Next Article in Journal
High-Performance Nanoscale Metallic Multilayer Composites: Techniques, Mechanical Properties and Applications
Previous Article in Journal
Fracture Resistance of 3D-Printed Occlusal Veneers Made from 3Y-TZP Zirconia
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Phytochemicals from Bark Extracts and Their Applicability in the Synthesis of Thermosetting Polymers: An Overview

by
Tomasz Szmechtyk
* and
Magdalena Małecka
Department of Physical Chemistry, Faculty of Chemistry, University of Łódź, Pomorska 163/165, 90-236 Łódź, Poland
*
Author to whom correspondence should be addressed.
Materials 2024, 17(9), 2123; https://doi.org/10.3390/ma17092123
Submission received: 1 March 2024 / Revised: 9 April 2024 / Accepted: 19 April 2024 / Published: 30 April 2024
(This article belongs to the Section Polymeric Materials)

Abstract

:
This review focuses on recent research on the phytochemicals found in bark from different trees and their potential to be used as substrates for the synthesis of thermosetting resins. Recent studies about the influence of each bark harvesting step on the extracted phytochemicals, from debarking to extraction, are investigated. A comparison of bark extracts in terms of the correlation between extraction conditions and efficiency (based on the total phenolic content (TPC) and extraction yield) is presented for six groups of trees (Norway spruce, pine species, other conifers, oak species, other deciduous trees of the north temperate zone, tropical and subtropical trees) and evaluated. The evaluation revealed that there is an interesting relationship between the extraction time and the type of solvent for some types of tree bark. It was found that a relatively short extraction time and a solvent temperature close to the boiling point are favourable. The latest research on the application of bark extracts in different types of thermosetting resins is described. This review discusses the attractiveness of bark extracts in terms of functional groups and the possibilities arising from extractable phytochemicals. In addition, different approaches (selective versus holistic) and methods of application are presented and compared.

1. Introduction

Petroleum-based compounds are still the main source of reagents for polymer synthesis. However, current research is increasingly focusing on the development of alternative bio-based materials. They are especially relevant for thermosetting resins, which cannot be widely recycled like thermoplastics. These activities result from both ecological and economic issues, and therefore, the management of waste of natural origin is of particular importance as a double benefit. A potential supplier of this type of bio-waste could be the wood industry, whose annual global production is 5 billion m3 of wood products [1]. The primary by-product is bark, which is not included in this annual production, as it is mainly used as fuel in sawmills. Recently, the use of bark in horticulture has increased the market share of bark, but its use is still relatively low (about 30% and 15% of hardwood and softwood bark supplies in the USA, respectively) [2].
The share of bark in the total volume of a tree or the total weight of a tree depends on the species, size and age of the tree and on the measurement and calculation methods [3,4]. Also, regarding the difference between inner and outer bark, the density of the bark is lower (in the case of inner bark) or much lower (in the case of outer bark) than the density of the wood [5]. Nevertheless, it can be assumed that the volume fraction of bark varies between 10% and 25% for most tree species at the age appropriate for felling. This gives the global amount of bio-waste ranging from 0.56 to 1.67 billion m3. Furthermore, the approximate quantitative determination of extractable compounds from bark is difficult due to the age and species of the tree, the maturity of the bark [6], the season of bark harvesting, soil conditions and the extraction method (the type of solvent is especially crucial) [6,7,8]. Rough estimations can be made with bark densities in the range of 200–700 kg/m3 and extraction yields in the range of 5–35% [6,7,9,10,11,12]. It can be assumed that the amount of extractable compounds from bark that is obtained annually by the wood industry is about 80 million metric tons (for values averaged over three ranges). This value corresponds to 20% of the world’s plastic production in 2021 (390.7 million metric tons) and almost double the global demand for thermosetting resins in 2021 (49.2 million metric tons) [13].
The aim of this review was to investigate the sourcing process of phytochemicals from bark as extracts and their applicability to thermosetting resins (Figure 1). We present the collected data as an introduction to a discussion about the advantages and disadvantages of phytochemicals and the ways they are applied in thermosets.

2. Bark Harvesting

2.1. Types of Bark

Most species of trees have persistent bark that should only be harvested industrially as a by-product of felling. It is important to avoid causing unnecessary damage to the bark because this can result in an increased risk of tree disease and pest infestation. Strategies for obtaining persistent bark are therefore limited by the felling time. Generally, it is important to consider circumstances such as the season of felling, weather conditions, or the place of growth to better understand their impact on bark phytochemistry. An exception to the aforementioned rule is cork oak (Quercus suber), which has highly valuable bark (cork), resistant heartwood and the ability to regrow new outer bark. Cork is harvested industrially from living trees, but the process is very complex, and trees are still exposed to harmful environmental conditions after harvesting [14].
Another group with different bark harvesting capabilities includes trees with exfoliating bark (Figure 2). The most notable representatives of this group are paperbark maple (Acer griseum), red maple (A. rubrum), most species of birch (Betula papyrifera, B. pendula, B. maximowicziana, B. utilis, B. pubescens, B. nigra, B. alleghaniensis, etc.), all eight species of plane tree (Platanus) [15], Chinese elm (Ulmus parvifolia), lacebark pine (Pinus bungeana), Sitka spruce (Picea sitchensis), various species of juniper (Juniperus communis, J. virginiana and J. oxycedrus), European yew (Taxus baccata) and shagbark hickory (Carya ovata). Exfoliation is a natural process and facilitates the harmless harvesting of the outer bark, but the bark should be carefully peeled in the early part of the growing season to avoid damaging the tree. Either way, the industrial harvesting of this type of bark from living trees is likely to be the subject of an environmental debate.

2.2. Debarking Methods

Bark that is harvested industrially during felling can be obtained using the dry or wet debarking process. The dry debarking (rotor debarking) method, used in the mechanical wood industry, only needs water if the logs are frozen [16]. The wet debarking (drum debarking) process, designed for the chemical wood industry, causes the release of water extractives with debarking effluent [17]. Consequently, the dry process provides the bark with higher amounts of phytochemicals, theoretically making it a more advantageous option. On the other hand, the liquid waste obtained after the wet process can still be a source of selected phytochemicals. Kemppainen et al. [18] performed extraction with hot tap water and proved that wet debarking can remove both carbohydrates and tannins from Norway spruce bark (Picea abies). Peeters et al. [19] also confirmed the similarity of the polyphenolic compositions of the water from the wet debarking process and bark press water to the compounds extracted (water as solvent) from Norway spruce bark. Bark extracts have a higher content of polyphenols than both types of wastewater. They found that bark press water, in particular, could still be a valuable source of phytochemicals with relatively low molecular weights. Multia [20] investigated the presence of stilbene glucosides in spruce bark and bark press water, the low levels of which suggest that they were largely extracted at an earlier stage (possibly during the wet debarking process). Neither of the above-mentioned wet processes is optimised as a source of phytochemicals, and due to this, common contaminants such as chlorides or metals [21] should be taken into account.

2.3. Seasonal Variations in Phytochemicals

The season of bark harvesting is another crucial factor. The examination of peach (Prunus persica) bark [22] indicates a marked correlation between the summer season and a high concentration of rutin (from May to August) and diacetylated p-coumaroylsucrose (from June to August). An inverse relationship was observed for catechin, persicoside, hesperetin-5-O-glucoside and 4′-O-methyltaxifolin-5-O-glucoside (high concentrations from September to April). Medic et al. [23] found increased contents of hydrojuglone glycosides (glucoside, rhamnoside, pentoside) and naphthoquinones in walnut (Juglans regia) bark extracts from the summer season (June–September). Most of the water-extractable phytochemicals from the bark of willow (Salix sp.) have higher concentrations in autumn and winter (triandrin, sucrose, raffinose, (+)-catechin, salicin), with the exception of glucose and fructose, whose contents are highest in July and relatively high throughout the year (lowest in September) [24]. The total phenolic content (TPC) and catechin content of water–ethanol (50% and 70% EtOH) extracts from the winter and spring bark of three Acacia species (A. farnesiana, A. longifolia, A. tortilis) were compared by Gabr et al. [25]. The only visible patterns were for the TPC of A. tortilis (significantly higher values for winter bark extracts) and the catechin content of A. farnesiana (higher values for spring bark extracts). The examination of Ashoka tree (Saraca asoca) bark showed increased epicatechin content in the winter season (Hemant and Shishir in the Hindu calendar) and increased gallic acid content in the monsoon season (Varsha) and early winter (Hemant) [26]. The complex study of stilbenes and carbohydrates in Norway spruce (P. abies) bark indicates higher contents of stilbenes and free sugars in samples collected in winter and total non-structural carbohydrates (mainly due to the presence of starch) in samples collected at the turn of spring and summer [27,28]. Furthermore, the tannin yield is higher in winter spruce bark extracts [29]. Halmemies et al. [30] observed a similar trend of higher values of total phenolic content (TPC), monosaccharides and stilbenes for winter spruce bark extracts, while the levels of organic acids, alcohols, flavonoids and distilbenes were similar for extracts in both seasons. Although recent studies of spruce bark provide relatively comprehensive information on extracts from different seasons, most studies of seasonal variability lack information about all relevant external factors (weather conditions, type of soil, pest activity). These factors, in addition to the different solvents, the variety of extraction procedures and the complex metabolism of trees, make predicting seasonal changes in the composition of phytochemicals challenging.

