images.
Progress in obtaining real-time measurements of the penetration of alkyd, emulsified alkyd, and linseed oil into wood have been obtained by a microautoradiographic method (Nussbaum et al. 1998). It was shown that solvent-borne primers and waterborne alkyd emulsion primers had similar rates of permeation into wood.
Mechanical interlock
According to a mechanical interlocking mechanism of adhesion, it is reasonable to expect that the adhesion of a coating to a wood surface would show a positive correlation to penetration into the pore structure. This idea is consistent with the penetrating nature of primers, which are intended to achieve a strong attachment to the wood surfaces (Allen 1984). de Meijer and Militz (2000a,b) found a positive correlation between greater penetration of wood and increased cohesive failure of the coating during subsequent testing of the adhesion. In other words, the attachment to the wood was found to be stronger than the internal strength of the coating itself.
Evidence supporting the mechanical interlock concept of adhesion of coatings onto wood has been obtained from a study involving the effects of sanding. Sögütlü et al. (2016) found a positive correlation between roughness (achieved with coarser sandpaper) and coating adhesion for three different varnish types. de Moura and Hernández (2005) reported that the adhesion of a high-solids polyurethane coating on sugar maple was favorably affected by sanding. The better adhesion was attributed to surface roughness resulting from the tearing out of fibrils by the abrasive action of the sandpaper. The authors’ explanation appears to be consistent also with the idea that the torn-out features could contribute to pores, as discussed in the previous subsection. Cool and Hernández (2016) likewise found that better adhesion was found when the wood surface had such features as open lumens and fibrillation. Such features were affected by the method of preparing the wood, such as oblique cutting, face milling, and helical planing. Helical planing gave the best adhesive strength. Vitosyte et al. (2013) found correlations between increased roughness of a birch wood surface and increased adhesive strength to an acrylic polyurethane varnish; such results are in qualitative agreement with a mechanical interlock mechanism.
Lucas-Washburn predictions
In principle, the permeation of a fluid into a porous material can be estimated based on the equations of Lucas (1918) and Washburn (1921), as discussed by de Meijer et al. (2001b). As shown in Fig. 22, these equations are based on a simplified model in which cylindrical, smooth pores of uniform diameter are used to represent the porosity of the material. The contact angle governing wetting within the pores is assumed to be the same angle as measured on the exterior of the material.
Fig. 22. Simplified physical model assumed for derivation of the Lucas-Washburn equations for prediction of the rates of wetting of porous solids based on the pore radius, the contact angle, and the fluid viscosity
The capillary force giving rise to penetration into the model pores is resisted by the viscous drag associated with Poisseuille-type laminar flow, which increases in proportion to the distance of penetration into the model pore. The integrated form of the Lucas-Washburn equation predicts a proportionality between the distance penetrated and the square-root of time. The Lucas-Washburn can be written in a derivative form or in an integrated form as shown below,
dL/dt = γLV R cos θ / (4η L ) (1)
L = [(R γLV cos θ] t /(2η)]1/2 (2)
where R is the equivalent radius of the pores (based on the cylindrical pore model), η is the dynamic viscosity, v is the average velocity of fluid flow into the capillary, L is the distance of permeation at time t, and t is the elapsed time after the initial wetting.
The viscosity term in the Lucas-Washburn equation can offer a likely explanation for observations reported by Cheng et al. (2016), who found that addition of a certain kind of nanofibrillated cellulose to a paint formulation had a negative effect on its adhesion to wood. Tamantini et al. (2023) reported similar results when cellulose nanocrystals were added to a waterborne acrylic wood coating. It is known that nanocellulose addition to an aqueous system can give a very strong increase in the coefficient of viscosity (Hubbe et al. 2017). Accordingly, the adverse effect on adhesion might be due to less distance of penetration of the formulation into the pores. Yona et al. (2001a) argued that penetration into pores was prevented by clogging and blockage of pores in the systems that they studied.
The contact angle term of the Lucas-Washburn equation might account for some other effects that have been observed. Xie et al. (2006), among various others, observed more favorable adhesion to wood with a solvent-borne coating formulation compared with a waterborne formulation. It is well known that solvent-borne fluids have very low angles of contact with typical wood surfaces. So, except for the effects of wetting agents, as mentioned earlier, the Lucas-Washburn equation can be used to explain some instances of poorer adhesion of waterborne coatings, especially if the wood is relatively hydrophobic. Work reported by Ghofrani et al. (2016) likewise supported the Lucan-Washburn equation, as well as the degree of penetration into pores, as an explanation for differences in adhesion of coatings. They found that removal of extractives, using a mixture of water and ethanol, resulted in improved adhesion. Such findings are consistent with the expected increase in capillary force, drawing the fluid into the pores. This is because removal of the hydrophobic material by the solvent mixture results in a lower contact angle, which is more favorable for wetting.
Resin Interactions with Wood
The next kind of hypothesis that can be considered is that the strength of adhesion may be significantly dependent on the development of bonds of a chemical nature. In principle, such bonds may be of two forms. One option is that the bonding may involve the diffusional entanglement among polymer segments, as was discussed earlier in this article. The other is that chemical bonding, such as covalent bonds, hydrogen bond interaction between ionic groups of opposite charge, or chemical complexation of various types may be significantly contributing to the observed adhesion.
Diffusional bonding with wood
Diffusional mixing of polymer segments, such as those belonging to adjacent latex particles, were already considered in Fig. 12, which is inspired by the work of Cabrera et al. (2022). Such an interdiffusion of polymer segments in an interfacial region was proposed by Shi et al. (2012) to account for adhesion between polymer fibers at a nano-scale. Krongauz (2021) found evidence of such a mechanism in the autohesion at the interface between plasticized poly(vinyl chloride) surfaces. Such a mechanism may be especially appropriate when envisioning what happens when a fresh layer of finish is being applied on top of another layer, especially when the time between the applications is a matter of hours or days. Such relatively prompt applications of successive coatings has been recommended as a way to favor good inter-layer adhesion (Cassens and Feist 1986). Presumably, a relatively fresh coating will have polymer segments facing the surface that retain a greater degree of mobility, especially when wetted and partly solubilized by the active solvents (Flexner 2005) present in the fresh coating. In such cases the main forces of interaction may be relatively weak, but cumulative van der Waals interactions (Gong et al. 2008; Klatt et al. 2017). Thus, Gong et al. (2008) demonstrated the participation of van der Waals forces during the initial approach of latex particles to each other during the curing of a coating. Klatt et al. (2017) calculated such forces for adhesion of a specific polymer film onto a rough surface. As discussed earlier, it can be expected that diffusional bonding will be strongly favored with the solubility parameters of the polymers associated with the wood (especially the lignin component) and the resins in the coating formulation have similar Hansen solubility parameters (Hansen 2007).
Molecular interactions
More energetic interactions, beyond the ubiquitous van der Waals interactions, have been shown to affect bonding at interfaces in many cases, including the anchoring of coatings onto wood. For instance, Kúdela and Liptáková (2006) carried out calculations that showed a dominant contribution of polar interactions in the adhesion of coating materials to wood. Probst et al. (1997) calculated the work of adhesion at the interfaces between primer and wood and primer and a top layer of polyester varnish. The interactions considered included not only van der Waals forces but also hydrogen bonding and electrostatic interactions. These authors also considered the detailed conformational shapes of different polymers, affecting their abilities to come into efficient molecular contact with each other. The electrostatic forces were shown more explicitly by Renneckar and Zhou (2009) for nanoscale layer-by-layer disposition of polyelectrolyte monolayers onto wood. Though the effects were demonstrated clearly, it can be argued that that type of coating is not representative of the state-of-the-art of wood coating methods.
A covalent bonding effect with the wood surface can be achieved by reaction with a silane reagent, thus achieving a sol-gel curing effect (Tshabalala and Gangstad 2003; Tshabalala et al. 2011; Cheumani-Yona et al. 2015, 2021; Slabejová et al. 2018; Yona et al. 2021a,b; Qu et al. 2023; Zhang et al. 2023). Though the relatively weak Si-C bonds present within such a film structure can represent a point of vulnerability, the films can make up for that by having a three-dimensional glass-like structure. It has been shown that treatment with a silane coupling agent can provide a suitable anchoring point for subsequent layers of coatings on wood (Zhang et al. 2022, 2023; Tao et al. 2024). Zheng et al. (2020) demonstrated a system in which a silane coupling agent improved the performance of an acetate-butyrate modified acrylic waterborne coating, presumably by a crosslinking effect.
Weak Boundary Layer Issues
All of the mechanisms of interfacial adhesion described above can fail in cases where the wood surface already is covered by weakly adsorbed monomeric substances such as waxes and wood resins. Monomeric wood extractives are often present at wood surfaces, especially after the passage of time, and their amount will depend on the wood species (Allen 1984; Williams et al. 1996; Kimerling and Bhatia 2004; Cordeiro et al. 2023). Such substances have been found to gradually diffuse and accumulate at the surface exposed to air (Hiltunen et al. 2008; Englund et al. 2009). This effect is illustrated in Fig. 23. Such an effect helps to explain why adhesion to a coating layer is often improved by fresh sanding of the surface (Cool and Hernández 2011; Cordeiro et al. 2023), thus removing the surface accumulation.
Fig. 23. Expected effect of a “weak boundary layer”, often consisting of monomeric compounds accumulated on a surface and getting in the way of intended adhesion
A strikingly different kind of weak layer between a coating layer and the surface of wood can result from the gradual breakdown of the lignin due to ultraviolet light, which to some extent will be able to pass through the finish. Kanbayashi et al. (2023) demonstrated such an effect by artificial weathering of cedar wood that had been coated with four common types of exterior finishes. Subsequent confocal Raman spectrometry revealed the degradation of lignin near to the paint interface, leaving the film vulnerable to detachment.
DEVELOPMENT OF FILM PHYSICAL PROPERTIES
Overview
In order to perform its function, an applied film of coating on wood needs to reorganize itself in response to evaporation and permeation processes. By such means it needs to solidify and develop strength, it needs to develop its hydrophobic or hydrophilic surface attributes, it will develop its appearance attributes, and it might even be designed so that it acquires a self-healing ability. Such issues will be considered in this section.
Self-assembly Overview
At the moment that a fresh layer of coating has been applied to a wood surface, or on top of previous such layers, various processes can take place that are no longer under human control. Rather, the dissolved polymers, particles, and emulsified matter within the formulation will move around in various ways during such processes as evaporation, diffusion into pores, and reorientation of amphiphilic compounds at phase boundaries. The sorting-out processes involved in such diffusional movements, especially if it leads to somewhat organized and beneficial effects, has been called self-assembly (Hubbe et al. 2023).
A key principle of self-assembly is illustrated in Fig. 24. As shown, it is proposed that certain of the component parts of the system to be assembled have a tendency to fit together in a prescribed favorable fashion. What is generally required for self-assembly to happen is a suitable level of mobilization of the parts, such that they are able to assume a large number of intermediate conditions on the way to adopting an intended structure.
Fig. 24. Schematic illustration of a self-assembly process, which depends on the presence of a certain level of chaotic energy and the sampling of different possible fits. Figure reused from Hubbe et al. 2023; copyright by the authors
As one example of how interactions at the nanoscale can affect the properties of coating layers, it has been shown that incorporation of ZnO nanoparticles was able to increase the glass transition temperature of an acrylic stain, once it had been evaporatively cured (Cristea et al. 2010). The effect was attributed to the ability of the nanoparticles to induce an increased level of crystallinity within the polymer material during its curing. Another example involved the development of a highly ordered pattern of latex spheres in a coating on wood, such as to present so-called “structural colors” that do not require any chromophoric materials or selective absorption of light (Liu et al. 2020). Song et al. (2024) reported the preparation of uniform lignin/titanium dioxide nanoparticles, by confined self-assembly, and their subsequent usage as an additive in polyurethane wood coatings. It is also possible to directly utilize self-assembled lignin nanoparticles (Trovagunta et al. 2024) as an additive in wood coatings (Song et al. 2022).
Hardness
The hardness of a wood coating can be important in various applications where the surface needs to be able to resist effects of tableware, footwear, and writing with a pen, to name a few. In principle, hardness can be regarded as being dependent on both the nature of the cured resin and any reinforcing particles in the formulation, as well as how the particles are arranged. Figure 25 illustrates some of the general principles that can be used to achieve increases in hardness in a formulated coating. There are a great variety of resin types used in wood coatings; though their properties are to some extent known, there is likely to be variation between and within different brands that nominally have the same identification. Likewise, the hardness of recycled versions of plastics to be used in wood coating formulations will need to be evaluated (Godinho et al. 2021). The hardness of a coating can be expected to change over time, due either to further curing or to other effects, such as weathering. Notably, Gunduz et al. (2019) found that certain coatings actually became harder in the course of weathering. The application of a coating, such as a varnish, often increases hardness (Pelit et al. 2023), which is consistent with the filling in of pore spaces need to the wood’s surface; however, the authors found cases in which a soft coating had the opposite effect of rendering the coated surface less hard. In general, factors that tend to increase the glass transition temperature and the crystallinity index of the cured resin in a coating can be expected to give higher hardness (Veigel et al. 2014).
Fig. 25. Some basic ways to achieve increases in hardness of a formulated coating
Role of crosslinking
Crosslinking can be regarded as a primary tool by which to increase the hardness of a wood coating. This is a particularly important issue in the coating of floors and furniture, which can encounter scuffing and impacts during normal usage. A notable feature of UV-cured coatings, which are popular for factory-prepared furniture and ready-to-use flooring, is that they provide extensive crosslinking (Hermann et al. 2021). Such systems are also known for their potential to achieve high hardness (Landry et al. 2015). Hermann et al. (2021) showed that high crosslink density, as needed to achieve high hardness, can be achieved by optimization of the mix of monomers and oligomers employed.
Addition of reinforcing materials to the formulation
A further strategy that can be used to promote hardness is to optimize the type and level of mineral content, noting the high hardness of certain minerals. For instance, Zhou et al. (2023) reported achievement of high hardness and scuff resistance when adding nano silicon carbide, a very hard mineral, to a UV-curing formulation for flooring. However, Ersoy et al. (2021) showed, somewhat surprisingly, that a strong contribution to hardness can be achieved with calcium carbonate, which is relatively low in cost. To go yet further in that unexpected direction, Yan et al. (2022a) found that incorporation of the relatively soft material rice husk powder, formulated with melamine, increased the hardness after curing of a mixture that included shellac microcapsules and a waterborne primer. Such findings appear to teach that the reinforcement itself may be as important to the development of hardness as is the hardness of the materials used for reinforcement. Landry et al. (2010) reported some hardness increases when using clay mineral at the 3% level in acrylate formulations; however, under many of the texted conditions there was no benefit to hardness when adding the mineral to the formulation. A related approach is to consider the highly platy nano-scale material graphene oxide; Xu et al. (2022) showed that its incorporation into a waterborne polyurethane paint formulation led to significant increases in hardness, which might be attributed to their efficient reinforcement in a type of nanocomposite.