2.4. Storage of Bark

Storage is also an important factor that affects the phytochemical content. The contents of the bark extractives of Scots pine (Pinus sylvestris) [31] and Norway spruce (P. abies) [17] (soluble in acetone) decrease after 8 weeks of storage (both as piles and outdoors, from August to October). Spruce bark is also more susceptible to the loss of water extractives during storage in summer than in winter, which is the result of higher UV radiation and higher temperatures [30]. The activity of insects and microorganisms is also a factor accelerating degradation [29], which can be observed as the self-heating of biomass piles [31].

2.5. Pest Infestations and Fungal Infections

The phytochemistry of living trees is affected by insects and microorganisms. Zhao et al. reported a significant loss of stilbene glucoside, lignan and flavonoid contents in the inner bark of Norway spruce (P. abies) inoculated with five types of blue-stain fungi, which are associated with bark beetles [32]. On the other hand, these trees increase the production of terpenoid oleoresins and polyphenols in response. The main end products are catechin, catechin–epicatechin dimers, and taxifolin and its glucoside, which are toxic to bark beetles and their fungal associates [33]. Also, American beech (Fagus grandifolia) increases catechin content upon Neonecteria sp. infection [34]. These significant changes in the composition of the extract may be of interest because infested trees are often felled to stop the spread of beetles, and their bark is undesirable waste.

3. Extract Preparation and Analysis

3.1. Pre-treatment

Bark is often obtained in the form of relatively large flakes or strips, requiring pre-treatment such as milling or grinding. Samples can be fresh, freeze-dried or air-dried prior to fragmentation, with the advantage of freeze-drying being its lower impact on the phytochemical content compared to air-drying. The ground bark is often fractionated to separate the finer particles, which provide more efficient extraction (better solvent penetration) [35]. Additionally, the selection of bark compartments (inner bark, outer bark) may be crucial for the production of desirable chemicals [5].

3.2. Solvent Selection

The solvents most frequently chosen for bark extraction are water and simple alcohols (methanol and ethanol). They are relatively inexpensive and provide a satisfactory yield of phytochemicals. Mixtures of these solvents are prepared in various proportions, as they are often more effective than either one alone [11,36,37,38]. Other less popular solvents are 1,4-dioxan, acetone, dimethylformamide, diethyl ether, ethyl acetate, chloroform, hexane, benzene, isopropanol and acetonitrile [37,39]. More complex mixtures, such as deep eutectic solvents based on choline chloride, are also being tested [35,40], but these mixtures have so far been more effective as lignin solvents [41]. Although the selection of the solvent is crucial, higher yields of polyphenols and other phytochemicals can be obtained with the optimised parameters (especially time and temperature) of the applied extraction method. Also, the stability of extractables (especially polyphenols) should be taken into account for more demanding conditions.

3.3. Extraction Method

Various methods for efficient extraction are widely applied. Simple extraction by maceration (ME) and conventional extraction (CE) with heating, stirring or shaking is still popular: 6.5% (5/77) and 28.6% (22/77) of all reported extractions for each bark–method combination (Table 1, Table 2, Table 3, Table 4, Table 5 and Table 6), respectively. With another 5 (6.5%) extractions without a specified method (CE/ME) [42], simple methods account for 41.6% of all reported extractions (32/77). More advanced but still relatively simple and inexpensive methods are Soxhlet extraction (SoxE, 5/77, 6.5%), ultrasound-assisted extraction (UAE, 27/77, 35.1%) and microwave-assisted extraction (MAE, 8/77, 10.4%). These three moderate methods account for 51.6% (39/77) of the extractions reported (mixed MAE-UAE method in one study [43] counted as one for this summary). Supercritical fluid extraction (SFE, 2/77, 2.6%, one mixed SFE-UAE method [44]), pressurised liquid extraction (PLE, 1/77, 1.3%) and accelerated solvent extraction (ASE, 4/77, 5.2%) are less popular (7/77, 9.1% in total). The effectiveness of all methods is restricted by the type of solvents and the set of parameters (again, time and temperature, as well as others specific to each method: pressure, frequency, power of microwaves, etc.).

3.4. Quantitative and Qualitative Analysis of Extracts

The qualitative and quantitative evaluation of extracted phytochemicals provides comprehensive information on their potential applications. The most common quantitative analysis is total phenolic content (TPC, expressed as milligrams of gallic acid per gram of dry weight of bark), which uses Folin–Ciocâlteu reagent to measure UV-Vis absorbance [45]. Another is the extraction yield, which is the percentage value of the extractive weight to the dry weight of extracted bark. Both were selected as key values in all the comparisons presented in the tables. Other popular quantitative methods are total flavonoid content (TFC) [36,42,43,44,46,47,48,49,50,51,52,53,54], total tannin content (TTC) [49,55,56,57] and antioxidant activity assays based on 2,2-diphenyl-1-picrylhydrazyl (DPPH) [36,37,38,40,42,43,44,47,48,49,50,51,52,53,54,55,56,57,58,59,60,61,62,63,64,65], 2,2’-azino-bis(3-ethylbenzothiazoline-6-sulfonic acid (ABTS) [36,37,38,43,46,47,48,53,55,56,57,58,62,64] and Ferric Reducing Antioxidant Power (FRAP) [36,37,38,42,43,44,46,47,48,53,58,61,64,66]. Antioxidant activity assays [11,60] with other assays and antimicrobial [36,38,44,55,56,57,63,67], inhibitory [38,53,55,57], cytotoxic [62,64] and antiplasmodial [53] activity measurement methods prove that the main areas of application of bark extractives are the medical, pharmaceutical and food industries. The only well-known quantitative method for the polymer industry is the estimation of the amount of reactive tannin towards formaldehyde, known as the Stiasny precipitation number [68,69]. The most common qualitative methods are liquid [6,36,37,38,42,44,57,58,61,62,63,67] and gas chromatography [42,44,63,64]), often combined with mass spectrometry (LC-MS, GC-MS) or with a photodiode array (LC-PDA). Less popular are phytochemical screening [52,59,60,64] and Fourier transform infrared spectroscopy (FTIR) [36,43,47], as they allow a more tentative identification of compounds or only the identification of classes of phytochemicals.

Chromatographic Identification of Phytochemicals: Conditions

The previously mentioned types of detectors—PDA and MS—can complement each other in liquid chromatography (as PDA-MS) [6,37,38,58,62], but a PDA alone was a slightly more common option here [36,42,44,57,61,63,67]. Two types of LC were predominantly used: ultra-pressure liquid chromatography (UPLC) and high-pressure liquid chromatography (HPLC). This was also reflected in the dimensions of the columns used. Most columns for UPLC were shorter. They were 100 mm long and had an internal diameter of 2.1 mm [36,62], 150 mm long with the same diameter [37] or 150 mm long with an internal diameter of 3 mm [6] or 4.6 mm [57,67]. HPLC columns were usually longer and had larger diameters, but this was not a rule. The standard dimensions of the normal-phase columns for HPLC were 250 mm × 4.6 mm [42,44,58]. The exceptions were the reverse-phase columns [38,61,63]. Anyway, all selected columns were C18. The separation temperature range was 25–35 °C. The flow rate ranged from 0.2 mL/min to 1.0 mL/min and the volume of injections from 5 µL to 20 µL. Methanol, acetonitrile, water and their solutions (concentrations from 0.1% to 2%) of acids (formic, acetic, trifluoroacetic and orthophosphoric) were chosen as mobile phases.
Gas chromatography was combined only with MS, and columns with the dimensions 30 m × 0.25 mm were used for all analyses. The heating procedures were different: from 40 °C to 260 °C (heating rate 4 °C/min, 10 min steps at 240 °C and 260 °C) [42], from 70 °C to 290 °C (heating rate 10 °C/min) [44] or from 60 °C to 300 °C (different heating rates were applied during the process: 3 °C/min, 2 °C/min and 10 °C/min) [64].

4. A Survey of Selected Recent Studies

Recent studies (since 2018) on bark extracts are compared in tables, which are divided as follows: Table 1—Norway spruce (Picea abies); Table 2—different pine species (Pinus sp.); Table 3—other conifers; Table 4—different oak species (Quercus sp.); Table 5—other deciduous trees of the north temperate zone; and Table 6—trees of tropical and subtropical zones. The solvent type, the temperature of the process (if available, not applicable for MAE and ME), the extraction method (with additional information) and the time were selected as input parameters. The TPC, extraction yield and identified abundant compounds were chosen as the results for comparison. The evaluation of the presented data is informative due to the complex relationships between all input parameters (not only those presented hereafter but also those from sections “Bark Harvesting”, “Pre-treatment” and many others) and the results (extraction efficiency). Almost all TPC results were converted to milligrams of gallic acid equivalent per gram of dry weight of bark (mg GAE/g DWB) to avoid confusing comparisons between different forms of TPC expression. The two most popular forms were the aforementioned mg GAE/g DWB and milligrams of GAE per gram dry weight of extract (mg GAE/g DWE). The form mg GAE/g DWB was selected as a better source of information on the value of each extraction set (bark, solvent, extraction method) in terms of polyphenol yield. Another advantage of DWB as the denominator herein is the better comparability of extracts as products of bark because processes such as the concentration or fractionation of extracts make comparisons based on DWE less relevant. Conversions were made according to the equation DWB = DWE × Ey, where Ey is the extraction yield. The locations of the tree species used to obtain all reported bark extracts are shown in Figure 3. Most of the bark samples of these tree species were collected in the European region, with the exception of the species shown in Table 6, which are mainly from Southeast Asia. This trend is likely due to the European Green Deal and the sustainability initiatives that have resulted from it. For Asian countries, interest arises from the need to take advantage of the favourable climate and local flora.

4.1. Evaluation of Norway Spruce Bark Extracts (Table 1)

Spinelli et al. [46] compared SFE and PLE with the most popular UAE for Norway spruce bark (Table 1), and UAE provided a competitive extraction yield with the highest total phenolic content (TPC) among them. Only PLE enables the production of comparable extracts with a higher total flavonoid content (TFC). A comparison of selected parameters for ASE [30,35] suggests that a shorter process time and water as a solvent (or a temperature closer to the boiling point of the solvent) may be favourable for higher TPC (for spruce bark). The highest TPC (110 mg GAE per g of dry weight of bark for bark obtained in winter) and extraction yield (35.7% for bark from summer) was obtained for ASE (1500 psi, 10 min static time) with high-purity water (at 120 °C) [30]. The compounds identified with relatively high yields were trans-resveratrol, dehydroabietic acid, glucose, gluconic acid, trans-isorhapontin and astringin (structures are presented in Scheme 1).
Table 1. Phytochemical extraction from Norway spruce (Picea abies) bark—comparison of selected recent studies (since 2018).
Table 1. Phytochemical extraction from Norway spruce (Picea abies) bark—comparison of selected recent studies (since 2018).
Ref.Solvent Type,
Temperature of
Extraction
Extraction Method a,
Time
TPC (mg GAE
/g DWB) b
Extraction Yield
(% DW)
Identified Abundant Compounds
[35]ethanol 96.6%, 100 °CASE, 1500 psi, 20 min (steam exposure)3.216.63n/a
ethanol 96.6%, 160 °CASE, 1500 psi, 30 min (steam exposure)2.3628.44n/a
ethanol 96.6%MAE3.21n/an/a
DES c—choline chloride–malic acid 1:1 (m/m), 60 °CCE (closed flask, continuous stirring),
1 h
9.0014.68n/a
DES c—choline chloride–maleic acid 1:1 (m/m), 60 °C20.0011.87
DES c—choline chloride–glycerol 1:2 (m/m), 60 °C17.0011.40
[46]ethanol–water 10:90 (v/v), 40 °CSFE, 100 bar,
105 min (dynamic time), 150 min (static time)
0.77 ± 0.022.86 ± 0.04trans-resveratrol d
ethanol–water 20:80 (v/v), 40 °C1.24 ± 0.073.07 ± 0.10
ethanol–water 40:60 (v/v), 40 °C2.50 ± 0.033.12 ± 0.02
water, 160 °CPLE, 50 bar, 5 min33.45 ± 1.4413.07 ± 0.86
ethanol, 180 °C46.32 ± 2.1712.79 ± 0.25
ethanol–water 70:30 (v/v), 54 °CUAE (39 kHz, bath, 200 W), 60 min54.97 ± 2.0012.33 ± 0.58
[17]acetone, ~56 °CSoxE, 15 minn/a11.83 ± 0.13 en/a
[40]DES13 c, 60 °CCE (closed flask, cont. stirring), 2 h5.31 ± 0.04n/an/a
DES14 c, 60 °C5.96 ± 0.07n/a
[30]ultra-high-quality water,
120 °C
ASE, 1500 psi, 10 min (static time)111.0
(winter) e
89.4
(summer) e
34.7
(winter) e
35.7
(summer) e
dehydroabietic acid, glucose,
gluconic acid, trans-isorhapontin,
astringin
a Types of extractions: SoxE—Soxhlet ex.; SFE—supercritical fluid ex.; PLE—pressurised liquid ex.; UAE—ultrasound-assisted ex.; CE—conventional ex.; ASE—accelerated solvent ex.; MAE—microwave-assisted ex. b TPC—total phenolic content, expressed as mg gallic acid equivalent per g dry weight of bark (mg GAE/g DWB); n/a—not available. c DES—deep eutectic solvent; DES 13 and DES 14—deep eutectic solvents based on choline chloride, lactic acid, 1,3-butanediol and water in molar ratios 1:4:1:1 and 1:5:1:1, respectively. All components were mixed to form a homogeneous liquid (60 °C; 30 min). Only the 2 best results are presented. d Trans-resveratrol was the only identified compound. e Extraction yield/TPC of fresh bark.

4.2. Evaluation of Pine Species Bark Extracts (Table 2)

The highest TPC (163.6 mg GAE per g of dry bark) was obtained for maritime pine (P. pinaster) conventionally extracted (115 min) in an equal-volume ethanol–water mixture [36]. The highest extraction yields (17.5–18.5%) were obtained using simple methods (SoxE, CE/ME) for maritime pine (P. pinaster) and Scots pine (P. sylvestris) in alcohol–water mixtures [11,42]. Despite this, such comparisons between species of the Pinus genus have limited rationality (specific site of growth and growth conditions, slightly different phytochemistry). CE, the most popular extraction method for pine species, can be more accurately evaluated. The results for maritime pine (P. pinaster) [36,70] suggest that an extraction time of approximately 2 h and a temperature close to the boiling point (lower TPC for water at 82 °C) are optimal for this method and this type of bark. Equal or nearly equal volume ratios of water and ethanol are also optimal for higher TPC, but the results of Japanese red pine reveal the potential of acetonitrile and isopropanol as alternatives to ethanol [37]. All the abundant compounds mentioned in Table 2 are presented in Scheme 2.
Table 2. Phytochemical extraction from bark of different pine species (Pinus sp.) *—comparison of selected recent studies (since 2018).
Table 2. Phytochemical extraction from bark of different pine species (Pinus sp.) *—comparison of selected recent studies (since 2018).
Ref.Solvent Type,
Temperature of
Extraction
Extraction Method a,
Time
TPC (mg GAE
/g DWB) b
Extraction Yield
(% DW)
Identified Abundant Compounds
[11]
mp
ethanol–water
50:50 (v/v), ~82 °C
SoxE, 4 h73.48 ± 1.8317.55 ± 0.16n/a
ethanol 96%, 78 °C63.38 ± 1.2617.08 ± 0.23
water, 100 °C50.09 ± 4.70ca. 8 c
[70]
mp
water, 95 °CCE (with ice on lid), 2 h101.1 ± 4.07.5n/a
[36]
mp
water, 82 °CCE,
115 min
48.1n/agallocatechin, taxifolin, ellagic acid
ethanol–water
30:70 (v/v), 82 °C
120.1n/ataxifolin, gallocatechin, naringenin, catechin, elagic acid
ethanol–water
50:50 (v/v), 82 °C
163.6n/ataxifolin, naringenin, catechin, ellagic acid
ethanol–water
70:30 (v/v), 82 °C
136.5n/a
ethanol–water
90:10 (v/v), 82 °C
123.8n/a
[61]
mp
water, ~100 °CCE, 15 min12.25 ± 0.03n/acatechin, taxifolin,
protocatechuic acid
[61]
sp
water, ~100 °CCE, 15 min14.77 ± 0.06n/acatechin, taxifolin,
caffeic acid
[61]
Sc
water, ~100 °CCE, 15 min5.42 ± 0.05n/acatechin, taxifolin, protocatechuic acid
[42]
Sc
methanol–water
65:35 (v/v)
n/a (CE/ME)8818.33myricetin,
eleutheroside
[49]
Sc
distilled water,
60 °C
CE (cont. stirring), 1 h12.33 ± 1.48 7.25 ± 1.37 5.80 ± 1.24 4.38 ± 0.94 dn/an/a
[50]
Sc
ethanol–water
60:40 (v/v), 50 °C
CE (cont. stirring), 20 minca 67 ccatechin, epicatechin
[37]
Jp
water, 60 °CCE
(extraction in heating
mantle),
9 h
9.046.18n/a
ethanol–water
20:80 (v/v), 60 °C
22.198.34
ethanol–water
40:60 (v/v), 60 °C
24.339.52
ethanol–water
60:40 (v/v), 60 °C
20.1511.76
ethanol–water
80:20 (v/v), 60 °C
14.7610.21
ethanol, 60 °C7.568.15
methanol–water
20:80 (v/v), 60 °C
16.768.36
methanol–water
40:60 (v/v), 60 °C
16.379.09
acetone–water
20:80 (v/v), 60 °C
20.268.41
acetone–water
40:60 (v/v), 60 °C
17.169.58
isopropanol–water
20:80 (v/v), 60 °C
24.5110.11
isopropanol–water
40:60 (v/v), 60 °C
29,4612.05
acetonitrile–water
20:80 (v/v), 60 °C
23.819.24
acetonitrile–water
40:60 (v/v), 60 °C
27.6811.37
[66]
Ap
ethanol–water
70:30 (v/v), room temperature
ME,
72 h
560.65 ± 44.00 en/acatechin, ferulic acid, taxifolin, caffeic acid f
[71]
Cp
ethanol–water
50:50 (v/v), 50 °C
UAE,
1 h
27.9 ± 0.311.14-vinyl guaiacol,
4-methylguaiacol,
guaiacol
* Compared pine species (abbreviations in the Ref. section): mp—maritime pine (P. pinaster); sp—stone pine (P. pinea); Sc—Scots pine (P. sylvestris); Jp—Japanese red pine (P. densiflora); Ap—Afghan pine (P. eldarica); Cp—Corsican pine (Pinus nigra subsp. laricio). a Types of extractions: SoxE—Soxhlet ex.; CE—conventional ex.; ME—ex. by maceration. b TPC—total phenolic content, expressed as mg gallic acid equivalent per g dry weight of bark (mg GAE/g DWB); n/a—not available. c Value obtained from an inaccurate graph. d In order from top to bottom: whole bark from the continental (cont.) zone, outer bark from the cont. zone, whole bark from the coastal (coa.) zone, outer bark from the coa. zone. e Incomparable, expressed as mg GAE per dry weight of extract (without extraction yield, TPC cannot be converted to mg GAE per dry weight of bark). f Quoted from their earlier article.