Plasticizers, glass transition temperature, and flexible resins
A potential downside of high hardness is a possible associated tendency for brittleness. Thus, it can be important to consider strategies that can be used to decrease the cured hardness of a coating. One of the key tools by which this can be done is by incorporation of a plasticizer (Cox 2003; Flexner 2005). In very general terms, a plasticizer can be defined as a non-crystalline compound having low volatility that has good solubility with the rest of the components in the formulation. For example, Flexner (2005) noted the common usage of cellulose acetate butyrate in nitrocellulose-based lacquer as a means to reduce the brittle character to an optimal extent.
Another approach to decreasing the hardness of a coating is to use an inherently flexible resin component (He et al. 2022). This can mean selecting a resin having a lower glass transition temperature (Nejad and Cooper 2010, 2017). In general, latex-type waterborne coatings are likely to be more flexible, and thus more resistant to cracking, than their oil-based counterparts (Williams et al. 1996).
Self-healing Attributes
Capsules is a coating formulation component
A recurring problem associated with wood coatings is their eventual development of cracks, peeling, or other damage. Several investigators have considered the possibility of developing a type of coating that has the ability to cure itself after being subjected to such damage. Highlights of such studies are provided in Table 6. A common feature of most such systems that have been considered is that they contain microcapsules filled with active coating formulation. Thus, in principle, when an event occurs causing damage to the coated surface, some of the local capsules will be ruptured, allowing for some filling of the crack or damaged zone.
Table 6. Progress in the Development of Self-Healing Wood Coatings
As illustrated in Fig. 26, self-assembly offers a different strategy by which to design self-healing systems. As shown, such mechanisms depend on the component parts of the system to be in a state of activation, making it possible for the system to sample many different arrangements of structure. The final structure, after a period of trial and error, can be governed by preferential rules of fitting, making some arrangements more energetically favorable. It is worth noting that none of the studies listed in Table 6 employed this kind of strategy. A possible reason is that typical coatings on wood may be so completely immobilized as to block such self-healing effects.
Fig. 26. Self-healing principle that is inherent in certain self-assembly mechanisms. Figure reused from Hubbe et al. 2023; copyright by the authors
COMPROMISES INHERENT IN ECO-FRIENDLY FINISHES
Challenges for Waterborne Finishes
As stated at the outset of this article, the decades since the late 1990s have witnessed a dramatic shift in wood coating formulations, going from exclusively oil-based mixtures to the present dominance of waterborne systems, especially in products designed for use by the general public. Various challenges, as well as successes, involving such systems already have come up in the course of considering different topics in this review article. The purpose of this section is to take stock of the present state of the art. Due to the advantages in terms of low volatile organic solvents and easy cleanup with soap and water, it can be expected that waterborne coatings for wood will continue their dominant position for the foreseeable future.
Among the issues that seem most in need of research input, relative to the performance of waterborne coating formulations in general, are the undesired swelling of the wood, the relatively high permeability of typical waterborne coatings, and opportunities to further improve the environmentally friendly character of such coating formulations.
Undesired Swelling of the Wood during Coating
Fiber-rise issues
As has been noted in this review, an inherent problem associated with the usage of waterborne coating formulations is related to the swelling of the wood when it is wetted by water (Flexner 1994, 2005; Cox 2003; Landry et al. 2013). For instance, Fig. 27 illustrates a problem called “fiber-rise,” in which the swelling of wood fibers in water causes them to protrude from a sanded surface of the wood. As shown, such protruding fibers often can exceed the thickness of the cured coating, which can result in a loss of gloss and a defective appearance. The practical effect is that the grain pattern of the wood can be raised in the course of its wetting by a water-based coating formulation (Landry et al. 2013). Ways to mitigate the problem are often labor-intensive and requiring extra time. For instance, the wood surface can be pre-wetted, allowed to dry overnight, sanded, and only then coated with a first waterborne layer (Flexner 2005). Although the mentioned procedure can be expected to decrease effects associated with the raising of the grain of the wood, it is not likely to completely resolve the problems. The topic of the swelling of cellulosic materials, including wood itself, was recently reviewed (Hubbe et al. 2024). A key issue highlighted in that review is the fact that the swelling of wood is strongly anisotropic, which results in a much greater expansion perpendicular to the grain direction.
Fig. 27. Illustration of fiber-rise caused when a sanded wood surface is wetting by a waterborne finish formulation
With an aim to strike a balance between a high quality of the coating, including its smoothness and adhesion to the wood, along with the minimization of environmental harm, it appears that one of the most promising approaches may be to optimize a suite of ethanol-based primer formulations that can be applied to the wood first. Ethanol is relatively benign, and it has been used successfully for the application of some stains, thus achieving less distortion and “fiber rise” problems at that stage of wood treatment (Cox 2003). Although in such applications it is clear that ethanol can be an effective means of distributing certain stain compounds, especially some of those that already can be distributed by water, it is unclear whether ethanol can be used as a substitute fluid carrier for the resins or resin-oil blends that are associated with many modern wood finishes. In particular, it may be inherently problematic to distribute a latex formulation in that way. Rather, the opportunity might lie in the development of a more narrowly focused class of primers to be used in cases where the goal is to minimize wood swelling in the course of applying the second and later coating layers, which can be expected to be waterborne in a high proportion of cases. Solubility principles (Hansen 2007), as outlined earlier, might be used as a means of narrowing down candidate resins suitable for such alcohol-borne primers.
Permeability
The relatively high permeability of many waterborne coatings for wood has been noted as an ongoing concern (Allen 1984; Ekstedt 2003; Donkers et al. 2013; Volkmer et al. 2015; Hysek et al. 2018). There appears to be a need for coating products that can achieve much higher barrier performance, relative to vapor exchange, while still retaining the environmental benefits of using water as the main solvent and carrier (Gezici-Koç et al. 2019). A more effective barrier to water vapor can be expected to decrease the extent of periodic swelling and shrinkage in response to changes in weather (Gezici-Koç et al. 2019). A particular challenge may reside in the fact that wetting of a coating layer, as in the case of rain, may render it yet more permeable to water vapor, presumably due to its swelling (Hysek et al. 2018).
A further incentive to developing coating formulations that offer greater resistance to permeation is that such coatings might simultaneously provide a better answer to observed problems with the bleeding of wood resins through waterborne coating layers (Allen 1984; Cassens and Feist 1986; Williams et al. 1996; Williams and Feist 1999; Kimerling and Bhatia 2004; Williams et al. 2005; Englund et al. 2009; Darmawan et al. 2018; Wu et al. 2020; Coniglio et al. 2023a,b; Dvorak et al. 2023). Wood resins rising to the surface may become more important, depending on the likely growth in popularity of some thermally modified wood products (Altgen and Militz 2017). Presently it appears that one of the best ways to address such issues is with a prime coating of pigmented shellac (Allen 1984), which falls into the category of an oil-based formulation.
One path by which developers of future waterborne coating formulations could take as a strategy to reduce vapor permeability would be to employ highly platy particles. Such materials might include exfoliated sodium montmorillonite, which has been shown to have an especially high ratio of plate diameter to thickness (Ploehn and Liu 2006). It can be dispersed in water and has been shown to be effective for developing vapor barrier properties (Bangar et al. 2023). Work by Cristea et al. (2010) already has shown that the incorporation of very different kinds of mineral particles, nano TiO2 and ZnO2, already can achieve some of the desired effect of decreasing the permeability of the coating. One of the challenges that needs to be faced during such developmental efforts is the possibility that vapor may be preferentially transported in a layer at the surface of the solid particles (Donkers et al. 2019). But presumably if the particles have a sufficiently high aspect ratio, such effects may be less important than the increased tortuosity, i.e. a much higher path length that diffusing molecules need to take in order to diffuse through the coating layer (Wolf and Strieder 1990; Ghanbarian et al. 2013; Simões da Silva et al. 2022).
Another angle by which future coating formulations might be optimized to address concerns of coating layer permeability is by focusing on the water uptake of the coating layer itself. Work by Gibbons et al. (2020) showed strong correlations between such water uptake and the surface tension and solids content of the coating formulations. Presumably, the degree of swelling of the coating itself, upon exposure to water, can be addressed by increased crosslinking and perhaps by increased hydrophobic composition.
Eco-friendly Coatings
Though the emergence of waterborne coatings for wood can be regarded as an important step forward in the direction of minimizing environmental impacts, there appear to be many opportunities for further progress in the direction of minimizing adverse environmental impact. These include further efforts to reduce volatile organic carbon (VOC) emissions and the replacement of some petroleum-sourced components with plant-based components. Figure 28 emphasizes some of the key goals with respect to achieving a high level of sustainability when using wood products, noting that wood coatings will play a role in such issues.
Fig. 28. Overall view of sustainability goals for wood products, including any effects related to the coatings. Part of this figure adapted from Hubbe (2023)
Plant-sourced components in coating formulations
Some progress in the development of coatings based on plant materials is highlighted in Table 7. In addition to plant-based approaches, there has been a specific concern to replace toxic isothionate products. A class of non-isothionate polyurethanes has been described by Zareanshahraki and Mannari (2021). Another such issue concerns the kinds of plasticizing agents used in lacquer coating products made with polyester resin and drying oils (Alves et al. 2023). The glycol ethers used for that purpose are known to be more toxic than alternatives such as propylene glycol ethers (Flexner 2005).
Table 7. Progress in the Development of Plant-based Wood Coatings
CONCLUDING STATEMENTS
A series of hypotheses were introduced at the beginning of this article. Now, after having reviewed published literature, these hypotheses can be reconsidered.
1. Understanding wood’s degradation and the factors that affect it can provide a foundation for understanding the role of protective finishes.
Support for this hypothesis mainly lies outside of the present work, which did not focus on reviewing wood degradation. Much is known about wood weathering from research that has been considered in earlier review articles (Feist 1982; Cogulet et al. 2018; Kropat et al. 2020). In addition, coating systems designed for exterior usage on wood will be considered in a future review article by the present authors.
2. By blocking all or most rainwater from wetting the wood itself, a finish is able to decrease the swings of cyclical changes in moisture content that would have been experienced by unprotected wood exposed to outdoor weather.
This hypothesis has been well supported by findings reported in the literature, especially the work of Lu and Lin (2008), de Meijer and Militz (2001a,b), Gezici-Koç et al. (2019), and Bobadilha et al. (2020).
3. By absorbing UV light, a suitably formulated finish can avoid or slow down the photodegradation of lignin near the surface of the wood, thus helping to maintain the integrity of the wood surface.
This hypothesis is well supported by numerous studies involving the addition of UV-absorbing additives in wood coatings. Such treatments will be considered in greater depth in a future review article by the present authors.
4. By presenting a suitably hard and contiguous film at the wood surface, the finish can protect the wood from scratching and abrasion.
This hypothesis is well supported by findings cited in this article, especially with respect to the development of hardness in wood coatings.
5. By drying and curing in a way that develops smoothness, the finish can contribute to a glossy appearance, and the level of gloss can be adjusted by formulation of the finish mixture.
This hypothesis, while agreeing with common experience, was not specifically addressed in the present work. Rather, it will be a key focus of a future review article by the present authors that deals with the coating of wood for indoor applications.
6. Surface preparation steps and many other details are foundational to achieving film and reliable adhesion of the finish to the wood to achieve an expected service life of multiple years or even decades in some cases.
As discussed in this article, key preparation steps will include fresh milling or sanding. One of the effects of such mechanical preparation is the exposure of surfaces not blocked by monomeric wood resins, which in some cases can present a weak interface, preventing the establishment of good adhesion. Also, published findings indicate that a moderately coarse final sanding is likely to provide better mechanical anchoring of the finish, especially in cases where there might be a problem.
7. To perform its various functions, an effective layer of finish will need to have undergone a curing process, which can involve evaporation of one or more carrier fluids (drying), oxidation and other chemical changes, and even covalent crosslinking in some cases.
This hypothesis is well supported by the reviewed findings. Curing processes that involve evaporation can be understood as also involving the self-assembly of polymeric segments and particulate components. Diffusional motions during evaporation of the carrier fluid are understood to be essential to achieve the needed density and interweaving among polymer segments to establish strong films and adhesion to the wood. In some other cases, curing can involve the establishment of chemical bonds.
8. To be able to spread effectively onto either wood or a previous layer of finish, the next layer to be spread needs to be formulated with a suitable surface tension, as well as having resins able to spread and interact with the functional groups facing the surface.
The present review of the literature did not find evidence of cases in which a fresh layer of coating formulation was not able to spread adequately onto another widely used type of wood coating. On the other hand, there will still be a need for studies that address ways to overcome poor wetting, spreading, and permeation of coating formulation onto and into the pore spaces of wood specimens having hydrophobic surfaces, often as a result of wood resins.
7. A final hypothesis to be considered in this article can be attributed to Flexner (1994, 2005). The background for this hypothesis involves the design of waterborne wood finishes. The general expectation of a waterborne finish is that it ought to perform its various roles equally well in comparison to traditional oil-borne formulations. The hypothesis states that a well-performing waterborne finish formulation will contain not only water, but also a minor amount of a slower-to-evaporate vehicle, which in addition needs to be a good solvent for the polymeric binder in the finish. Such a system can be expected to permit intermixing among polymer segments of adjacent macromolecules, thus resulting in a contiguous film.
While this hypothesis appears to be sound, it does not appear to have been the focus of critical studies. Due to the likely importance of such a mechanism relative to the further development of waterborne coatings for wood, it is recommended that further research should focus on confirming and explaining the role of solvent components that are less volatile than water and able to bring about tackiness of resin components either in a continuous manner or especially in the situation where their concentration has been increased due to the evaporation of water during the curing process.
ACKNOWLEDGMENTS
The authors gratefully acknowledge the following people who studied an earlier version of the article and contributed corrections and suggestions: Asniza Mustapha, Forest Products Division, Forest Research Institute Malaysia (FRIM), 52109 Kepong, Selangor, Malaysia; Mohd Khairun Anwar Uyup, Forest Products Division, Forest Research Institute Malaysia, 52109 Kepong, Selangor, Malaysia; Oladele Nathaniel Amoo-Onidundu, Forestry Research Institute of Nigeria; Mehmet Karamanoglu, Kastamonu Üniversitesi, Türkiye; and Varun Thakur, Dr. Yashwant Singh Parmar University of Horticulture and Forestry Nauni, Solan (HP), India.