4.3. Evaluation of Other Conifers’ Bark Extracts (Table 3)

In this table, studies of European larch bark extracts (Larix decidua) and different firs (Abies sp. and Cunninghamia lanceolata) are presented together. Interestingly, the TPC results from larch studies are very diverse (from 6 to 145 mg GAE/g DWB) for relatively similar extraction conditions. However, temperature and time might be crucial parameters: all three extracts with low TPC were probably overheated (heating above 50 °C or exposure to microwaves for more than 1 h) [48]. The best TPC values were obtained by the UAE method (a horn sonicator provides a yield of polyphenols similar to that obtained by the ultrasound bath but 4 times faster). Polyphenols probably extracted from larch bark are temperature-sensitive, and longer exposure causes their decomposition and/or evaporation (the accompanying low extraction yields for these three extracts may also be indicative of this). Unfortunately, a qualitative analysis provides limited data: only astringin, 4-vinyl guaiacol, 4-methylguaiacol and guaiacol were identified (Scheme 3), but the last three are considered volatile. Only two studies on firs (Abies sp.) were collected, and the only possible conclusion is a relatively high extraction yield (around 20%) for alcohol–water mixtures as solvents. Compounds identified with relatively high yields were isorhamnetin glucoside, quercetin glycoside, isorhamnetin, (+)-catechin and myricetin (Scheme 3). The TPCs of Chinese fir (Cunninghamia lanceolata) bark extract indicate the better solubility of its polyphenols in ethanol (almost twice as high TPC compared to water extracts).
Table 3. Phytochemical extraction from bark of other conifers *—comparison of selected recent studies (since 2018).
Table 3. Phytochemical extraction from bark of other conifers *—comparison of selected recent studies (since 2018).
Ref.Solvent Type,
Temperature of
Extraction
Extraction Method a,
Time
TPC (mg GAE
/g DWB) b
Extraction Yield
(% DW)
Identified Abundant Compounds
[58]
El
ethanol–water
80:20 (v/v)
UAE (horn), 15 min145.22 ± 6.11n/aastringin
[71]
El
ethanol–water
50:50 (v/v), 50 °C
UAE,
1 h
143.7 ± 426.14-vinyl guaiacol,
4-methylguaiacol, guaiacol
[43]
El
ethanol–water
50:50 (v/v)
MAE-UAE (simultaneous, power:
100–300 W), 30–120 s
90 ± 315.1 ± 0.1n/a
[48]
El
ethanol–water
50:50 (v/v),
58.26 °C
CE (with orbital shaker, speed 120 rpm),
94.27 min
0.837.73n/a
ethanol–water
50:50 (v/v),
65 °C
UAE (bath) 94.76 min0.375.87
ethanol–water
50:50 (v/v)
MAE (power: 100 W),
62.66 min
0,888.21
[6]
sf
ethanol–water
50:50 (v/v)
ASEn/a21.63 cisorhamnetin glucoside, quercetin glycoside, isorhamnetin,
(+)-catechin
[42]
Cf
methanol–water, 65:35 (v/v)n/a (CE/ME)7319.10myricetin
[72]
Ch
1% NaOH (aq), 90 °CCE, 2 h149.321.26 ± 0.81n/a
ethanol 95%SoxE,
(4–5 cycles/h), 7 h
285.65.04 ± 0.22
water, 90 °CCE, 2 h162.72.46 ± 0.14
* Compared other conifers (abbreviations in the Ref. section): El—European larch (Larix decidua); sf—silver fir (Abies alba); Cf—Caucasian fir (Abies nordmanniana); Ch—Chinese fir (Cunninghamia lanceolata). a Types of extractions: UAE—ultrasound-assisted ex.; CE—conventional ex.; ASE—accelerated solvent ex.; MAE—microwave-assisted ex.; SoxE—Soxhlet ex. b TPC—total phenolic content, expressed as mg gallic acid equivalent per g dry weight of bark (mg GAE/g DWB); n/a—not available. c Average of mean values for all disc–tree configurations.

4.4. Evaluation of Oak Species Bark Extracts (Table 4)

The highest TPC value (79.3 ± 0.8 mg GAE/g DWB) was obtained from the bark of common oak (Quercus robur) as a result of a short (20 min) CE (room temperature) in an ethanol–water mixture (60:40, v/v) [51]. The short process time may be crucial for a better extraction of polyphenols from this type of bark, because a similar experiment at an elevated temperature (50 °C) gave comparable results (approx. 77 mg GAE/g DWB) [50]. Furthermore, the TPC values for longer extraction times were significantly lower, which may indicate that the polyphenols obtained are sensitive even to long-term exposure to mild heating. A relatively high TPC value for UAE heated for a long time (1 h, 50 °C) [47] led us to assume that UAE may be a more effective technique if it is used for a shorter period of time. The extraction yields were mentioned in only four studies, and the highest one (18.75%) was obtained by an unspecified extraction (CE/ME) in a mixture of methanol with water (65:35, v/v). The bark of other oak species was extracted by assisted extraction (UAE or MAE), and the TPC values exceeded 200 mg GAE/g DWB, with two exceptions (both were heated for one hour) [47,71]. These results partially confirm the hypothesis mentioned above about the advantage of UAE over CE for common oak bark extraction. On the other hand, a complex comparison of TPC values obtained by MAE and UAE (shorter and longer process times for the two methods) reveals microwaves to be a well-suited assisting technique for the fast heating of most oak species bark, especially for ethanol–water mixtures as solvents (highest TPC values for all oak species were obtained by MAE, 650 W, 18 min). Catechin and gallic acid are compounds identified for at least two species of oak. Other identified and abundant compounds for single species were p-hydroxybenzoic acid, epicatechin, myricetin (common oak), 4-vinyl guaiacol (northern red oak), vanillic acid (oak of Daléchamp) and caffeic acid (Hungarian oak). Their structures are presented in Scheme 4.
Table 4. Phytochemical extraction from bark of different oak species (Quercus sp.) *—comparison of selected recent studies (since 2018).
Table 4. Phytochemical extraction from bark of different oak species (Quercus sp.) *—comparison of selected recent studies (since 2018).
Ref.Solvent Type,
Temperature of
Extraction
Extraction Method a,
Time
TPC (mg GAE
/g DWB) b
Extraction Yield
(% DW)
Identified Abundant Compounds
[50]
co
ethanol–water
60:40 (v/v), 50 °C
CE (cont. stirring), 20 minca 77 cn/acatechin, epicatechin,
p-hydroxybenzoic acid
[49]
co
distilled water,
60 °C
CE (cont.
stirring),
1 h
18.09 ± 3.50 10.28 ± 2.05 17.68 ± 3.14 8.25 ± 1.55 dn/an/a
[51]
co
water, rtCE (cont.
stirring),
20 min
60.4 ± 1.3n/an/a
ethanol–water
60:40 (v/v), rt
79.3 ± 0.8n/a
[42]
co
methanol–water
65:35 (v/v)
n/a (CE/ME)4818.75myricetin
[47]
co
ethanol–water
50:50 (v/v), 50 °C
UAE,
1 h
61.25 ± 1.5010.03 ± 0.31n/a
[47]
no
ethanol–water
50:50 (v/v), 50 °C
UAE,
1 h
8.84 ± 0.103.20 ± 0.07n/a
[71]
no
ethanol–water
50:50 (v/v), 50 °C
UAE,
1 h
91.9 ± 3.217.34-vinyl guaiacol
[55]
no
water,
70 °C
UAE (40 kHz), 15 min203.58 ± 3.25n/an/a
ethanol–water
50:50 (v/v), 70 °C
UAE (40 kHz), 15 min226.79 ± 1.54n/a
waterMAE (850 W), 30 min216.47 ± 1.19n/a
ethanol–water
70:30 (v/v)
MAE (650 W), 18 min321.08 ± 3.23n/a
[56]
To
waterMAE (850 W), 30 min382.26 ± 0.97n/an/a
ethanol–water
70:30 (v/v)
MAE (650 W), 18 min403.73 ± 7.35n/a
[57]
oD
waterUAE (40 kHz),
15 min
ca 290 cn/agallic acid, catechin,
vanillic acid
ethanol–water
70:30 (v/v)
UAE (40 kHz),
15 min
ca 285 cn/a
waterMAE (850 W), 30 minca 315 cn/a
ethanol–water
70:30 (v/v)
MAE (650 W), 18 minca 370 cn/a
[57]
Ho
waterUAE (40 kHz),
15 min
ca 310 cn/acaffeic acid, catechin,
gallic acid
ethanol–water
70:30 (v/v)
UAE (40 kHz),
15 min
ca 330 cn/a
waterMAE (850 W), 30 minca 350 cn/a
ethanol–water
70:30 (v/v)
MAE (650 W), 18 minca 355 cn/a
* Compared oak species (abbreviations in Ref section): co—common oak (Quercus robur); no—northern red oak (Quercus rubra); To—Turkey oak (Quercus cerris); oD—oak of Daléchamp (Quercus dalechampii); Ho—Hungarian oak (Quercus frainetto). a Types of extractions: UAE—ultrasound-assisted ex.; CE—conventional ex.; MAE—microwave-assisted ex. b TPC—total phenolic content, expressed as mg gallic acid equivalent per g dry weight of bark (mg GAE/g DWB); n/a—not available. c Value obtained from an inaccurate graph. d In order from top to bottom: whole bark from the continental (cont.) zone, outer bark from the cont. zone, whole bark from the coastal (coa.) zone, outer bark from the coa. zone; rt—room temperature.

4.5. Evaluation of Bark Extracts from Other Deciduous Trees of North Temperate Zone (Table 5)

The highest comparable (expressed as mg GAE per gram of dry weight of bark) TPC value (174.25 ± 16.95) was obtained for the sweet chestnut bark extract obtained by UAE with a horn sonicator (15 min) [58]. The TPC of the sweet chestnut bark bath/UAE-heated extract was significantly lower (58.87 ± 2.24 mg GAE/g DWB), which may suggest that the temperature (at least 50 °C) is more harmful to these polyphenols than to European larch polyphenols.
A complex study of European beech extracts obtained by MAE was carried out, taking into account different microwave power values [38]. Water MAE beech extracts had the highest TPC values for the shortest extraction time (for all microwave power values). For equal-volume water–ethanol MAE beech extracts, a reverse trend was observed. The 80% ethanol MAE beech extracts were not clearly correlated with microwave power and time. Catechin and vanillic acid were identified in beech bark in two independent studies. All structures of abundant compounds mentioned in Table 5 (except unspecified eleutheroside) are presented in Scheme 5.
Table 5. Phytochemical extraction from bark of other deciduous trees of north temperate zone *—comparison of selected recent studies (since 2018).
Table 5. Phytochemical extraction from bark of other deciduous trees of north temperate zone *—comparison of selected recent studies (since 2018).
Ref.Solvent Type,
Temperature of
Extraction
Extraction Method a,
Time
TPC (mg GAE
/g DWB) b
Extraction Yield
(% DW)
Identified Abundant Compounds
[44]
at
subcritical water, 150 °C SFE-UAE c (40 bar, 3 Hz), 40 min 31.47 ± 1.86n/agallic acid, catechin, benzoic acid, guaiacol
[58]
wc
ethanol–water
80:20 (v/v)
UAE (horn),
15 min
112.88 ± 17.27n/a-O-hexosides of: (luteolin, apigenin, daidzein, taxifolin, kaempferol), scopolin
[58]
sc
ethanol–water
80:20 (v/v)
UAE (horn),
15 min
174.25 ± 16.95n/atrigalloyl-HHDP-glucose, quinic acid, vescalagin
[47]
sc
ethanol–water
50:50 (v/v), 50 °C
UAE, 1 h58.87 ± 2.249.27 ± 0.18n/a
[42]
Ob
methanol–water
65:35 (v/v)
– (CE/ME)42.0415.50n/a
[67]
Eb
distilled water,
85–90 °C
CE (water bath with shaking),
45 min
22.95 ± 0.07n/acatechin, vanillic acid, taxifolin, syringin
[38]
Eb
waterMAE (300 c, 450 d, 600 e,
800 f W),
2, 3, 4 min
47.44–51.53 c
48.19–56.79 d
52.79–59.10 e
55.68–72.31 f
n/acatechin, vanillic acid,
(-)-epicatechin
ethanol–water 50:50 (v/v)67.27–76.46 c
66.43–77.53 d
71.91–72.43 e
72.46–73.32 f
n/acatechin,
(-)-epicatechin,
vanillic acid
ethanol–water
80:20 (v/v)
61.87–64.77 c
66.00–67.86 d
69.81–70.95 e
64.12–66.07 f
n/acatechin,
(-)-epicatechin,
vanillic acid
[42]
sp
methanol–water
65:35 (v/v)
–(CE/ME)10019.43myricetin, eleutheroside, quercetin
[71]
Cp
ethanol–water
50:50 (v/v), 50 °C
UAE, 1 h56 ± 0.416.84-vinyl guaiacol,
4-methylguaiacol
[49]
al
distilled water,
60 °C
CE (continuous stirring),
1 h
29.00 ± 5.33
13.42 ± 1.41
12.18 ± 1.53
4.92 ± 0.51 g
n/an/a
[70]
we
water, 95 °CCE (with ice on lid), 2 h407.05 h5.18n/a
[47]
ca
ethanol–water
50:50 (v/v), 50 °C
UAE, 1 h49.91 ± 1.6315.77 ± 0.14n/a
[47]
Ib
ethanol–water
50:50 (v/v), 50 °C
UAE, 1 h21.99 ± 0.155.09 ± 0.06n/a
[73]
sb
methanol, 50 °CUAE (bath, 35 kHz), 3 h79.4317.74 ± 1.64n/a
[47]
bl
ethanol–water
50:50 (v/v), 50 °C
UAE, 1 h5.49 ± 0.183.08 ± 0.18n/a
[71]
bl
ethanol–water
50:50 (v/v), 50 °C
UAE, 1 h25.3 ± 0.39.54-vinyl guaiacol
[71]
ww
ethanol–water
50:50 (v/v), 50 °C
UAE, 1 h102.6 ± 4.223.84-vinyl guaiacol, phenol
[62]
ww
methanol–water
70:30 (v/v), 40 °C
UAE (with chloroformpre-extraction), 30 min23.30 ± 0.17 k34.6caffeoyl hexose, A-type procyanidin dimers, caffeoyl hexose-deoxyhexoside,
(−)-epicatechin,
[63]
sm
acetone, rtCE
(stirring),
6 h
190 ± 10
inner bark (IB)
292.67 ± 11.02
outer bark (OB) h
6.06 ± 0.89
(IB)
7.82 ± 0.33
(OB)
IB: caffeine, p-hydroxybenzoic acid, palmitic acid
OB: gallic acid, p-hydroxybenzoic acid, phthalic acid
* Compared deciduous trees of north temperate zone (abbreviations in Ref. section): at—apple tree (Malus domestica); wc—wild cherry (Prunus avium); sc—sweet chestnut (Castanea sativa); Ob—Oriental beech (Fagus orientalis); Eb—European beech (Fagus sylvatica); sp—silver poplar (Populus alba); Cp—Canadian poplar (Populus × canadensis); al—alder (Alnus glutinosa); we—wych elm (Ulmus glabra); ca—common ash (Fraxinus excelsior); Ib—Iberian white birch (Betula celtiberica); sb—silver birch (Betula pendula); bl—black locust (Robinia pseudoacacia); ww—white willow (Salix alba); sm—sugar maple (Acer saccharum). a Types of extractions: SFE—supercritical fluid ex.; UAE—ultrasound-assisted ex.; CE—conventional ex.; MAE—microwave-assisted ex. b TPC—total phenolic content, expressed as mg gallic acid equivalent per g dry weight of bark (mg GAE/g DWB); n/a— not available. c–f Assignment of MAE power for TPC value ranges; rt—room temperature. g In order from top to bottom: whole bark from the continental (cont.) zone, outer bark from the cont. zone, whole bark from the coastal (coa.) zone, outer bark from the coa. zone. h As mg GAE/g (not specified what material a gram is), incomparable. k As mg/g DW (not as GAE, obtained by UPLC-PDA-Q/TOF-MS), incomparable.