REFERENCES CITED
Acarali, N., and Demir, S. (2021). “Physical and chemical effects of quartet structure (bamboo/zinc borate/shellac/surfactant) on organic coatings,” J. Indian Chem. Soc. 98(3), article 100043. DOI: 10.1016/j.jics.2021.100043
Agnol, L. D., Dias, F. T. G., Ornaghi, H. L., Sangermano, M., and Bianchi, O. (2021). “UV-curable waterborne polyurethane coatings: A state-of-the-art and recent advances review,” Prog. Organic Coatings 154, article 106156. DOI: 10.1016/j.porgcoat.2021.106156
Ahola, P. (1991). “Adhesion between paint and wood substrate,” JOCCA – Surf. Coatings Int. 74(5), 173-176.
Ahola, P. (1995). “Adhesion between paints and wooden substrates – Effects of pretreatments and weathering of wood,” Mater. Struc. 28(180), 350-356. DOI: 10.1007/BF02473151
Ahuja, A. (2024). Utilization of Naturally Occurring Shellac-based Materials for Paper-based Packaging of Liquid Products, Ph.D. Dissertation, Dept. of Paper Technol., Indian Inst. Technol. Roorkee, India.
Akkus, M., Akbulut, T., and Candan, Z. (2019). “Application of electrostatic powder coating on wood composite panels using a cooling method. Part 1: Investigation of water intake, abrasion, scratch resistance, and adhesion strength,” BioResources 14(4), 9557-9574. DOI: 10.15376/biores.14.4.9557-9574
Akkus, M., Akbulut, T., and Candan, Z. (2022). “Formaldehyde emission, combustion behavior, and artificial weathering characteristics of electrostatic powder coated wood composite panels,” Wood Mater. Sci. Eng. 17(6), 540-550. DOI: 10.1080/17480272.2021.1901142
Alade, A. A., Hoette, C., and Militz, H. (2024). “Coatings adhesion on chemically modified Scots pine (Pinus sylvestris L.) woods,” Forests 15(3), article 526. DOI: 10.3390/f15030526
Ali, K. M. I., Khan, M. A., Rahman, M., and Ghani, M. (1997). “Ultraviolet curing of epoxy coating on wood surface,” J. Appl. Polymer Sci. 66(10), 1997-2004. DOI: 10.1002/(SICI)1097-4628(19971205)66:10<1997::AID-APP16>3.0.CO;2-S
Alinejad, M., Henry, C., Nikafshar, S., Gondaliya, A., Bagheri, S., Chen, N. S., Singh, S. K., Hodge, D. B., and Nejad, M. (2019). “Lignin-based polyurethanes: Opportunities for bio-based foams, elastomers, coatings and adhesives,” Polymers 11(7), article 1202. DOI: 10.3390/polym11071202
Allen, N. S. (1996). “Photoinitiators for UV and visible curing of coatings: Mechanisms and properties,” J. Photochem. Photobiol. A – Chem. 100, 101-107. DOI: 10.1016/S1010-6030(96)04426-7
Allen, S. (1984). Wood Finisher’s Handbook, Sterling Pub. Co., New York.
Altgen, M., and Militz, H. (2017). “Thermally modified Scots pine and Norway spruce wood as substrate for coating systems,” J. Coatings Technol. Res. 14(3), 531-541. DOI: 10.1007/s11998-016-9871-8
Alves, T. R., Pacheli Heitmann, A., Ayres, E., Coura, I. R., de Souza, P. P., and de O. Patricio, P. S. (2023). “Waterborne polyurethane dispersions based on polypropylene glycol/polycarbonate diol: Evaluation of their use as wood protective coatings,” J. Appl. Polym. Sci. 140(40), article e54530. DOI: 10.1002/app.54530
Arminger, B., Jaxel, J., Bacher, M., Gindl-Altmutter, W., and Hansmann, C. (2020). “On the drying behavior of natural oils used for solid wood finishing,” Prog. Organic Coatings 148, article 105831. DOI: 10.1016/j.porgcoat.2020.105831
Atar, M., Cinar, H., Dongel, N., and Yalinkilic, A. C. (2015). “The effect of heat treatment on the pull-off strength of optionally varnished surfaces of five wood materials,” BioResources 12(4), 7151-7164. DOI: 10.15376/biores.10.4.7151-7164
Ayata, Ü. (2019). “Effects of artificial weathering on the surface properties of ultraviolet varnish applied to lemonwood (Citrus limon (L.) Burm.),” BioResources 14(1), 8313-8323. DOI: 10.15376/biores.14.4.8313-8323
Ayrilmis, N. (2022). “A review on electrostatic powder coatings for the furniture industry,” Int. J. Adhesion Adhesives 113, article 113. DOI: 10.1016/j.ijadhadh.2021.103062
Bangar, S. P., Whiteside, W. S., Chaudhary, V., Akhila, P. P., and Sunooj, K. V. (2023). “Recent functionality developments in montmorillonite as a nanofiller in food packaging,” Trends Food Sci. Technol. 140, article 104148. DOI: 10.1016/j.tifs.2023.104148
Bansal, R., Nair, S., and Pandey, K. K. (2022). “UV resistant wood coating based on zinc oxide and cerium oxide dispersed linseed oil nano-emulsion,” Mater. Today Commun. 30, article 103177. DOI: 10.1016/j.mtcomm.2022.103177
Basri, E., Martha, R., Damayanti, R., Rahayu, I., Darmawan, W., and Gérardin, P. (2022). “Durability and wettability of varnishes on the modified and aged surfaces of short rotation teak wood,” Pigment Resin Technol., Early access. DOI: 10.1108/PRT-09-2022-0110
Baumstark, R., and Tiarks, F. (2002). “Studies for a new generation of acrylic binders for exterior wood coatings,” Macromol. Symp. 187, 177-186. DOI: 10.1002/1521-3900(200209)187:1<177::AID-MASY177>3.0.CO;2-1
Bessières, J., Maurin, V., George, B., Molina, S., Masson, E., and Merlin, A. (2013). “Wood-coating layer studies by X-ray imaging,” Wood Sci. Technol. 47(4), 853-867. DOI: 10.1007/s00226-013-0546-7
Bobadilha, G. S., Stokes, C. E., Kirker, G., Ahmed, S. A., Ohno, K. M., and Lopes, D. J. V. (2020). “Effect of exterior wood coatings on the durability of cross-laminated timber against mold and decay fungi,” BioResources 15(4), 8420-8433. DOI: 10.15376/biores.15.4.8420-8433
Bottcher, P. (1984). “Improving longterm behavior of exterior wood structures made from domestic wood species by renewal paint coating,” Holz als Roh- und Werkstoff 42(3), 85-91. DOI: 10.1007/BF02628859
Brischke, C., Bayerbach, R., amd Rapp, A. O. (2006). “Decay-influencing factors: A basis for service life prediction of wood and wood-based products,” Wood Mater. Sci. Eng. 1(3-4), 91-107. DOI: 10.1080/17480270601019658
Budakçi, M., and Sönmez, A. (2010). “Determining adhesion strength of some wood varnishes on different wood surfaces,” J. Faculty Eng. Archit. Gazi Univ. 25(1), 111-118.
Bulian, F., and Graystone, J. A. (2009). Wood Coatings: Theory and Practice, 1st Ed., Elsevier, Amsterdam. DOI: 10.1016/B978-0-444-52840-7.00001-1
Cabrera, I., Rückel, M., Boyko, V., Baumstark, R., and Willerich, I. (2022). “Quick curing mechanisms for all-season paints and renders,” Materials 15(20), article 7397. DOI: 10.3390/ma15207397
Caicedo-Casso, E., Sargent, J., Dorin, R. M., Wiesner, U. B., Phillip, W. A., Boudouris, B. W., and Erk, K. A. (2019). “A rheometry method to assess the evaporation-induced mechanical strength development of polymer solutions used for membrane applications,” J. Appl. Polymer Sci. 136(6), article 47038. DOI: 10.1002/app.47038
Cassens, D. L., and Feist, W. C. (1986). Finishing Wood Exterior: Selection, Application, and Maintenance, U. S. Dept. of Agriculture, Forest Service, Washington, DC.
Cassie, A. B. D., and Baxter, S. (1944). “Wettability of porous surfaces,” Trans. Faraday Soc. 40, 546-551. DOI: 10.1039/tf9444000546
Cavus, V. (2021). “Weathering performance of mulberry wood with UV varnish applied and its mechanical properties,” BioResources 16(4), 6790-6798. DOI: 10.15376/biores.16.4.6791-6798
Chang, C. W., Kuo, W. L., and Lu, K. T. (2019). “On the effect of heat treatments on the adhesion, finishing and decay resistance of Japanese cedar (Cryptomeria japonica D. Don) and Formosa acacia (Acacia confuse Merr. (Leguminosae)),” Forests 11(7), article 586. DOI: 10.3390/f10070586
Chang, C. W., and Lu, K. T. (2012). “Natural castor oil based 2-package waterborne polyurethane wood coatings,” Prog. Organic Coat. 75(4), 435-443. DOI: 10.1016/j.porgcoat.2012.06.013
Chang, Y. J., Yan, X. X., and Wu, Z. H. (2023). “Application and prospect of self-healing microcapsules in surface coating of wood,” Colloid Interface Sci. Comm. 56, article 100736. DOI: 10.1016/j.colcom.2023.100736
Chen, K., Zhu, H. X., Zhang, Z. Y., Shao, Y. Q., Yu, Q. H., Cao, X. L., Pan, S. Y., Mu, X., Gao, Z. H., and Wang, D. (2023). “Self-healing polyurethane coatings based on dynamic chemical bond synergy under conditions of photothermal response,” Chem. Eng. J. 474, article 145811. DOI: 10.1016/j.cej.2023.145811
Cheng, A. K., Liu, Z. L., Qiu, Y. H., Xiang, J. P., Zhou, Y. H., Lin, X. Y., Wang, X. J., Tu, D. Y., and Zhou, Q. F. (2024). “Effect of unilateral surface-densification process on wood-surface coating properties,” Wood Mater. Sci. Eng. Early access. DOI: 10.1080/17480272.2024.2314754
Cheng, A. K., Tu, D. Y., Zhu, Z. P., Zhou, Q. F., Wei, W., Hu, C. S., and Liu, X. H. (2023). “Study on machining properties and surface coating properties of heat treated densified poplar wood,” Wood Res. 67(6), 1032-1045. DOI: 10.37763/wr.1336-4561/67.6.10321045
Cheng, D., Wen, Y. B., An, X. Y., Zhu, X. H., and Ni, Y. H. (2016). “TEMPO-oxidized cellulose nanofibers (TOCNs) as a green reinforcement for waterborne polyurethane coating (WPU) on wood,” Carbohyd. Polym. 151, 326-334. DOI: 10.1016/j.carbpol.2016.05.083
Cheumani-Yona, A. M., Budija, F., Hrastnik, D., Kutnar, A., Pavlic, M., Pori, P., Tavzes, C., and Petric, M. (2015). “Preparation of two-component polyurethane coatings from bleached liquefied wood,” BioResources 10(2), 3347-3363. DOI: 10.15376/biores.10.2.3347-3363
Cheumani-Yona, A. M, Zigon, J., Kamlo, A. N., Pavlic, M., Dahle, S., and Petric, M. (2021). “Preparation, surface characterization, and water resistance of silicate and sol-silicate inorganic-organic hybrid dispersion coatings for wood,” Materials 14(13), article 3559. DOI: 10.3390/ma14133559
Chevalier, Y., and Bolzinger, M.-A. (2013). “Emulsions stabilized with solid nanoparticles: Pickering emulsions,” Colloids Surf. A: Physicochem. Eng. Aspects 439, 23-34. DOI: 10.1016/j.colsurfa.2013.02.054
Clausen, C. A. (1996). “Bacterial associations with decaying wood: A review,” Int. Biodeter. Biodegrad. 37(1-2), 101-107. DOI: 10.1016/0964-8305(95)00109-3
Cogulet, A., Blanchet, P., Landry, V., and Morris, P. (2018). “Weathering of wood coated with semi-clear coating: Study of interactions between photo and biodegradation,” Int. Biodeter. Biodegrad. 129, 33-41. DOI: 10.1016/j.ibiod.2018.01.002
Coniglio, R., Gaschler, W., and Clavijo, L. (2023a). “Water-based system to prevent the yellowing of opaque coatings on knotted pine wood,” J. Coatings Technol. Res. 20(2), 781-788. DOI: 10.1007/s11998-022-00723-w
Cool, J., and Hernández, R. E. (2011). “Performance of three alternative surfacing processes on black spruce wood and their effects on water-based coating adhesion,” Wood Fiber Sci. 43(4), 365-378.
Cool, J., and Hernández, R. E. (2016). “Impact of three alternative surfacing processes on weathering performance of an exterior water-based coating,” Wood Fiber Sci. 48(1), 43-53.
Corcione, C. E., and Frigione, M. (2012). “UV-cured polymer-boehmite nanocomposite as protective coating for wood elements,” Prog. Organic Coat. 74(4), 781-787. DOI: 10.1016/j.porgcoat.2011.06.024
Cordeiro, L. A., de Miranda, B., Carneiro, M. E., Missio, A. L., Klock, U., and de Cademartori, P. H. G. (2023). “The effect of sanding on the wettability and surface quality of imbuia, red oak and pine wood veneers,” Maderas – Cienca y Technologia 25. DOI: 10.4067/s0718-221×2023000100421
Cox, R. M. (2003). Building an Industrial Wood Finish, Forest Products Society, Madison, WI.