4.6. Evaluation of Tropical and Subtropical Tree Bark Extracts (Table 6)

The highest TPC value (373 ± 4.2 mg GAE/g DWB) was obtained for goran (Ceriops decandra). Most of the presented studies cannot be expressed as mg GAE per dry weight of bark and, due to this, cannot be compared to each other. Slight differences were observed between the TPCs of extracts with different fruit colours [60] and human effects on plant growth [54], but more research is needed. Phytochemical screening was the dominant qualitative method for studies in this table, and only groups of compounds were identified.
Table 6. Phytochemical extraction from bark of tropical and subtropical trees *—comparison of selected recent studies (since 2018).
Table 6. Phytochemical extraction from bark of tropical and subtropical trees *—comparison of selected recent studies (since 2018).
Ref.Solvent Type,
Temperature of
Extraction
Extraction Method a,
Time
TPC (mg GAE
/g DWB) b
Extraction Yield
(% DW)
Identified Abundant
Compounds c
[59]
ma
ethanol, ~78 °CSoxE
(3 times
refluxed)
60.25n/atannins, flavonoids, carbohydrates, steroids
ethyl acetate, ~77 °C63.00n/atannins, flavonoids
chloroform, ~61 °C36.25n/atannins, flavonoids, steroids
petroleum ether,
42–62 °C
29.75n/atannins, flavonoids
[60]
wa
methanolME0.34 (pink)
0.34 (red) d
7.58 (pink)
6.48 (red) d
phenols, flavonoids, saponins, triterpenoids, alkaloids
ethyl acetate0.26 (pink)
0.30 (red) d
2.90 (pink)
2.44 (red) d
n-hexane0.17 (pink)
0.21 (red) d
0.52 (pink)
0.54 (red) d
[52]
em
methanol, rtME287.16 ± 2.14 fn/aphenols, steroids,
alkaloids
methanol–
ethyl acetate e
362.88 ± 1.89 f
methanol–
n-hexane e
15.47 ± 0.38 f
[64]
re
methanolME366.43 ± 11.52 f4.24alkaloids, steroids, tannins, xanthones, reducing sugars
[53]
ka
n-hexaneME187.37 ± 0.06 gn/axanthones
dichloromethane127.84 ± 0.05 g
ethyl acetate116.65 ± 0.06 g
methanol73.40 ± 0.11 g
[65]
kt
ethanolUAE451.07 ± 3.35 gn/an/a
[65]
ct
ethanolUAE327.60 ± 2.79 gn/an/a
[65]
jf
ethanolUAE90.33 ± 0.23 gn/an/a
[74]
go
water, 80 °CCE, 3 h373 ± 4.2n/an/a
[54]
za
methanolSoxE,
3 days
185.15 ± 1.22 g (wild)
171.13 ± 6.73 g (cultivated)
n/an/a
* Compared tropical and subtropical trees (abbreviations in Ref. section): ma—mangrove apple (Sonneratia caseolaris); wa—water apple (Syzygium aqueum); em—Elaeocarpus mastersii King; re—resin tree (Dipterocarpus alatus); ka—kandis (Garcinia forbesii); kt—kusum tree (Schleicera oleosa); ct—cashew tree (Anacardium occidentale); jf—jackfruit (Artocarpus heterophyllus); go—goran (Ceriops decandra); za—Zanthoxylum armatum. a Types of extractions: UAE—ultrasound-assisted ex.; SoxE—Soxhlet ex.; ME—ex. by maceration; CE—conventional ex. b TPC—total phenolic content, expressed as mg gallic acid equivalent per g dry weight of bark (mg GAE/g DWB); n/a—not available. c Groups of compounds (phytochemical screening). d Colour of fruits of water apple; rt—room temperature. e Concentrated methanol extract was partitioned using n-hexane and ethyl acetate. f As mg GAE/g DW (not specified DW—extract/bark), incomparable. g Comparable only with others in the same study, expressed as mg GAE per dry weight of extract (cannot be converted to mg GAE per dry weight of bark without extraction yield).

4.7. Summary of Survey

Although the identification of phytochemicals was limited to the standards selected by the researchers, some trends could be observed. Polyphenols and their derivatives were the most abundant type of extractives for all groups of trees. Other common compounds were guaiacol and their derivatives, stilbenes and their derivatives, and simple phenolic acids. All the above-mentioned compounds are aromatic, but cycloaliphatic extractives (quinic acid, dehydroabietic acid) and aliphatic extractives (palmitic acid) are also present. The majority of the extractives have at least two hydroxyl groups, which makes them interesting in terms of polymer synthesis. This means that bark extracts, as a relatively uniform mixture, can be a source of substances with similar properties, which have potential as commercial substrates for the polymer industry.

5. Application of Bark Extracts in Thermosetting Polymers

5.1. Types of Thermosetting Polymers

Thermosetting polymers, also called resins, are a group of synthetic materials that include polyurethanes (PUR), epoxy and polyester resins. Also, formaldehyde-based resins, such as phenol–formaldehyde (PF), urea–formaldehyde (UF) and melamine–urea–formaldehyde (MUF), constitute a group classified as thermosets. Other polymer materials that have a dense, crosslinked structure, a high molar mass and good mechanical properties can be called resins.

5.2. Formaldehyde-Based Resins and Alternative Adhesives

PF, UF and MUF resins, which are popular adhesives for plywood and veneer, are rather obvious types of thermosets wherein bark phytochemicals, especially polyphenols and tannins, can be used. The main drawback of formaldehyde-based resins is formaldehyde emission, which can be reduced by modifications or removed for formaldehyde-free alternatives (e.g., amine-based). Bark phytochemicals can be useful in both solutions.
Hajriani et al. paid attention to the importance of the optimal formulation of tannin from the bark of Merkus pine (Pinus merkusii), formaldehyde and resorcinol mixtures for PF adhesives [69]. They observed a correlation between the tannin structure with the presence of resorcinol and a relatively low (approx. 5%) formaldehyde demand, which should result in low emissions. Hendrik et al. proposed a similar formaldehyde–resorcinol-based PF adhesive with tannin bark extract from a fast-growing tree—mangium (Acacia mangium). The obtained resin partially meets the requirements of Japanese standards (JAS 234, in terms of moduli of elasticity and rupture) and also shows very low formaldehyde emission [68]. Commercial PF was copolymerised with tannins from black wattle (Acacia mearnsii) bark in different ratios (20%, 30% and 40% of PF resin solids), and the obtained composites were characterised by better thermal stability and faster curing with lower shear strength [75].
The UF resin was modified with small amounts (2.5% and 5%) of bark extracts of maritime pine (Pinus pinaster) and wych elm (Ulmus glabra) to obtain phenol–urea–formaldehyde (PUF) resins [70]. PUF adhesive with 2.5% maritime pine extract had improved bonding shear strength (more than 20% higher) for pine plywood. In addition, reduced formaldehyde emission was observed for all compositions with wych elm bark extract (from 1 mg to 5.5 mg per 100 g of oven-dried panel) and maritime pine extract (from 3 mg to 5 mg per 100 g of oven-dried panel). Tannins extracted from the bark of goran (Ceriops Decandra) were also added to commercial UF resin and compared to commercial UF and a tannin-only adhesive [74]. The PUF obtained with 25% tannin content was optimal in terms of water resistance and mechanical and adhesive properties. Also, the addition (10%) of finely ground beech bark as a filler for a UF adhesive for plywood provided lower formaldehyde emission and better mechanical properties, which are the result of the presence of bark extractives in the UF adhesive [76]. However, the higher content of filler (15–20%) was too viscous to prepare with laboratory equipment.
Janceva et al. synthesised adhesives from bark extracts (with amounts of condensed tannins (CTs)) of grey alder (Alnus incana) and black alder (Alnus glutinosa) combined with polyethyleneimine (CTs-PEI) [77]. CTs-PEI was also mixed with ultra-low-emission PF resin (ULEFR). These mixed adhesives (40–60% CTs-PEI substitution) show comparable mechanical properties (modulus of elasticity, shear strength) to clear ULEFR. In the case of CTs-PF resin, formaldehyde emissions were two times lower compared to those of conventional PF adhesives. Another formaldehyde-free alternative adhesive based on Monterey pine (Pinus radiata) bark extract and hexamethylenetetramine (7%) with a low solid content (30%) had properties similar to those of commercial PF resin [78]. Garcia et al. conducted a systematic study of polyphenolic resins synthesised with commercial CTs of maritime pine bark (P. pinaster) and eleven different aldehydes as formaldehyde alternatives [79]. They observed an increase in the Tg and bulk density of resins as a function of the aldehyde chain length. The type of aldehyde chain (aliphatic or aromatic, functionality, length) may also be useful for tailoring the properties of resins (e.g., lower ratio of carbon and hydrogen to other elements improves fire resistance). All collected data are summarised in Table 7.
Table 7. Formaldehyde-based resins and alternative adhesives—summary.
Table 7. Formaldehyde-based resins and alternative adhesives—summary.
Polymer MatrixExtract SourceExtract ContentPropertiesRef.Comments
PFmerkus pine barkover 50% (not specified)liquid adhesive (solid content from 14% to 17%) with Stiasny number from 50% to 83% and low formaldehyde demand (5%)[69]only formulations of tannins with resorcinol and formaldehyde were tested, no mechanical tests of adhesive
PFmangium barkca. 80%laminate with mangium wood partially meets the JAS 234 standard—slightly lower moduli of elasticity and rupture and good level of formaldehyde emission (very low)[68]-
PFblack wattle bark20%, 30% and 40%copolymers with better thermal stability, faster curing and lower shear strength than commercial PF[75]-
PFmaritime pine bark40%, 50%, 67%Tg and bulk density correlation with aldehyde chain length and functionality and unsaturation[78]PF
PUFmaritime pine bark2.5% and 5%improved bonding shear strength (more than 20% higher), reduced formaldehyde emission[70]tested as adhesive for pine plywood
PUFwych elm bark2.5% and 5%reduced formaldehyde emission[70]tested as adhesive for pine plywood
PUFgoran bark 25%reduced moisture content, comparable water absorption and better mechanical properties than commercial UF[74]all properties for particleboards; mechanical tests: tensile strength, moduli of elasticity and rupture
UFbeech bark (passive extraction)10%reduced formaldehyde emission, equal or higher mechanical properties (modulus of rupture, thickness swelling, bonding quality) compared to plywood with commercial UF[76]tested as adhesive for beech plywood
ULEFRgrey alder and black alder bark40–60%comparable mechanical properties (modulus of elasticity, shear strength) to clear ULEFR, reduced formaldehyde emission[77]tested as adhesive for birch plywood and pine wood particleboards
amine-based PF alternativeMonterey pine bark90–95%liquid adhesive (solid content—30%) with properties similar to those of commercial PF resins, reduced formaldehyde emission)[78]-