Cristea, M. V., Riedl, B., and Blanchet, P. (2010). “Enhancing the performance of exterior waterborne coatings for wood by inorganic nanosized UV absorbers,” Prog. Organic Coat. 69(4), 432-441. DOI: 10.1016/j.porgcoat.2010.08.006
Czarniak, P., Borysiuk, P., El Bayda, H., Perisse, F., and Menecier, S. (2022). “Analysis of technological effects of cold plasma treatment of wood based materials finished with finishing foil and paints,” Int. J. Adhes. Adhesives 118, article 103248. DOI: 10.1016/j.ijadhadh.2022.103248
Darmawan, W., Nandika, D., Noviyanti, E., Alipraja, I., Lumongga, D., Gardner, D., and Gérardin, P. (2018). “Wettability and bonding quality of exterior coatings on jabon and sengon wood surfaces,” J. Coatings Technol Res. 15(1), 95-104. DOI: 10.1007/s11998-017-9954-1
Decker, C. (2005). “New developments in UV radiation curing of protective coatings,” Surface Coat. Int. Part B – Coat. Trans. 88(1), 9-17. DOI: 10.1007/BF02699702
de Meijer, M., Thurich, K., and Militz, H. (1998). “Comparative study on penetration characteristics of modern wood coatings,” Wood Sci. Technol. 32(5), 347-365. DOI: 10.1007/BF00702791
de Meijer, M., and Militz, H. (2000a). “Wet adhesion of low-VOC coatings on wood – A quantitative analysis,” Prog. Organic Coat. 38(3-4), 223-240. DOI: 10.1016/S0300-9440(00)00108-9
de Meijer, M., and Militz, H. (2000b). “Moisture transport in coated wood. Part 1: Analysis of sorption rates and moisture content profiles in spruce during liquid water uptake,” Holz als Roh- und Werkstoff 58(5), 354-362. DOI: 10.1007/s001070050445
de Meijer, M. (2001). “Review on the durability of exterior wood coatings with reduced VOC-content,” Prog. Organic Coatings 43(4), 217-225. DOI: 10.1016/S0300-9440(01)00170-9
de Meijer, M., and Militz, H. (2001). “Moisture transport in coated wood. Part 2: Influence of coating type, film thickness, wood species, temperature and moisture gradient on kinetics of sorption and dimensional change,” Holz als Roh- und Werkstoff 58(6), 467-475. DOI: 10.1007/s001070050461
de Meijer, M., Thurich, K., and Militz, H. (2001a). “Quantitative measurements of capillary coating penetration in relation to wood and coating properties,” Holz als Roh- und Werkstoff 59(1-2), 35-45. DOI: 10.1007/s001070050469
de Meijer, M., van de Velde, B., and Militz, H. (2001b). “Rheological approach to the capillary penetration of coating into wood,” J. Coatings Technol. 73(914), 39-51. DOI: 10.1007/BF02698437
de Moura, L. F., and Hernández, R. E. (2005). “Evaluation of varnish coating performance for two surfacing methods on sugar maple wood,” Wood Fiber Sci. 37(2), 355-366.
Dilik, T., Erdinler, S., Hazir, E., Koç, H., and Hiziroglu, S. (2015). “Adhesion strength of wood based composites coated with cellulosic and polyurethane paints,” Adv. Mater. Sci. Eng. 2015, article 745675. DOI: 10.1155/2015/745675
Dixit, A., Wazarkar, K., and Sabnis, A. S. (2021). “Antimicrobial UV curable wood coatings based on citric acid,” Pigment Resin Technol. 50(6), 533-544. DOI: 10.1108/PRT-07-2020-0067
Donkers, P. A. J., Huinink, H. P., Erich, S. J. F., Reuvers, N. J. W., and Adan, O. C. G. (2013). “Water permeability of pigmented waterborne coatings,” Prog. Organic Coat. 76(1), 60-69. DOI: 10.1016/j.porgcoat.2012.08.011
Dossin, R., Lavoratti, A., Cercena, R., Dal-bó, A. G., Zimmermann, M. V. G., and Zattera, A. J. (2022). “Influence of different photoinitiatiors in the UV-LED curing of polyester acrylated varnishes,” J. Coat. Technol. Res. 20(1), 393-402. DOI: 10.1007/s11998-022-00679-x
Dvorak, O., Kvietkova, M. S., Horak, P., Kubista, K., Panek, M., and Sterbova, I. (2023). “Effect of larch wood extractive leaching on accelerated weathering aging durability of oil-based coatings,” Cent. Eur. For. J. 69(2), 126-131. DOI: 10.2478/forj-2022-0018
Dvorchak, M. J. (1997). “Using ‘high performance two-component waterborne polyurethane’ wood coatings,” J. Coatings Technol. 69(866), 47-52. DOI: 10.1007/BF02696145
Ekstedt, J. (2003). “Influence of coating system composition on moisture dynamic performance of coated wood,” J. Coatings Technol. 75(938), 27-37. DOI: 10.1007/BF02697719
Englund, F., Bryne, L. E., Ernstsson, M., Lausmaa, J., and Wålinder, M. (2009). “Spectroscopic studies of surface chemical composition and wettability of modified wood,” Wood Mater. Sci. Eng. 4(1–2), 80-85. DOI: 10.1080/17480270903337659
Erich, S. J. F., Adan, O. C. G., Pel, L., Huinink, H. P., and Kopinga, K. (2006). “NMR imaging of coatings on porous substrates,” Chem. Mater. 18(18), 4500-4504. DOI: 10.1021/cm060744x
Ersoy, O., Fidan, S., Köse, H., Guler, D., and Özdöver, Ö. (2021). “Effect of calcium carbonate particle size on the scratch resistance of rapid alkyd-based wood coatings,” Coatings 11(3), article 340. DOI: 10.3390/coatings11030340
Evans, P. D., Haase, J. G., Shakri, B. M. S. A., and Kiguchi, M. (2015). “The search for durable exterior clear coatings for wood,” Coatings 5(4), 830-864. DOI: 10.3390/coatings5040830
Eyann, L., Ariff, Z, M., Shafiq, M. D., Shuib, R. K., and Musa, M. S. (2023). “Investigating the effect of methacrylic acid on the properties of waterborne epoxy-acrylate core-shell emulsion, film and coating,” Prog. Organic Coatings 187, article 108096. DOI: 10.1016/j.porgcoat.2023.108096
Fallah, F., Khorasani, M., and Ebrahimi, M. (2017). “Improving the mechanical properties of waterborne nitrocellulose coating using nano-silica particles,” Prog. Organic Coatings 109, 110-116. DOI: 10.1016/j.porgcoat.2017.04.016
Feist, W. C. (1982). Weathering of Wood in Structural Uses, USDA Forest Service, Forest Products Laboratory, Madison, WI.
Feist, W. C., and Mraz, E. A. (1980). Durability of Exterior Natural Wood Finishes in the Pacific Northwest, Forest Products Laboratory, Madison, WI. DOI: 10.2737/FPL-RP-366
Flexner, B. (1994). Understanding Wood Finishing: How to Select and Apply the Right Finish, Rodale Press, New York.
Flexner, B. (2005). Understanding Wood Finishing: How to Select and Apply the Right Finish, Readers Digest, Pleasantville, NY.
Fortmann, R., Roache, N., Chang, J. C. S., and Guo, Z. (1998). “Characterization of emissions of volatile organic compounds from interior alkyd paint,” J. Air Waste Manag. Assoc. 48(10), 931-940. DOI: 10.1080/10473289.1998.10463741
Fumero, A., and García, T. (2005). “Evaluation of the factors that affect the drying of an alkyd base varnish,” Ingenieria UC 12(1), 17-28.
Ganguli, P., and Chaudhuri, S. (2021). “Nanomaterials in antimicrobial paints and coatings to prevent biodegradation of man-made surfaces: A review,” Materials Today – Proc. 45, 3769-3777. DOI: 10.1016/j.matpr.2021.01.275
Gezici-Koç, Ö., Erich, S. J. F., Huinink, H. P., van der Ven, L. G. J., and Adan, O. C. G. (2019). “Moisture content of the coating determines the water permeability as measured by 1D magnetic resonance imaging,” Prog. Organic Coatings 130, 114-123. DOI: 10.1016/j.porgcoat.2019.01.043
Ghanbarian, B., Hunt, A. G., Ewing, R. P., and Sahimi, M. (2013). “Tortuosity in porous media: A critical review,” Soil Sci. Soc. Amer. J. 77(5), 1461-1477. DOI: 10.2136/sssaj2012.0435
Ghofrani, M., Mirkhandouzi, F. Z., and Ashori, A. (2016). “Effects of extractives removal on the performance of clear varnish coatings on boards,” J. Composite Mater. 50(21), 3019-3024. DOI: 10.1177/0021998315615205
Gholamiyan, H., Tarmian, A., Ranjbar, Z., Abdulkhani, A., Azadfallah, M., and Mai, C. (2016). “Silane nanofilm formation by sol-gel processes for promoting adhesion of waterborne and solvent-borne coatings to wood surface,” Holzforschung 70(5), 429-437. DOI: 10.1515/hf-2015-0072
Gibbons, M. J., Nikafshar, S., Saravi, T., Ohno, K., Chandra, S., and Nejad, M. (2020). “Analysis of a wide range of commercial exterior wood coatings,” Coatings 10(11), article 1013. DOI: 10.3390/coatings10111013
Godinho, B., Gama, N., Barros-Timmons, A., and Ferreira, A. (2021). “Recycling of different types of polyurethane foam wastes via acidolysis to produce polyurethane coatings,” Sustain. Mater. Technol. 29, article e00330. DOI: 10.1016/j.susmat.2021.e00330
Godnjavec, J., Znoj, B., Veronovski, N., and Venturini, P. (2012). “Polyhedral oligomeric silsesquioxanes as titanium dioxide surface modifiers for transparent acrylic UV blocking hybrid coating,” Prog. Organic Coat. 74(4), 654-659. DOI: 10.1016/j.porgcoat.2011.09.032
Gomez-Lopez, A., Panchireddy, S., Grignard, B., Calvo, I., Jerome, C., Detrembleur, C., and Sardon, H. (2021). “Poly(hydroxyurethane) adhesives and coatings: State-of-the-art and future directions,” ACS Sustain. Chem. Eng. 9(29), 9541-9562. DOI: 10.1021/acssuschemeng.1c02558
Gong, X. B., Davis, H. T., and Scriven, L. E. (2008). “Role of van der Waals force in latex film formation,” J. Coatings Technol. Res. 5(3), 271-283. DOI: 10.1007/s11998-008-9095-7
Goodell, B. (2003). “Brown-rot fungal degradation of wood: Our evolving view,” in: Wood Deterioration and Preservation, B. Goodell, D. D. Nicholas, and T. P. Schultz (eds.), Vol. 845, pp. 97-118. DOI: 10.1021/bk-2003-0845.ch006
Grigsby, W. J. (2018). “Photooxidative stability provided by condensed tannin additives in acrylic-based surface coatings on exterior exposure,” J. Coatings Technol. Res. 15(6), 1273-1282. DOI: 10.1007/s11998-018-0086-z
Grigsby, W., and Steward, D. (2018). “Applying the protective role of condensed tannins to acrylic-based surface coatings exposed to accelerated weathering,” J. Polym. Environ. 26(3), 895-905. DOI: 10.1007/s10924-017-0999-0
Grüll, G., Tscherne, F., Spitaler, I., and Forsthuber, B. (2014). “Comparison of wood coating durability in natural weathering and artificial weathering using fluorescent UV-lamps and water,” Eur. J. Wood Wood Prods. 72(3), 367-376. DOI: 10.1007/s00107-014-0791-y
Grüneberger, F., Künniger, T., Huch, A., Zimmermann, T., and Arnold, M. (2015). “Nanofibrillated cellulose in wood coatings: Dispersion and stabilization of ZnO as UV absorber,” Prog. Organic Coatings 87, 112-121. DOI: 10.1016/j.porgcoat.2015.05.025
Grüneberger, F., Künniger, T., Zimmermann, T., and Arnold, M. (2014). “Nanofibrillated cellulose in wood coatings: Mechanical properties of free composite films,” J. Mater. Sci. 49(18), 6437-6448. DOI: 10.1007/s10853-014-8373-2
Gunduz, A., Baysal, E., Turkoglu, T., Kucuktuvek, M., Altay, C., Peker, H., and Toker, H. (2019). “Accelerated weathering performance of Scots pine preimpregnated with copper based chemicals before varnish coating. Part II: Coated with water based varnish,” Wood Res. 64(6), 987-998.
Gupta, P. C., Islam, M., and Prasad, R. (1983). “The adhesive property of shellac on wood to wood surfaces,” J. Oil Colour Chem. Assoc. 66(5), 141-142.
Gurleyen, L. (2021). “Effects of artificial weathering on the color, gloss, adhesion, and pendulum hardness of UV system parquet varnish applied to doussie (Afzelia africana) wood,” BioResources 16(1), 1616-1627. DOI: 10.15376/biores.16.1.1616-1627
Gustafsson, L. M., and Börjesson, P. (2007). “Life cycle assessment in green chemistry: A comparison of various industrial wood surface coatings,” Int. J. Life Cycle Assess. 12(3), 151-159. DOI: 10.1065/lca2006.11.280
Haase, J. G., Leung, L. H., and Evans, P. D. (2019). “Plasma pre-treatments to improve the weather resistance of polyurethane coatings on black spruce wood,” Coatings 9(1), article 8. DOI: 10.3390/coatings9010008
Hansen, C. M. (2007). Hansen Solubility Parameters: A User’s Handbook, 2nd Ed., CRC Press, Boca Raton, FL. DOI: 10.1201/9781420006834
Hazir, E. (2021). “Improvement of heat-treated wood coating performance using atmospheric plasma treatment and design of experiments method,” Polymers 13(9), article 1520. DOI: 10.3390/polym13091520
Hazir, E., and Koc, K. H. (2019). “Evaluation of wood surface coating performance using water based, solvent based and powder coating,” Maderas – Ciencia y Technologia 21(4), 467-480. DOI: 10.4067/S0718-221X2019005000404
Hazir, E., Seker, S., Koc, K. H., Dilik, T., Erdinler, E. S., and Ozturk, E. (2021). “Optimization of plasma treatment parameters to improve the wood-coating adhesion strength using Taguchi integrated desirability function approach,” J. Adhes. Sci. Technol. 35(5), 451-467. DOI: 10.1080/01694243.2020.1816668
He, Y. R., Wu, Y. Z., Qu, W., and Zhang, J. P. (2022). “Crack-resistant amino resin flame-retardant coatings using waterborne polyurethane as a co-binder resin,” Mater. 15(12), article 4122. DOI: 10.3390/ma15124122
Hermann, A., Giljean, S., Pac, M. J., Marsiquet, C., Beaufils-Marquet, M., Burr, D., and Landry, V. (2021). “Understanding indentation, scratch and wear behavior of UV-cured wood finishing products,” Prog. Oran. Coatings 161, article 106504. DOI: 10.1016/j.porgcoat.2021.106504
Hermann, A., Giljean, S., Pac, M. J., Marsiquet, C., Burr, D., and Landry, V. (2023). “Physico-mechanical characterisation of basecoats for tailored UV-cured multilayered wood coating systems,” Prog. Organic Coatings 182, article 107673. DOI: 10.1016/j.porgcoat.2023.107673
Herrera, R., Muszynska, M., Krystofiak, T., and Labidi, J. (2015). “Comparative evaluation of different thermally modified wood samples finishing with UV-curable and waterborne coatings,” Appl. Surf. Sci. 357, 1444-1453. DOI: 10.1016/j.apsusc.2015.09.259
Herrera, R., Sandak, J., Robles, E., Krystofiak, T., and Labidi, J. (2018). “Weathering resistance of thermally modified wood finished with coatings of diverse formulations,” Prog. Organic Coatings 119, 145-154. DOI: 10.1016/j.porgcoat.2018.02.015
Hiltunen, E., Mononen, K., Alvila, L., and Pakkanen, T. T. (2008). “Discolouration of birch wood: Analysis of extractives from discoloured surface of vacuum-dried European white birch (Betula pubescens) board,” Wood Sci. Technol. 42(2), 103-115. DOI: 10.1007/s00226-007-0143-8
Howard, E. M., McCrillis, R. C., Krebs, K. A., Fortman, R., Lao, H. C., and Guo, Z. (1998). “Indoor emissions from conversion varnishes,” J. Air Waste Manag. Assoc. 48(10), 924-930. DOI: 10.1080/10473289.1998.10463747
Huang, Y. H., Feng, Q. M., Ye, C. Y., Nair, S. S., and Yan, N. (2019). “Incorporation of ligno-cellulose nanofibrils and bark extractives in water-based coatings for improved wood protection,” Prog. Organic Coatings 138, article 105210. DOI: 10.1016/j.porgcoat.2019.105210
Huang, Y. H., Ma, T. T., Li, L. P., Wang, Q. W., and Guo, C. G. (2022). “Facile synthesis and construction of renewable, waterborne and flame-retardant UV-curable coatings in wood surface,” Prog. Organic Coatings 172, article 107104. DOI: 10.1016/j.porcoat.2022.107104
Huang, Y. S., Zhou, Q. F., Li, L. P., Wang, Q. W., and Guo, C. G. (2023). “Construction of waterborne flame-retardant itaconate-based unsaturated polyesters and application for UV-curable hybrid coatings on wood,” Prog. Organic Coat. 183, article 107826. DOI: 10.1016/j.porgcoat.2023.107826
Hubbe, M. A. (2023). “Sustainable composites: A review with critical questions to guide future initiatives,” Sustainability 15, article 11088. DOI: 10.3390/su151411088
Hubbe, M. A., Gardner, D. J., and Shen, W. (2015a). “Contact angles and wettability of cellulosic surfaces: A review of proposed mechanisms and test strategies,” BioResources 10(4), 8657-8749. DOI: 10.15376/biores.10.4.Hubbe_Gardner_Shen
Hubbe, M. A., and Grigsby, W. (2020). “From nanocellulose to wood particles: A review of particle size vs. the properties of plastic composites reinforced with cellulose-based entities,” BioResources 15(1), 2030-2081. DOI: 10.15376/biores.15.1.2030-2081
Hubbe, M. A., Rojas, O. J., and Lucia, L. A. (2015b). “Green modification of surface characteristics of cellulosic materials at the molecular or nano scale: A review,” BioResources 10(3), 6095-6229. DOI: 10.15376/biores.10.3.Hubbe
Hubbe, M. A., Sjöstrand, B., Lestelius, M., Håkansson, H., Swerin, A., and Henriksson, G. (2024). “Swelling of cellulosic fibers in aqueous systems: A review of chemical and mechanistic factors,” BioResources 19(3), 6859-6945. DOI: 10.15376/biores.19.3.Hubbe
Hubbe, M. A., Tayeb, P., Joyce, M., Tyagi, P., Kehoe, M., Dimic-Misic, K., and Pal, L. (2017). “Rheology of nanocellulose-rich aqueous suspensions: A review,” BioResources 12(4), 9556-9661. DOI: 10.15376/biores.12.1.2143-2233
Hubbe, M. A., Trovagunta, R., Zambrano, F., Tiller, P., and Jardim, J. (2023). “Self-assembly fundamentals in the reconstruction of lignocellulosic materials: A review,” BioResources 18(2), 4262-4331. DOI: 10.15376/biores.18.2.Hubbe
Hwang, H. D., Moon, J. I., Choi, J. H., Kim, H. J., Do Kim, S., and Park, J. C. (2009). “Effect of water drying conditions on the surface property and morphology of waterborne UV-curable coatings for engineered flooring,” J. Ind. Eng. Chem. 15(3), 381-387. DOI: 10.1016/j.jiec.2008.11.002
Hysek, S., Fidan, H., Pánek, M., Böhm, M., and Trgala, K. (2018). “Water permeability of exterior wood coatings: Waterborne acrylate dispersions for windows,” J. Green Building 13(3), 3-16. DOI: 10.3992/1943-4618.13.3.1
Ibrahim, M. S., Mohamed, H. A., Moustafa, M. E., Mohamed, T. Y., and Ashraf, A. (2020). “A comparative study of ultraviolet and electron beam irradiation on acrylate coatings,” Egyptian J. Chem. 63(5), 1931-1940. DOI: 10.21608/ejchem.2019.15990.1966
Irmouli, Y., George, B., and Merlin, A. (2012). “Artificial ageing of wood finishes monitored by IR analysis and color measurements,” J. Appl. Polym. Sci. 124(3), 1938-1946. DOI: 10.1002/app.34797
Jamali, A., and Evans, P. D. (2020). “Plasma treatment reduced the discoloration of an acrylic coating on hot-oil modified wood exposed to natural weathering,” Coatings 10(3), article 248. DOI: 10.3390/coatings10030248
Jämsä, S., Ahola, P., and Viitaniemi, P. (1999). “Performance of coated heat-treated wood,” Surf. Coat. Int. 82(6), 297-300. DOI: 10.1007/BF02720126
Jordan, J., Helander, R., and Laaksonen, P. (2021). “Colour stability of wood coatings pigmented with natural indigo from Isatis tinctoria after accelerated weathering,” Coloration Technol. 138(2), 210-218. DOI: 10.1111/cote.12585
Jung, S. J., Lee, S. J., Cho, W. J., and Ha, C. S. (1998). “Synthesis and properties of UV-curable waterborne unsaturated polyester for wood coating,” J. Appl. Polym. Sci. 69(4), 695-708. DOI: 10.1002/(SICI)1097-4628(19980725)69:4<695::AID-APP8>3.3.CO;2-F
Kanbayashi, T., Ishikawa, A., Matsunaga, M., Kobayashi, M., and Kataoka, Y. (2019). “Application of confocal Raman microscopy for the analysis of the distribution of wood preservative coatings,” Coatings 9(10), article 621. DOI: 10.3390/coatings9100621
Karaman, I., Kiliç, K., and Sögütlü, C. (2023). “Prediction of adhesion strength of some varnishes using soft computing models,” Drvna Industrija 74(2), 153-166. DOI: 10.5552/drvind.2023.0029
Karamanoglu, M., Birinci, E., Kesik, H. I., and Kaymakci, A. (2022). “Effect of treatment temperature on the initial performance of layers of water-based paints in heat-treated pine and beech wood,” BioResources 17(1), 1491-1506. DOI: 10.15376/biores.17.1.1494-1506
Karbalaei, H., Tarmian, A., Rasouli, D., and Pourmahdian, S. (2022). “Effects of UV-curing epoxy acrylate and urethane acrylate coatings incorporated with ZnO nanoparticles on weathering resistance of thermally modified timber,” Wood Mater. Sci. Eng. 17(6), 868-877. DOI: 10.1080/17480272.2021.1968491
Kardar, P., Ebrahimi, M., and Bastani, S. (2014). “Curing behaviour and mechanical properties of pigmented UV-curable epoxy acrylate coatings,” Pigment Resin Technol. 43(4), 177-184. DOI: 10.1108/PRT-07-2013-0054
Kelkar, B. U., Shukla, S. R., Yadav, S. M., and Bansal, R. (2023). “Performance of laminated bamboo lumber and bamboo strand lumber coated with solvent and water-based polyurethane against accelerated UV and natural weathering,” Indust. Crops Prod. 192, article 116058. DOI: 10.1016/j.indcrop.2022.116058
Kesik, H. I., and Akyildiz, M. H. (2015). “Effect of the heat treatment on the adhesion strength of water based wood varnishes,” Wood Res. 60(6), 987-994.
Kettner, F., Plaschkies, K., Gerullis, S., Pfuch, A., and Küzün, B. (2020). “Raising the permanent adhesion of coatings on resinous wood by APPCVD promotion,” Int. J. Adhesion Adhesives 102, article 102642. DOI: 10.1016/j.ijadhadh.2020.102642
Kibleur, P., Blykers, B., Boone, M. N., Van Hoorebeke, L., Van Acker, J., and van den Bulcke, J. (2022). “Detecting thin adhesive coatings in wood fiber materials with laboratory-based dual-energy computed tomography (DECT),” Sci. Rep. 12(1), article 15969. DOI: 10.1038/s41598-022-20422-1
Kielmann, B. C., and Mai, C. (2016a). “Application and artificial weathering performance of translucent coatings on resin-treated and dye-stained beech-wood,” Prog. Organic Coatings 95, 54-63. DOI: 10.1016/j.porgcoat.2016.02.019
Kielmann, B. C., and Mai, C. (2016b). “Natural weathering performance and the effect of light stabilizers in water-based coating formulations on resin-modified and dye-stained beech-wood,” J. Coatings Technol. Res. 13(6), 165-1074. DOI: 10.1007/s11998-016-9818-0
Kiliç, K., and Sögütlü, C. (2022). “Varnish adhesion strength in natural aged wood material,” J. Polytech. Politeknik Dergisi 26(2), 847-854. DOI: 10.2339/politeknik.1072314
Kimerling, A. S., and Bhatia, S. R. (2004). “Block copolymers as low-VOC coatings for wood: Characterization and tannin bleed resistance,” Prog. Organic Coat. 51(1), 15-26. DOI: 10.1016/j.porgcoat.2004.03.006
Klatt, J., Barcellona, P., Bennett, R., Bokareva, O. S., Feth, H., Rasch, A., Reith, P., and Buhnnann, S. Y. (2017). “Strong van der Waals adhesion of a polymer film on rough substrates,” Langmuir 33(21), 5298-5303. DOI: 10.1021/acs.langmuir.7b01381
Kleive, K. (1986). “Weathered wooden surfaces – Their influence on the durability of coating systems,” J. Coatings Technol. 740, 39-43.
Klinc, M., Pavlic, M., Petric, M., and Pohleven, F. (2017). “Influence of microwave heating in wood preservation on traditional surface coatings,” Acta Silvae et Ligni 122, 21-33. DOI: 10.20315/ASetL.112.3
Knaebe, M., and Williams, R. S. (1993). “Determining paint adhesion to wood using a uniform double-cantilever beam technique,” J. Testing Eval. 21(4), 272-279. DOI: 10.1520/JTE11952J
Köhler, R., Sauerbier, P., Militz, H., and Viöl, W. (2017). “Atmospheric pressure plasma coating of wood and MDF with polyester powder,” Coatings 7(10), article 171. DOI: 10.3390/coatings7100171
Kristyna, S., Stepán, H., Eliska, O., Milos, P., and Hakan, F. (2020). “Effect of artificial weathering and temperature cycling on the adhesion strength of waterborne acrylate coating systems used for wooden windows,” J. Green Building 15(1), 3-14. DOI: 10.3992/1943-4618.15.1.1
Krongauz, V. V. (2021). “Kinetics and mechanism of plasticized poly(vinyl chloride) films autohesion interface effect,” J. Thermal Anal. Calorim. 147(6), 4177-4195. DOI: 10.1007/s10973-021-10832-0
Kropat, M., Hubbe, M. A., and Laleicke, F. (2020). “Natural, accelerated, and simulated weathering of wood: A review,” BioResources 15(4), 9998-10062. DOI: 10.15376/biores.15.4.Kropat
Krystofiak, T., Lis, B., Muszynska, M., and Proszyk, S. (2016). “The effect of aging tests on gloss and adhesion of lacquer coatings on window elements from pine wood,” Drewno 59(197), 127-137. DOI: 10.12841/wood.1644-3985.C33.19
Kubelka, P., and Munk, F. (1931). “An article on optics of paint layers,” Z. Tech. Phys. 12(593-601), 259-274.
Kúdela, J., and Liptáková, E. (2006). “Adhesion of coating materials to wood,” J. Adhesion Sci. Technol. 20(8), 875-895. DOI: 10.1163/156856106777638725
Kulichikhin, V. G., and Malkin, A. Y. (2022). “The role of structure in polymer rheology: Review,” Polymers 14(6), article 1262. DOI: 10.3390/polym14061262
Lagana, R., Svocák, J., and Kúdela, J. (2021). “A study of the surface properties of a solid wood coating system with self-healing microcapsules,” in: 9th Hardwood Proceedings, Vol. 9 – Pt II: An Underutilized Resource – Hardwood Oriented Research, R. Nemeth, et al. (eds.), Univ. Sopron Niensis, pp. 61-67.