5.3. Polyurethanes

Polyphenols, tannins and lignin [80] are potential substrates for the synthesis of PUR as polyol alternatives due to the presence of hydroxyl groups, which react with diisocyanates or their less toxic derivatives (NIPU—non-isocyanate polyurethane).
PUR based on tannins extracted from mangium (Acacia mangium) bark and polymeric diphenylmethane diisocyanate (pMDI) was used in the impregnation process of ramie fibres, enhancing their thermal and mechanical properties [81]. Mangium tannin-based PUR for ramie fibre impregnation was compared with the isocyanate-free version of PUR from this extract (NIPU). NIPU was prepared in a reaction with dimethyl carbonate and hexamethylenediamine instead of pMDI [82,83]. Both types of bio-based PUR improved the mechanical parameters of the fibres (isocyanate-based—higher modulus of elasticity; isocyanate-free—higher tensile strength). PUR based on pMDI provided better thermal properties. The most recent study focused on the optimisation of mangium tannin-based NIPU [83].
Higher degradation temperatures were also observed for PUR foam with added hydroxybutylated tannin from Monterey pine (Pinus radiata) bark [84]. Hydroxyalkylation was selected for the modification of tannins, which improved their viscosity to a level similar to that of commercial polyols in PUR foams. However, the addition of these homoalkylated tannins still resulted in foam brittleness and a higher density. D’Souza et al. observed similar characteristics (higher degradation temperatures and density, brittleness) of foams made by adding (about 13%) liquefied lodgepole pine (Pinus contorta) bark infested with mountain pine beetle [85]. Liquefaction was conducted at two temperatures (90 °C and 130 °C), and different product profiles were obtained. Also, the produced PUR foams had different properties. The 90 °C bark-based foam contained secondary hydroxyls (from sugars) and a foaming behaviour more similar to that of polypropylene glycolfoam based on glycerol (PPG-G). On the other hand, a large amount of low-molecular-weight compounds was disadvantageous to foam properties. The 130 °C bark-based foam (high-molecular-weight, well-functionalised compounds) had a lower density and better elastic properties (elastic modulus and compression strength).
Phytochemicals can also be used as natural photostabilisers for PUR-based coatings. Bark extracts from Chinese fir (Cunninghamia lanceolata) were added to a UV-curable waterborne polyurethane-acrylate (PUA) coating, and a relationship between higher TPC and better photostabilisation efficiency was observed, especially for the ethanolic extract [72]. For mimosa (Accacia mollissima) bark extract added to PUR varnish, a decrease in surface glossiness was observed, with the exception of the mixture with 10% bark extract, which provides better impregnation and increased glossiness than the tested commercial impregnating agent [86]. Bio-based aromatic diisocyanates were also obtained from guaiacol and vanillyl alcohol, both available in bark extracts [87]. All collected data about polyurethanes are summarised in Table 8.
Table 8. Polyurethanes—summary.
Table 8. Polyurethanes—summary.
Polymer MatrixExtract
Source
Extract ContentPropertiesRef.Comments
PUAChinese fir bark2%coating: increased
photostability
[72]better for higher TPC
PURmimosa bark10%coating: better impregnation and glossiness[86]other extract contents were suboptimal
PURMonterey pine bark25%, 50%, 75%, 100%foam: improved thermal stability, increased strength and brittleness, deformed cellular structure *[84]* only for higher loadings
PURlodgepole pine barkca 13%foam: higher degradation temperature and density, increased brittleness[85]-
PURmangium barknon-specifiedimproved thermal stability and crystallinity[81]used for impregnation of ramie fibres, different times of impregnation
PURmangium barknon-specifiedimproved thermal stability,
tensile strength and elastic modulus
[82]used for impregnation of ramie fibres
NIPUmangium bark30.7% and 44.4%[82]
NIPUmangium bark30.7%as above, optimisation[83]used for impregnation of ramie fibres

5.4. Polyester Resins

The presence of hydroxyl and carboxylic groups in phytochemicals can be attractive for the synthesis of polyesters. Tannins and other naturally occurring biopolyesters can be applied without other synthetic substrates.
Han et al. selected salicylic acid (present in willow (Salix sp.) bark) as a poly(salicylate) homopolymer substrate with a narrow molecular weight distribution and a high glass transition temperature (Tg) [88]. They obtained a six-membered cyclic o-carboxyanhydride of salicylic acid as the intermediate product, which was polymerised via ring-opening polymerisation. Lang et al. proposed a biopolyester adhesive based on torrefied bark of silver birch (Betula pendula) as an alternative to hot-melt glue or a tackifying agent [89]. Birch bark is also a source of betulin, which reacts with aliphatic acid dichlorides (from 5 to 8 carbons in the methylene chain) and gives a biopolyester with a Tg comparable to that of polycarbonates (150–165 °C) [90]. Also, thermal treatment of such biopolyesters (2 h at 250 °C) provides better chemical resistance to tetrahydrofuran (THF), CHCl3 and N-methylpyrrolidone. Moreover, the polycondensation product of betulin and sebacoyl dichloride is a transparent biopolyester. Furthermore, suberin fatty acid hydrolysates (from the outer bark of silver birch) crosslinked with maleic anhydride and octadecyltrichlorosilane can constitute competitive biopolyester coatings for the wood industry, providing improved hydrophobicity and stable anti-ageing properties [91]. Gosecki et al. also proposed a vitrimer based on birch suberin monomers and commercial polyols in the presence of a tin catalyst [92]. The obtained recyclable polymers show good hydrolytic stability and creep resistance. Cho et al. isolated depolymerised suberin derivatives from the powdered bark of cork oak (Quercus suber) and obtained syntactic biopolyester foams by polycondensation with glycerol (5.77 wt%) [93]. These foams were similar to cork in terms of density and porosity but had a different chemical structure (higher content of suberin), which caused different viscoelastic properties (higher tan δ and a more viscous character).
Poly(δ-decalactone) and poly(δ-dodecalactone), two other types of polyesters, can be synthesised via the ring-opening polymerisation of δ-decalactone and δ-dodecalactone, respectively [94]. Both lactones were reported as oil ingredients harvested from the bark of the Massoia tree (Cryptocarya massoia) [95]. Garcia et al. extracted polyflavonoids from Monterey pine (Pinus radiata) bark and proposed their synthesis with poly(lactic acid) (PLA) [96]. The obtained polyester blends were compared to their modified versions, where hydroxypropylated (propylene oxide was used) polyflavonoids were synthesised with PLA instead of raw ones. In addition, blends with poly(ethylene glycol) (PEG) were obtained. All types of blends induced the crystallisation of PLA, but the PLA–polyflavonoid–PEG blends also had reduced transitional temperatures because of the breakage of the intermolecular hydrogen bonds and increased chain mobility. Furthermore, the bark tannins of Monterey pine (Pinus radiata) were hydroxypropylated (also with propylene oxide) by Bridson et al. [97]. Hydroxypropylated tannins were blended with PLA using the melt-spinning method. Surprisingly, a more amorphous character was observed for blends with hydroxypropylated bark tannins (as opposed to the conclusions of Garcia et al. [96]). These differences may be related to the type of selected PLA, the extraction parameters and the blending procedure. This suggests that the impact of polyphenols on PLA blends is more complex and needs further investigation. All collected data are summarised in Table 9.
Table 9. Polyester resins—summary.
Table 9. Polyester resins—summary.
Polymer MatrixExtract SourceExtract ContentPropertiesRef.Comments
polyesterwillow barkalmost 100%high Tg, rapid polymerisation with a narrow distribution of molecular weights[88]-
polyestertorrefied bark of silver birch100%high adhesion—alternative to hot-melt glue or tackifying agent[89]-
polyestersilver birch bark52%Tg at 150–165 °C, good chemical resistance[90]made from unsaturated betulin with sebacoyl chloride and pyridine
polyestersilver birch bark70%improved hydrophobicity and stable anti-ageing properties[91]made from suberin fatty acids with maleic anhydride
polyestersilver birch barknon-specifiedgood hydrolytic stability and creep resistance[92]made from suberin with polyol and tin catalyst
polyestercork oak barknon-specifiedfoams similar to cork[93]made from depolymerised suberin with glycerol and bismuth catalyst
poly(lactic acid) blendMonterey pine bark10%, 20%, 30%, 40%improved processability of PLA, more crystalline character[96]also hydroxypropylated extract was tested
poly(lactic acid) blendMonterey pine bark25%amorphous character[97]hydroxypropylated extract, obtained by melt-spinning