Landry, V., and Blanchet, P. (2012). “Surface preparation of wood for application of waterborne coatings,” Forest Prod. J. 62(1), 39-45. DOI: 10.13073/FPJ-D-10-00011.1
Landry, V., Blanchet, P., Boivin, G., Bouffard, J. F., and Vlad, M. (2015). “UV-LED curing efficiency of wood coatings,” Coatings 5(4), 1019-1033. DOI: 10.3390/coatings5041019
Landry, V., Blanchet, P., and Cormier, L. M. (2013). “Water-based and solvent-based stains: Impact on the grain raising in yellow birch,” BioResources 8(2), 1997-2009. DOI: 10.15376/biores.8.2.1997-2009
Landry, V., Blanchet, P., and Riedl, B. (2010). “Mechanical and optical properties of clay-based nanocomposites coatings for wood flooring,” Prog. Organic Coat. 67(4), 381-388. DOI: 10.1016/j.porgcoat.2009.12.011
Landry, V., Boivin, G., Schorr, D., Mottoul, M., Mary, A., Abid, L., Carrère, M., and Laratte, B. (2023). “Recent developments and trends in sustainable and functional wood coatings,” Current Forestry Reports 9(5), 319-331. DOI: 10.1007/s40725-023-00195-0
Landry, V., Riedl, B., and Blanchet, P. (2008a). “Alumina and zirconia acrylate nanocomposites coatings for wood flooring: Photocalorimetric characterization,” Prog. Organic Coat. 61(1), 76-82. DOI: 10.1016/j.porgcoat.2007.09.013
Landry, V., Riedl, B., and Blanchet, P. (2008b). “Nanoclay dispersion effects on UV coatings curing,” Prog. Organic Coat. 62(4), 400-408. DOI: 10.1016/j.porgcoat.2008.02.010
Li, X., Wang, D., Zhao, L. Y., Hou, X. Z., Liu, L., Feng, B., Li, M. X., Zheng, P., Zhao, X., and Wei, S. Y. (2021). “UV LED curable epoxy soybean-oil-based waterborne PUA resin for wood coatings,” Prog. Organic Coat. 151, article 105942. DOI: 10.1016/j.porgcoat.2020.105942
Liu, Y., Hu, J., and Wu, Z. H. (2020). “Fabrication of coatings with structural color on a wood surface,” Coatings 10(1), article 32. DOI: 10.3390/coatings10010032
Lokhande, G., Chambhare, S., and Jagtap, R. (2017). “Synthesis and properties of phosphate-based diacrylate reactive diluent applied to UV-curable flame-retardant wood coating,” J. Coating Technol. Res. 14(1), 255-266. DOI: 10.1007/s11998-016-9849-6
Lu, K. T., and Chang, J. P. (2020). “Synthesis and antimicrobial activity of metal-containing linseed oil-based waterborne urethane oil wood coatings,” Polymers 12(3), article 663. DOI: 10.3390/polym12030663
Lu, K. T., and Lin, S. L. (2008). “Synthesis of polyurethane resins and its improvement on dimensional stability and finishing performances of medium- and small-diameter softwoods,” J. Appl. Polym. Sci. 108(3), 2029-2036. DOI: 10.1002/app.27169
Lu, K. T., Liu, C. T., and Lin, S. M. (2004). “The effect of alkyd resins on the properties of AA-NC semi-IPNs as binders in wood finish,” J. Appl. Polym. Sci. 93(4), 1923-1927. DOI: 10.1002/app.20651
Lucas, R. (1918). “Ueber das Zeitgesetz des kapillaren Aufstiegs von Flüssigkeiten,” Kolloid Zeitschrift 23(1), 15-22. DOI: 10.1007/BF01461107
Mahajan, D., Srivats, D. S., and More, A. (2023). “Synthesis of vanillin-based UV curable polyurethane dispersions for wood coating applications,” J. Coatings Technol. Res. Early Access. DOI: 10.1007/s11998-023-00780-9
Mali, P., Sonawane, N. S., Patil, V., Lokhande, G., Mawale, R., and Pawar, N. (2021). “Morphology of wood degradation and flame retardants wood coating technology: An overview,” Int. Wood Prod. J. 13(1), 21-40. DOI: 10.1080/20426445.2021.2011552
Marschner, S. R., Westin, S. H., Arbree, A., and Moon, J. T. (2005). “Measuring and modeling the appearance of finished wood,” ACM Trans. Graphics 24(3), 727-734. DOI: 10.1145/1073204.1073254
McCrillis, R. C., Howard, E. M., Guo, Z. S., Krebs, K. A., Fortmann, R., and Lao, H. C. (1999). “Characterization of curing emissions from conversion varnishes,” J. Air Waste Manag. Assoc. 49(1), 70-75. DOI: 10.1080/10473289.1999.10463777
Miccichè, F., van Haveren, J., Oostveen, E., Ming, W., and van der Linde, R. (2006). “Oxidation and oligomerization of ethyl linoleate under the influence of the combination of ascorbic acid 6-palmitate/iron-2-ethylhexanoate,” Appl. Catal. Gen. 297, 174-181. DOI: 10.1016/j.apcata.2005.09.008
Mihaila, A., Danu, M., Ibanescu, C., Anghel, I., Sofran, I. E., Balanescu, L. V., Tudorachi, N., and Lisa, G. (2022). “Thermal characterization and rheological behavior of some varnishes and paints used for wood protection,” Int. J. Environ. Sci. Technol. 19(7), 6299-6314. DOI: 10.1007/s13762-021-03579-6
Mihaila, A., Lisa, C., Ipate, A. M., Zaltariov, M. F., Rusu, D., Mamaliga, I., and Lisa, G. (2019a). “Determination of the effective diffusion coefficient during the drying of paint and varnish films applied on fir wood,” Prog. Oran. Coatings 137, article 105344. DOI: 10.1016/j.porgcoat.2019.105344
Mihaila, A., Lisa, C., Mamaliga, I., and Lisa, G. (2019b). “Kinetics of drying of certain lacquers and paints in isothermal conditions using a thermogravimetric analyser,” J. Thermal Anal. Calorim. 138(3), 2315-2322. DOI: 10.1007/s10973-019-08779-4
Mikkonen, K. S., Kirjoranta, S., Xu, C. L., Hemming, J., Pranovich, A., Bhattarai, M., Peltonen, L., Kilpeläinen, P., Maina, N., and Tenkanen, M. (2019). “Environmentally-compatible alkyd paints stabilized by wood hemicelluloses,” Indust. Crops. Prod. 133, 212-220. DOI: 10.1016/j.indcrop.2019.03.017
Miklecic, J., Loncaric, A., Veselicic, N., and Jirous-Rajkovic, V. (2022). “Influence of wood surface preparation on roughness, wettability and coating adhesion of unmodified and thermally modified wood,” Drvna Industrija 73(3), 261-269. DOI: 10.5552/drvind.2022.0016
Montazeri, M., and Eckelman, M. J. (2018). “Life cycle assessment of UV-curable bio-based wood flooring coatings,” J. Cleaner Prod. 192, 932-939. DOI: 10.1016/j.jclepro.2018.04.209
Motawie, A. M., and Sadek, E. M. (1999). “Adhesives and coatings based on phenolic epoxy resins,” Polym. Advan. Technol. 10(4), 223-228. DOI: 10.1002/(SICI)1099-1581(199904)10:4<223::AID-PAT865>3.0.CO;2-7
Mudri, N. H., Abdullah, L. C., Aung, M. M., Salleh, M. Z., Biak, D. R. A., and Rayung, M. (2020). “Comparative study of aromatic and cycloaliphatic isocyanate effects on physico-chemical properties of bio-based polyurethane acrylate coatings,” Polymers 12(7), article 1494. DOI: 10.3390/polym12071494
Napper, D. H. (1977). “Steric stabilization,” J. Colloid Interface Sci. 58(2), 390-407. DOI: 10.1016/0021-9797(77)90150-3
Nazari, B., Kumar, V., Bousfield, D. W., and Toivakka, M. (2016). “Rheology of cellulose nanofibers suspensions: Boundary driven flow,” J. Rheology 60(6), 1151-1159. DOI: 10.1122/1.4960336
Nejad, M., and Cooper, P. (2010). “Coatings to reduce wood preservative leaching,” Environ. Sci. Technol. 44(16), 6162-6166. DOI: 10.1021/es101138v
Nejad, M., and Cooper, P. (2011). “Exterior wood coatings. Part-2: Modeling correlation between coating properties and their weathering performance,” J. Coatings Technol. Res. 8(4), 459-467. DOI: 10.1007/s11998-011-9331-4
Nejad, M., and Cooper, P. (2017). “Exterior wood coatings,” in: Wood in Civil Engineering, pp. 111-129. DOI: 10.5772/67170
Nussbaum, R. M., Sutcliffe, E. J., and Hellgren, A. C. (1998). “Microautoradiographic studies of the penetration of alkyd, alkyd emulsion and linseed oil coatings into wood,” J. Coatings Technol. 70(878), 49-57. DOI: 10.1007/BF02697811
Orlova, Y., Harmon, R. E., Broadbelt, L. J., and Iedema, P. D. (2021). “Review of the kinetics and simulations of linseed oil autoxidation,” Prog. Org. Coatings 151, article e106041. DOI: 10.1016/j.porgcoat.2020.106041
Pan, P., and Yan, X. X. (2023). “Preparation of antibacterial nanosilver solution microcapsules and their impact on the performance of andoung wood surface coating,” Polymers 15(7), article 1722. DOI: 10.3390/polym15071722
Pan, P., Yan, X. X., and Peng, W. W. (2022). “Tung oil microcapsules prepared with different emulsifiers and their effects on the properties of coating film,” Coatings 12(8), article 1166. DOI: 10.3390/coatings12081166
Pandey, K. K., and Srinivas, K. (2015). “Performance of polyurethane coatings on acetylated and benzoylated rubberwood,” Eur. J. Wood Wood Prod. 73(1), 111-120. DOI: 10.1007/s00107-014-0860-2
Paquet, C., Brown, S., Klemberg-Sapieha, J. E., Morin, J. F., and Landry, V. (2021). “Self-healing UV-curable acrylate coatings for wood finishing system, Part 2: Impact of monomer structure and self-healing parameters on self-healing efficiency,” Coatings 11(11), article 1328. DOI: 10.3390/coatings11111328
Paquet, C., Schmitt, T., Klemberg-Sapieha, J. E., Morin, J. F., and Landry, V. (2020). “Self-healing UV curable acrylate coatings for wood finishing system, Part 1: Impact of the formulation on self-healing efficiency,” Coatings 10(8), article 770. DOI: 10.3390/coatings10080770
Patil, A. M., and Jagtap, R. N. (2021). “PU-coating performance of bio-based hyperbranched alkyd resin on mild steel and wood substrate,” J. Coatings Technol. Res. 18(3), 741-752. DOI: 10.1007/s11998-020-00438-w
Pelit, H., Koc, E., and Cakicier, N. (2023). “Adhesion strength and pendulum hardness of some coatings in wood heat-treated by different methods,” BioResources 18(4), 7353-7366. DOI: 10.15376/biores.18.4.7353-7366
Peng, W. W., and Yan, X. X. (2022). “Preparation of tung oil microcapsule and its effect on wood surface coating,” Polymers 14(8), article 1536. DOI: 10.3390/polym14081536
Peng, X. R., and Zhang, Z. K. (2019). “Improvement of paint adhesion of environmentally friendly paint film on wood surface by plasma treatment,” Prog. Organic Coatings 134, 255-263. DOI: 10.1016/j.porgcoat.2019.04.024
Philipp, C., and Eschig, S. (2012). “Waterborne polyurethane wood coatings based on rapeseed fatty acid methyl esters,” Prog. Organic Coat. 74(4), 705-711. DOI: 10.1016/j.porgcoat.2011.09.028
Ploehn, H. J., and Liu, C. Y. (2006). “Quantitative analysis of montmorillonite platelet size by atomic force microscopy,” Indust. Eng. Chem. Res. 45(21), 7025-7034. DOI: 10.1021/ie051392r
Poaty, B., Vardanyan, V., Wilczak, L., Chauve, G., and Riedl, B. (2014). “Modification of cellulose nanocrystals as reinforcement derivatives for wood coatings,” Prog. Organic Coatings 77(4), 813-820. DOI: 10.1016/j.porgcoat.2014.01.009
Podgorski, L., and Roux, M. (1999). “Wood modification to improve the durability of coatings,” JOCCA – Surf. Coat. Int. 82(12), 590-596. DOI: 10.1007/BF02692672
Poussard, L., Lazko, J., Mariage, J., Raquez, J. M., and Dubois, P. (2016). “Biobased waterborne polyurethanes for coating applications: How fully biobased polyols may improve the coating properties,” Prog. Organic Coat. 97, 175-183. DOI: 10.1016/j.porgcoat.2016.04.003
Pratiwi, L. A., Darmawan, W., Priadi, T., George, B., Merlin, A., Gérardin, C., Dumarçay, S., and Gérardin, P. (2019). “Characterization of thermally modified short and long rotation teaks and the effects on coatings performance,” Maderas – Ciencia Tecnol. 21(2), 209-222. DOI: 10.4067/S0718-221X2019005000208
Probst, F., Laborie, M. P., Pizzi, A., Merlin, A., and Deglise, X. (1997). “Molecular mechanics experimental methods applied to varnish/primer/wood interactions,” Holzforschung 51(5), 459-466. DOI: 10.1515/hfsg.1997.51.5.459
Qin, Y. Z., and Yan, X. X. (2022). “Effect of the addition of shellac self-healing and discoloration microcapsules on the performance of coatings applied on ebiara solid board,” Coatings 12(11), article 1627. DOI: 10.3390/coatings12111627
Qu, L. J., Yang, J. H., Li, L. M., Zhu, W. K., Liu, L., and Li, S. (2023). “A clear coating with high hydrophobicity and anti-weathering performance by low-temperature siloxane deposition,” Polym. Degrad. Stabil. 214, article 110383. DOI: 10.1016/j.polymdegradstab.2023.110383
Quintos, A. L., Jimenez-Jr, J. P., Salvatierra, B. S., and Tagolabong, A. A. (2023). “Performance of film-forming finishes on thermally modified bolo (Gigantochloa levis),” J. Tropical For. Sci. 35(3), 339-349. DOI: 10.26525/jtfs2023.35.3.339
Ranarijaona, M. M., Rafanoela, S. H., Herinirina, L. C., Duclos, M. C., Lavaud, A. L., Goux-Henry, C., Métay, E., Vestalys, V. R., and Lemaire, M. (2023). “CNSL oxyacetic derivatives, new bio-based binder for paint preparation,” Green Mater. Early access. DOI: 10.1680/jgrma.23.00046
Rawat, R. S., Chouhan, N., Talwar, M., Diwan, R. K., and Tyagi, A. K. (2019). “UV coatings for wooden surfaces,” Prog. Organic Coatings 135, 490-495. DOI: 10.1016/j.porgcoat.2019.06.051
Raychura, A. J., Dholakiya, B. Z., Patel, K. I., and Jauhari, S. (2018). “Development of non-traditional vegetable-oil-based two-pack polyurethane for wood-finished coating: An alternative approach,” Chemistryselect 3(39), 10837-10842. DOI:10.1002/slct.201801452
Remadevi, O. K., Siddiqui, M. Z., Nagaveni, H. C., Rao, M. V., Shiny, K. S., and Ramani, R. (2015). “Efficacy of shellac-based varnishes for protection of wood against termite, borer and fungal attack,” J. Indian Acad. Wood Sci. 12(1), 9-14. DOI: 10.1007/s13196-015-0138-2
Ren, X. F., and Soucek, M. (2014). “Soya-based coatings and adhesives,” in: Soy-based Chemicals and Materials, R. P. Brenin (ed.), ACS Symp. Ser. 1178, Ch. 10, pp. 207-254. DOI: 10.1021/bk-2014-1178.ch010
Renneckar, S., and Zhou, Y. (2009). “Nanoscale coatings on wood: Polyelectrolyte adsorption and layer-by-layer assembled film formation,” ACS Appl. Mater. Interfaces 1(3), 559-566. DOI: 10.1021/am800119q
Riedl, B., Angel, C., Prégent, J., Blanchet, P., and Stafford, L. (2014). “Effect of wood surface modification by atmospheric-pressure plasma on waterborne coating adhesion,” BioResources 9(3), 4908-4923. DOI: 10.15376/biores.9.3.4908-4923
Rijckaert, V., Stevens, M., Van Acker, J., de Meijer, M., and Militz, H. (2001). “Quantitative assessment of the penetration of water-borne and solvent-borne wood coatings in Scots pine sapwood,” Holz als Roh- und Werkstoff 59(4), 278-287. DOI: 10.1007/s001070100208
Romani, F., Corrieri, R., Braga, V., and Ciardelli, F. (2002). “Monitoring the chemical crosslinking of propylene polymers through rheology,” Polymer 43(4), 1115-1131. DOI: 10.1016/S0032-3861(01)00679-6
Sabani, S., Önen, A. H., and Güngör, A. (2012). “Preparation of hyperbranched polyester polyol-based urethane acrylates and applications on UV-curable wood coatings,” J. Coatings Technol. Res. 9(6), 703-716. DOI: 10.1007/s11998-012-9418-6
Salleh, N. G. N., Alias, M. S., Gläsel, H. J., and Mehnert, R. (2013). “High performance radiation curable hybrid coatings,” Radiation Phys. Chem. 84, 70-73. DOI: 10.1016/j.radphyschem.2012.06.042
Sarma, B., Dolui, S. K., and Sharma, A. K. (2001). “Comparative evaluation of the effect of butylated and isobutylated derivatives of M-F resins on wood-finish properties,” J. Sci. Indust. Res. 60(2), 153-158.
Scharff, R. (1974). Complete Book of Wood Finishing, 2nd Ed., McGraw-Hill, New York.
Schreiner, C., Scharf, S., Stenzel, V., and Rössler, A. (2017). “Self-healing through microencapsulated agents for protective coatings,” J. Coatings Technol. Res. 14(4), 809-816. DOI: 10.1007/s11998-017-9921-x
Shi, Q., Wong, S. C., Ye, W., Hou, J. W., Zhao, J., and Yin, J. H. (2012). “Mechanism of adhesion between polymer fibers at nanoscale contacts,” Langmuir 28(10), 4663-4671. DOI: 10.1021/la204633c
Shukla, S., Pandey, K. K., and Kishan Kumar, V. S. (2019). “Chemical and application properties of some solvent and water based coatings on wooden substrate,” Drvna Industrija 70(2), 107-114. DOI: 10.5552/drvind.2019.1809
Silva, T. A. R., Marques, A. C., dos Santos, R. G., Shakoor, R. A., Taryba, M., and Montemor, M. F. (2023). “Development of biopolyurethane coatings from biomass-derived alkylphenol polyols – A green alternative,” Polymers 15(11), article 2561. DOI: 10.3390/polym15112561
Simões da Silva, M. T. Q., Cardosa, M. d. R., Veronese, C. M. P., and Mazer, W. (2022). “Tortuosity: A brief review,” Mater. Today 58(4), 1344-1349. DOI: 10.1016/j.matpr.2022.02.228
Singh, A. P., and Dawson, B. S. W. (2006). “Microscopic assessment of the effect of saw-textured Pinus radiata plywood surface on the distribution of a film-forming acrylic stain,” JCT Res. 3(3), 193-201. DOI: 10.1007/BF02774508
Singh, A. P., Ratz, A., and Dawson, B. S. W. (2007). “A novel method for high-resolution imaging of coating distribution within a rough-textured plywood surface,” J. Coatings Technol. Res. 4(2), 207-210. DOI: 10.1007/s11998-007-9019-y
Slabejová, G., Smidriaková, M., and Pánis, D. (2018). “Quality of silicone coating on the veneer surfaces,” BioResources 13(1), 776-788. DOI: 10.15376/biores.13.1.776-788
Slabejová, G., Smidriaková, M., and Svocák, J. (2020). “Interlayer with microcapsules and its influence on the surface finish quality of wood,” Acta Facultatis Xylologiae Zvolen 62(2), 61-74. DOI: 10.17423/afx.2020.62.2.06
Smith, I., Snow, M., Asiz, A., and Vasic, S. (2007). “Failure mechanisms in wood-based materials: A review of discrete, continuum, and hybrid finite-element representations,” Holzforschung 61(4), 352-359. DOI: 10.1515/HF.2007.055
Sögütlü, C., Nzokou, P., Koc, I., Tutgun, R., and Döngel, N. (2016). “The effects of surface roughness on varnish adhesion strength of wood materials,” J. Coatings Technol. Res. 13(5), 863-870. DOI: 10.1007/s11998-016-9805-5
Song, X. X., Guo, W. X., Zhu, Z. P., Han, G. P., and Cheng, W. L. (2024). “Preparation of uniform lignin/titanium dioxide nanoparticles by confined assembly: A multifunctional nanofiller for a waterborne polyurethane wood coating,” Int. J. Biol. Macromol. 258(1), article 128827. DOI: 10.1016/j.ijbiomac.2023.128827
Song, X. X., Tang, S., Chi, X., Han, G. P., Bai, L., Shi, S. Q., Zhu, Z. P., and Cheng, W. L. (2022). “Valorization of lignin from biorefinery: colloidal lignin micro- nanospheres as multifunctional bio-based fillers for waterborne wood coating enhancement,” ACS Sustain. Chem. Eng. 10(35), 11655-11665. DOI: 10.1021/acssuschemeng.2c03590
Song, X. X., Wei, J. J., Mao, Z. Y., Chi, X., Zhu, Z. P., Han, G. P., and Cheng, W. L. (2023). “Effect of hot-air drying conditions on the drying efficiency and performance of a waterborne coating on pine wood,” Forests 14(9), article 1752. DOI: 10.3390/f14091752
Sow, C., Riedl, B., and Blanchet, P. (2011). “UV-waterborne polyurethane-acrylate nanocomposite coatings containing alumina and silica nanoparticles for wood: Mechanical, optical, and thermal properties assessment,” J. Coatings Technol. Res. 8(2), 211-221. DOI: 10.1007/s11998-010-9298-6
Steiner, A. J., Lunzer, F., Vanheertum, T., Kittler, T., Sterckx, L., and Gariepy, K. (2020). “Novel oil-modified acrylic multi-domain dispersions for wood coatings,” Surface Coatings Int. 103(3), 128-131.
Stenberg, C., Svensson, M., Wallström, E., and Johansson, M. (2005). “Drying of linseed oil wood coatings using reactive diluents,” Surf. Coatings Int. Part B – Coat. Trans. 88(2), 119-126. DOI: 10.1007/BF02699543
Stirling, R., and Temiz, A. (2014). “Fungicides and insecticides used in wood preservation,” in: Deterioration and Protection of Sustainable Biomaterials, Schultz, T. P., Goodell, B., and Nicholas, D. D. (eds.), Vol 1158, pp. 185-201. DOI: 10.1021/bk-2014-1158.ch010
Su, Y. P., Zhang, S. T., Chen, Y. W., Yuan, T., and Yang, Z. H. (2020). “One-step synthesis of novel renewable multi-functional linseed oil-based acrylate prepolymers and its application in UV-curable coatings,” Prog. Organic Coatings 148, article 105820. DOI: 10.1016/j.porgcoat.2020.105820
Sun, F. C., Fu, J. H., Peng, Y. X., Jiao, X. M., Liu, H., Du, F. P., and Zhang, Y. F. (2021). “Dual-functional intumescent fire-retardant/self-healing water-based plywood coatings,” Prog. Organic Coatings 154, article 106187. DOI: 10.1016/j.porgcoat.2021.106187
Taghiyari, H., Esmailpour, A., and Papadopoulos, A. (2019). “Paint pull-off strength and permeability in nanosilver-impregnated and heat-treated beech wood,” Coatings 9(11), article 723. DOI: 10.3390/coatings9110723
Tamantini, S., Bergamasco, S., Zikeli, F., Humar, M., Cavalera, M., and Romagnoli, M. (2023). “Cellulose nano crystals (CNC) as additive for a bio-based waterborne acrylic wood coating: Decay, artificial weathering, physical and chemical tests,” Nanomater. 13(3), article 442. DOI: 10.3390/nano13030442
Tao, X., Tian, D. X., Liang, S. Q., Li, S. M., Peng, L. M., and Fu, F. (2024). “Enhanced paint adhesion of puffed wood-based metal composites via surface treatment with silane coupling agent,” Wood Mater. Sci. Eng. 19(3), 573-579. DOI: 10.1080/17480272.2023.2269414
Tao, Y., Yan, X. X., and Chang, Y. J. (2021). “Effect of coating process on mechanical, optical, and self-healing properties of waterborne coating on basswood surface with MF-coated shellac core microcapsule,” Polymers 13(23), article 4228. DOI: 10.3390/polym13234228
Tathe, D. S., and Jagtap, R. N. (2015). “Biobased reactive diluent for UV-curable urethane acrylate oligomers for wood coating,” J. Coating Technol. Res. 12(1), 187-196. DOI: 10.1007/s11998-014-9616-5
Teaca, C. A., Rosu, D., Mustata, F., Rusu, T., Rosu, L., Rosca, I., and Varganici, C. D. (2019). “Natural bio-based products for wood coating and protection against degradation: A review,” BioResources 14(2), 4873-4901. DOI: 10.15376/biores.14.2.Teaca
Trovagunta, R., Marquez, R., Tolosa, L., Barrios, N., Zambrano, F., Suarez, A., Pal, L., Gonzalez, R., and Hubbe, M. A. (2024). “Lignin self-assembly phenomena and valorization strategies for pulping, biorefining, and materials development: Part 1. The physical chemistry of lignin self-assembly,” Adv. Colloid Interface Sci. 332, article 103247. DOI: 10.1016/j.cis.2024.103247
Tshabalala, M. A., and Gangstad, J. E. (2003). “Accelerated weathering of wood surfaces coated with multifunctional alkoxysilanes by sol-gel deposition,” J. Coatings Technol. 75(943), 37. DOI: 10.1007/BF02730098
Tshabalala, M. A., Libert, R., and Schaller, C. M. (2011). “Photostability and moisture uptake properties of wood veneers coated with a combination of thin sol-gel films and light stabilizers,” Holzforschung 65(2), 215-220. DOI: 10.1515/HF.2011.022
Turku, I., Rohumaa, A., Tirri, T., and Pulkkinen, L. (2024). “Progress in achieving fire-retarding cellulose-derived nano/micromaterial-based thin films/coatings and aerogels: A review,” Fire – Switzerland 7(1), article 31. DOI: 10.3390/fire7010031
Van den Bulcke, J., Rijckaert, V., Van Acker, J., and Stevens, M. (2003). “Quantitative measurement of the penetration of water-borne coatings in wood with confocal lasermicroscopy and image analysis,” Holz als Roh- und Werkstoff 61(4), 304-310. DOI: 10.1007/s00107-003-0394-5
Van den Bulcke, J., Rijckaert, V., Van Acker, J., and Stevens, M. (2006). “Adhesion and weathering performance of waterborne coatings applied to different temperate and tropical wood species,” JCT Res. 3(3), 185-191.
Van den Bulcke, J., Van Acker, J., and Stevens, M. (2008). “Experimental and theoretical behavior of exterior wood coatings subjected to artificial weathering,” J. Coatings Technol Res. 5(2), 221-231. DOI: 10.1007/s11998-007-9074-4
van Meel, P. A., Erich, S. J. F., Huinink, H. P., Kopinga, K., de Jong, J., and Adan, O. C. G. (2011). “Moisture transport in coated wood,” Prog. Organic Coat. 72(4), 686-694. DOI: 10.1016/j.porgcoat.2011.07.011
Vardanyan, V., Poaty, B., Chauve, G., Landry, V., Galstian, T., and Riedl, B. (2014). “Mechanical properties of UV-waterborne varnishes reinforced by cellulose nanocrystals,” J. Coatings Technol. Res. 11(6), 841-852. DOI: 10.1007/s11998-014-9598-3
Varganici, C. D., Rosu, L., Rosu, D., Mustata, F., and Rusu, T. (2021). “Sustainable wood coatings made of epoxidized vegetable oils for ultraviolet protection,” Environ. Chem. Lett. 19(1), 307-328. DOI: 10.1007/s10311-020-01067-w
Veigel, S., Grüll, G., Pinkl, S., Obersriebnig, M., Müller, U., and Gindl-Altmutter, W. (2014). “Improving the mechanical resistance of waterborne wood coatings by adding cellulose nanofibres,” Reactive Func. Polym. 85, 214-220. DOI: 10.1016/j.reactfunctpolym.2014.07.020
Veigel, S., Lems, E. M., Grüll, G., Hansmann, C., Rosenau, T., Zimmermann, T., and Gindl-Altmutter, W. (2017). “Simple green route to performance improvement of fully bio-based linseed oil coating using nanofibrillated cellulose,” Polymers 9(9), article 425. DOI: 10.3390/polym9090425
Vidholdová, Z., Slabejová, G., and Kaloc, J. (2017). “Influence of wood pre-weathering on selected surface properties of the system wood – coating film,” Acta Facultatis Xylologiae Zvolen 59(2), 67-77. DOI: 10.17423/afx.2017.59.2.07
Vitosyte, J., Ukvalbergiene, K., and Keturakis, G. (2013). “The effects of surface roughness on adhesion strength of coated ash (Fraxinus excelsior L.) and birch (Betula L.) wood,” Mater. Sci. – Medziagotyra 18(4), 347-351. DOI: 10.5755/j01.ms.18.4.3094
Volkmer, T., Glaunsinger, M., Mannes, D., Sonderegger, W., and Niemz, P. (2015). “Investigations on water vapour permeability of wood coatings for exterior use,” Bauphysik 37(3), 186-U104. DOI: 10.1002/bapi.201510020
Wang, J., Wu, H. G., Liu, R., Long, L., Xu, J. F., Chen, M. G., and Qiu, H. Y. (2019). “Preparation of a fast water-based UV cured polyurethane-acrylate wood coating and the effect of coating amount on the surface properties of oak (Quercus alba L.),” Polymers 11(9), article 1414. DOI: 10.3390/polym11091414
Wang, L., Han, Y., and Yan, X. X. (2022). “Effects of adding methods of fluorane microcapsules and shellac resin microcapsules on the preparation and properties of bifunctional waterborne coatings for basswood,” Polymers 14(18), article 3919. DOI: 10.3390/polym14183919
Wang, L., Kelly, P. V., Ozveren, N., Zhang, X. F., Korey, M., Chen, C., Li, K., Bhandari, S., Tekinalp, H., Zhao, X. H., Wang, J. W., Seydibeyoglu, M. O., Alyarnac-Seydibeyoglu, E., Gramlich, W. M., Tajvidi, M., Webb, E., Ozcan, S., and Gardner, D. J. (2023). “Multifunctional polymer composite coatings and adhesives by incorporating cellulose nanomaterials,” Matter 6(2), 344-372. DOI: 10.1016/j.matt.2022.11.024
Waring, R. G. (1963). Modern Wood Finishing, Bruce Pub. Co., Milwaukee.
Washburn, E. W. (1921). “The dynamics of capillary flow,” Physics Review 17(3), 273-283. DOI: 10.1103/PhysRev.17.273
Wenning, A. (2002). “Tailor-made UV-curable powder clear coatings for metal applications,” Macromol. Symp. 187, 597-603. DOI: 10.1002/1521-3900(200209)187:1<597::AID-MASY597>3.0.CO;2-E
Weththimuni, M. L., Canevari, C., Legnani, A., Licchelli, M., Malagodi, M., Ricca, M., and Zeffiro, A. (2016). “Experimental characterization of oil-colophony varnishes: A preliminary study,” Int. J. Conservation Sci. 7, 813-826.
Weththimuni, M. L., Milanese, C., Licchelli, M., and Malagodi, M. (2021). “Improving the protective properties of shellac-based varnishes by functionalized nanoparticles,” Coatings 11(4), article 419. DOI: 10.3390/coatings11040419
Williams, R. S., Winandy, J. E., and Feist, W. C. (1987). “Paint adhesion to weathered wood,” J. Coatings Technol. 59(749), 43-49.
Williams, R. S., and Feist, W. C. (1993). “Durability of paint or solid-color stain applied to preweathered wood,” For. Prod. J. 43(1), 8-14.
Williams, R. S., and Feist, W. C. (1994). “Effect of preweathering, surface-roughness, and wood species on the performance of paint and stains,” J. Coatings Technol. 66(828), 109-121.
Williams, R. S., Knaebe, M. T., and Feist, W. C. (1996). Finishes for Exterior Wood: Selection, Application, and Maintenance: A Comprehensive Guide to the Painting/Staining and Maintenance of Homes, Decks, Log Structures, and More, Forest Products Society, Madison, WI.
Williams, R. S., and Feist, F. C. (1999). “Water repellents and water-repellent preservatives for wood,” Forest Products Laboratory, GTR-109. DOI: 10.2737/FPL-GTR-109
Williams, R. S., and Feist, W. C. (2001). “Duration of wood preweathering: Effect on the service life of subsequently applied paint,” J. Coatings Technol. 73(920), 65-72. DOI: 10.1007/BF02698377
Williams, R. S., Swan, L., Sotos, P., Knaebe, M., and Feist, W. C. (2005). “Performance of finishes on western juniper lumber and particleboard during outdoor exposure,” For. Prod. J. 55(11), 65-72.
Wolf, J. R., and Strieder, W. (1990). “Surface and void tortuosities for a random fiber bed – Overlapping, parallel cylinders of several radii,” J. Membrane Sci. 49(1), 103-115. DOI: 10.1016/S0376-7388(00)80781-4
Wright, N. C., Li, J. C., and Guo, M. R. (2006). “Microstructural and mold resistant properties of environment-friendly oil-modified polyurethane based wood-finish products containing polymerized whey proteins,” J. Appl. Polym. Sci. 100(5), 3519-3530. DOI: 10.1002/app.22432
Wu, Q. L., Li, W. B., and Yan, X. X. (2023). “Effect of microcapsules on mechanical, optical, self-healing and electromagnetic wave absorption in waterborne wood paint coatings,” Coatings 13(9), article 1478. DOI: 10.3390/coatings13091478
Wu, X. Y., Yang, F., Gan, J., Zhao, W. Y., and Wu, Y. (2021). “A flower-like waterborne coating with self-cleaning, self-repairing properties for superhydrophobic applications,” J. Mater. Res. Technol. – JMR&T 14, 1820-1829. DOI: 10.1016/j.jmrt.2021.07.096
Wu, Y., Wu, X. Y., Yang, F., Zhang, H. Q., Feng, X. H., and Zhang, J. L. (2020). “Effect of thermal modification on the nano-mechanical properties of the wood cell wall and waterborne polyacrylic coating,” Forests 11(12), article 1247. DOI: 10.3390/f11121247
Xia, Y. X., Li, W. B., and Yan, X. X. (2023). “Effect of shellac microcapsules blended with carbonyl iron powder and carbon nanotubes on the self-healing and electromagnetic wave absorption properties of waterborne coatings on fiberboard surface,” Coatings 13(9), article 1572. DOI: 10.3390/coatings13091572
Xia, Y. X., Yan, X. X., and Peng, W. W. (2022). “Preparation of cellulose modified wall material microcapsules and its effect on the properties of wood paint coating,” Polymers 14(17), article 3534. DOI: 10.3390/polym14173534
Xie, Y. J., Krause, A., Militz, H., and Mai, C. (2006). “Coating performance of finishes on wood modified with an N-methylol compound,” Prog. Organic Coatings 57(4), 291-300. DOI: 10.1016/j.porgcoat.2006.06.010
Xu, D. D., Liang, G. T., Qi, Y. R., Gong, R. Z., Zhang, X. Q., Zhang, Y. M., Liu, B. X., Kong, L. L., Dong, X. Y., and Li, Y. F. (2022). “Enhancing the mechanical properties of waterborne polyurethane paint by graphene oxide for wood products,” Polymers 14(24), article 5456. DOI: 10.3390/polym14245456
Yan, X. X., Chang, Y. J., and Qian, X. Y. (2019). “Effect of the concentration of pigment slurry on the film performances of waterborne wood coatings,” Coatings 9(10), article 635. DOI: 10.3390/coatings9100635
Yan, X. X., Qian, X. Y., and Chang, Y. J. (2019). “Preparation and characterization of urea formaldehyde @ epoxy resin microcapsule on waterborne wood coatings,” Coatings 9(8), article 475. DOI: 10.3390/coatings9080475
Yan, X. X., and Peng, W. W. (2020). “Preparation of microcapsules of urea formaldehyde resin coated waterborne coatings and their effect on properties of wood crackle coating,” Coatings 10(8), article 764. DOI: 10.3390/coatings10080764
Yan, X. X., and Wang, L. (2020). “Preparation of shellac resin microcapsules coated with urea formaldehyde resin and properties of waterborne paint films for Tilia amurensis Rupr.,” Membranes 10(10), article 278. DOI: 10.3390/membranes10100278
Yan, X. X., Tao, Y., and Qian, X. Y. (2020a). “Preparation and optimization of waterborne acrylic core microcapsules for waterborne wood coatings and comparison with epoxy resin core,” Polymers 12(10), article 2366. DOI: 10.3390/polym12102366
Yan, X. X., Wang, L., and Qian, X. Y. (2020b). “Effect of microcapsules with different core-wall ratios on properties of waterborne primer coating for European linden,” Coating 10(9), article 826. DOI: 10.3390/coatings10090826
Yan, X. X., and Peng, W. W. (2021). “Effect of microcapsules of a waterborne core material on the properties of a waterborne primer coating on a wooden surface,” Coatings 11(6), article 657. DOI: 10.3390/coatings11060657
Yan, X. X., Peng, W. W., and Qian, X. Y. (2021a). “Effect of water-based acrylic acid microcapsules on the properties of paint film for furniture surface,” Appl. Sci. – Basel 11(16), article 7586. DOI: 10.3390/app11167586
Yan, X. X., Tao, Y., and Chang, Y. J. (2021b). “Effect of shellac waterborne coating microcapsules on the optical, mechanical and self-healing properties of waterborne primer on Tilia europaea L. wood,” Coatings 11(7), article 785. DOI: 10.3390/coatings11070785
Yan, X. X., Zhao, W. T., and Qian, X. Y. (2021c). “Effect of water-based emulsion core microcapsules on aging resistance and self-repairing properties of water-based coatings on linden,” Appl. Sci. – Basel 11(10), article 4662. DOI: 10.3390/app11104662
Yan, X. X., Zhao, W. T., and Wang, L. (2021d). “Preparation and performance of thermochromic and self-repairing dual function paint film with lac resin microcapsules and fluorane microcapsules,” Polymers 13(18), article 3109. DOI: 10.3390/polym13183109
Yan, X. X., Li, W. B., Han, Y., and Yin, T. Y. (2022a). “Preparation of melamine/rice husk powder coated shellac microcapsules and effect of different rice husk powder content in wall material on properties of wood waterborne primer,” Polymers 14(1), article 72. DOI: 10.3390/polym14010072
Yan, X. X., Zhao, W. T., and Wang, L. (2022b). “Mechanism of thermochromic and self-repairing of waterborne wood coatings by synergistic action of waterborne acrylic microcapsules and fluorane microcapsules,” Polymers 14(1), article 56. DOI: 10.3390/polym14010056
Yang, L., and Kruse, B. (2004). “Revised Kubelka-Munk theory. I. Theory and application,” J. Optical Soc. Amer. 21(10), 1933-1941. DOI: 10.1364/JOSAA.21.001933
Yaremchuk, L., Hogaboam, L., Slabejová, G., and Sedliacik, J. (2023). “Comparative analysis of the quality properties of oil-based and alkyd coating materials for wood,” Acta Facultatis Xylologiae Zvolen 65(1), 63-71. DOI: 10.17423/afx.2023.65.1.06
Yona, A. M. C., Zigon, J., Dahle, S., and Petric, M. (2021a). “Study of the adhesion of silicate-based coating formulations on a wood substrate,” Coatings 11(1), article 61. DOI: 10.3390/coatings11010061
Yona, A. M. C., Zigon, J., Kamlo, A. N., Pavlic, M., Dahle, S., and Petric, M. (2021b). “Preparation, surface characterization, and water resistance of silicate and sol-silicate inorganic-organic hybrid dispersion coatings for wood,” Mater. 14(13), article 3559. DOI: 10.3390/ma14133559
Yoo, Y., and Youngblood, J. P. (2017). “Tung oil wood finishes with improved weathering, durability, and scratch performance by addition of cellulose nanocrystals,” ACS Appl. Mater. Interfac. 9(29), 24936-24946. DOI: 10.1021/acsami.7b04931
Zareanshahraki, F., and Mannari, V. (2021). “Formulation and optimization of radiation-curable nonisocyanate polyurethane wood coatings by mixture experimental design,” J. Coatings Technol. Res. 18(3), 695-715. DOI: 10.1007/s11998-020-00453-x
Zhang, J. Y., Long, C., Zhang, X., Liu, Z., Zhang, X. L., Liu, T., Li, J. Z., and Gao, Q. (2022). “An easy-coating, versatile, and strong soy flour adhesive via a biomineralized structure combined with a biomimetic brush-like polymer,” Chem. Eng. J. 450(4), article 138387. DOI: 10.1016/j.cej.2022.138387
Zhang, H. Q., Feng, X. H., Wu, Y., Wu, Z. H. (2024). “Self-matting waterborne polyurethane acrylate wood coating by 222 nm far-UVC irradiation in ambient air,” Prog. Organic Coat. 189, article 018305. DOI: 10.1016/j.porgcoat.2024.108305
Zhang, L., Huang, Y. B., Sun, P., Hai, Y., and Jiang, S. H. (2021). “A self-healing, recyclable, and degradable fire-retardant gelatin-based biogel coating for green buildings,” Soft Matter 17(20), 5231-5239. DOI: 10.1039/d1sm00435b
Zhang, M. Y., Lu, J. X., Li, P., Li, X. J., Yuan, G. M., and Zuo, Y. F. (2022). “Construction of high-efficiency fixing structure of waterborne paint on silicate-modified poplar surfaces by bridging with silane coupling agents.” Prog. Organic Coat. 167, article 106846. DOI: 10.1016/j.porgcoat.2022.106846
Zhang, M. Y., Lyu, J. X., Zuo, Y. F., Li, X. G., and Li, P. (2023). “Effect of KH560 concentration on adhesion between silicate modified poplar and waterborne varnish,” Prog. Organic Coatings 174, article 107267. DOI: 10.1016/j.porgcoat.2022.107267
Zheng, B., Ge, S. S., Wang, S., Shao, Q., Jiao, C. Y., Liu, M., Das, R., Dong, B. B., and Guo, Z. H. (2020). “Effect of γ-aminopropyltriethoxysilane on the properties of cellulose acetate butyrate modified acrylic waterborne coatings,” React. Func. Polym. 154, article 104657. DOI: 10.1016/j.reactfunctpolym.2020.104657
Zheng, R. B., Tshabalala, M. A., Li, Q. Y., and Wang, H. Y. (2015). “Weathering performance of wood coated with a combination of alkoxysilanes and rutile TiO2 hierarchical nanostructures,” BioResources 10(4), 7053-7064. DOI: 10.15376/biores.10.4.7053-7064
Zhou, K. K., Xu, M. P., Wu, W. J., and Wang, J. (2023). “Preparation of nano silicon carbide modified UV paint and its application performance on wood flooring surface,” Polymers 15(23), article 4584. DOI: 10.3390/polym15234584
Zhu, X. D., Liu, Y., and Shen, J. (2016). “Volatile organic compounds (VOCs) emissions of wood-based panels coated with nanoparticles modified water based varnish,” Eur. J. Wood Wood Prod. 74(4), 601-607. DOI: 10.1007/s00107-016-1012-7
Zhu, Y., Li, W. B., Xia, Y. X., Hang, J. Y., Yan, X. X., and Li, J. (2024). “Effect of blending of shellac, carbonyl iron powder, and carbonyl iron powder/carbon nanotube microcapsules on the properties of coatings,” Coatings 14(1), article 75. DOI: 10.3390/coatings14010075
Zigon, J., Kovac, J., and Petric, M. (2022). “The influence of mechanical, physical and chemical pre-treatment processes of wood surface on the relationships of wood with a waterborne opaque coating,” Prog. Organic Coatings 162, article 106574. DOI: 10.1016/j.porgcoat.2021.106574
Zigon, J., Pavlic, M., Kibleur, P., Van den Bulcke, J., Petric, M., Van Acker, J., and Dahle, S. (2021). “Treatment of wood with atmospheric plasma discharge: Study of the treatment process, dynamic wettability and interactions with a waterborne coating,” Holzforschung 75(7), 603-613. DOI: 10.1515/hf-2020-0182
Zigon, J., Petric, M., and Dahle, S. (2018). “Dielectric barrier discharge (DBD) plasma pretreatment of lignocellulosic materials in air at atmospheric pressure for their improved wettability: A literature review,” Holzforschung 72(11), 979-991. DOI: 10.1515/hf-2017-0207