5.5. Epoxy Resins

The hydroxyl groups of phytochemicals can be epoxidised with chlorohydrin to obtain epoxy compounds for resin synthesis [98] or react with an epoxy ring to form epoxy resins (in the presence of a specific catalyst). Also, the reaction between carboxyl groups and epoxy rings is a common synthesis method of epoxy resins.
Bridson et al. [99] confirmed the reactivity of the Monterey pine (Pinus radiata) bark tannin with simple aliphatic epoxides (propylene oxide, butylene oxide and hexylene oxide) in the presence of triethylamine (hydroxylation). The hydroxylated products had reduced viscosity compared to natural tannins, but viscosity and Tg increased with the increasing aliphatic chain length of the epoxide. Extractives (water-soluble fraction and fraction insoluble in water) from Monterey pine (Pinus radiata) bark were added to commercial epoxy-polyamide resin as copolymers [100]. Both fractions are competitive as corrosion inhibitors (the water-soluble fraction performs better). This effect is probably the result of the formation of complexes between Fe3+ ions and adjacent hydroxyl groups of the B-ring of polyphenols. Epoxidised extracts from the bark of various trees were reported recently. Shnawa prepared tannin-based epoxy resin from eucalyptus (Eucalyptus sp.) bark and compared it with a commercial product [101]. The crosslinking process with different ratios of the amine as a curing agent was more sluggish for tannin-based epoxy as a result of the steric effects of the tannin structure, according to the author. Despite this, the commercial resin also had reduced viscosity (presence of 1,6-hexanediol diglycidyl ether, 10–30%), which could also improve reactivity. Only blends of commercial epoxy with up to 20% tannin-based epoxy maintained curing properties similar to those of neat commercial epoxy. Bog-myrtle (Myrica gale) bark extract was another mixture of tannins functionalised with epoxy groups. The obtained epoxy was copolymerised with chitosan and applied as a paper coating that provides resistance to oil and water [102].
Another solution was the addition of the aforementioned epoxy to a commercial one as a strengthening agent (5%) [103]. It was successful: a significant improvement in flexural, tensile and impact strengths was obtained with the maintenance of chemical and thermal resistance. Chen et al. epoxidised eucommia ulmoides gum sourced from Eucommia ulmoides bark and added it to the epoxy coating as a nanofiller [104]. Composites with up to 1% of such filler had improved tensile strength and enhanced corrosion performance as a result of a higher crosslink density.
An interesting type of compound is magnolol extracted from houpu magnolia (Magnolia officinalis) bark, which can be turned into epoxy by the epoxidation of allyl groups (Prilezhaev reaction with 3-chloroperoxybenzoic acid in dichloromethane solution, without catalyst) instead of hydroxyl groups [105]. Unreacted hydroxyl groups and created epoxy groups reveal self-curing behaviour above 120 °C. The self-curing activation energy was relatively low (94.8 kJ/mol) compared to the activation energy (77.0 kJ/mol) of the curing reaction of this epoxy with an amine (4,4’-methylenedianiline), which obviously acts as a catalyst. Furthermore, self-cured magnolol-based epoxy resin had lower flammability compared to other tested systems. In addition, both magnolol-based epoxy resins had antibacterial properties. The commercial availability of magnolol and the unique properties of magnolol-based epoxy have resulted in new research articles on the topic. Cao et al. tested ester functional groups (instead of hydroxyl groups) for magnolol-based epoxy as an alternative [106]. They observed the necessity of an amine hardener as a catalyst for self-curing (only 0.5% or 1%), but the obtained thermosets had higher stiffness and toughness compared to commercial epoxy resins based on the diglycidyl ester of bisphenol A (DGEBA). Zhang et al. proposed magnolol-based epoxy as a reinforcement for bio-based aerogel made of a chitosan skeleton [107]. Wei et al. synthesised a reactive flame-retardant—diglycidyl ether of magnolol phosphine oxide—and added it to commercial epoxy [108]. Epoxy systems with 10% and 15% flame-retardant had significantly lower heat production and slightly increased smoke production in the cone calorimeter test. Also, the limited oxygen index in the UL-94 test was measured (37.0% and 41.5%, respectively, both V-0 rating). All data collected on epoxies are summarised in Table 10.
Table 10. Epoxy resins—summary.
Table 10. Epoxy resins—summary.
Polymer MatrixExtract
Source
Extract ContentPropertiesRef.Comments
epoxyMonterey pine bark22%, 28%, 33%, 50%oligomers with various viscosities and Tg, affected by chain length[99]reactivity tests
epoxyMonterey pine barknon-specifiedcompetitive as corrosion inhibitor[100]-
epoxyeucalyptus bark20%, 40%, 60%slightly faster curing for 20% content[101]bark extract was epoxidised
epoxybog-myrtle bark non-specifiedoil and water resistance[102]as paper coating
epoxybog-myrtle bark 5%significant improvement in flexural, tensile and impact strengths[103]as strengthening agent
epoxyeucommia ulmoides bark1%improved tensile strength and enhanced corrosion performance [104]as nanofiller
epoxyhoupu magnolia barkover 90%lower flammability and antibacterial properties[105]self-curing ability
epoxyhoupu magnolia barkover 90%higher stiffness and toughness compared to DGEBA[106]added an amine catalyst for self-curing
epoxy-chitosan copolymerhoupu magnolia bark9%, 16.6%, 23.1%, 28.6%, 33.3%improved flame retardancy[107]as hard segments for chitosan-based aerogel
epoxyhoupu magnolia bark10%, 15%significantly lower heat production and slightly increased smoke production, V-0 rating for UL-94 test[108]as a reactive flame-retardant

5.6. Other Thermosets

Magnolol can also be a precursor for other thermosets. Magnolol and honokiol functionalised with 4-nitrophthalonitrile (hydroxyl group substitution) are able to form phthalonitrile thermosets via self-curing reactions [109]. These resins exhibit high thermal stability (Tg > 500 °C) compared to similar resins obtained from petroleum-based precursors. The allyl groups of magnolol are also useful for thiol-ene click chemistry. Adhesive synthesised from magnolol and a tetrafunctional thiol—pentaerythritol tetra-(3-mercaptopropionate)—can be used to bond different types of materials (PVC, steel, wood, glass), even in a wet environment (adhesion strength greater than 7.5 MPa and 6.0 MPa after immersion in hot water for 3 h) [110]. Furthermore, an antibacterial effect was also observed. Weems et al. tested five different terpenes with a multifunctional thiol to obtain thermosetting resins via thiol-ene photopolymerisation [111]. They obtained promising materials for 3D printing with mechanical properties more similar to those of elastomers and thermoplastics. Recent studies proposed the biomimetic crosslinking of silicone elastomers by polyphenols with an allyl group (i.e., eugenol) in the presence of a platinum catalyst [112], which can also be attractive for the synthesis of thermoset–silicone copolymers. Also, combustion behaviour studies of cork and phloem from the bark of different trees [113] suggest that bark after extraction, without relatively flammable extractives, can be an interesting ecological alternative for flame-retardant fillers. The data collected showing the application of phytochemicals in other resins are summarised in Table 11.
Table 11. Other resins—summary.
Table 11. Other resins—summary.
Polymer
Matrix
Extract
Source
Extract
Content
PropertiesRef.
phthalonitrile resinhoupu magnolia bark42.80%high thermal stability (Tg > 500 °C)[109]
thiol-ene resinhoupu magnolia bark35.3%, 45.0%, 52.2%, 59.2%, 68.6%good adhesive in a wet environment, antibacterial properties[110]
thiol-ene resinlimonene36.2%3D printing material with mechanical properties similar to those of elastomers and thermoplastics [111]

6. Discussion and Perspectives

The evaluation of data collected for six groups of trees revealed an interesting relationship between the extraction time and the type of solvent for some types of bark trees. Relatively short extraction times and a temperature close to the boiling point of the solvent are favourable, although each type of bark has unique and distinctive extraction needs. However, as stated by Warlo et al. [114], bark extraction should be designed for the stage of existing industrial processes for a better use of resources. The description of phytochemicals present in the extractives of different trees in this article, as well as their structures and TPC values, is an introduction to the discussion of the attractiveness of the most popular bark in terms of possibilities arising from extractable phytochemicals. Polyphenols, tannins and alcohols need complex investigation as carriers of reactive hydroxyl groups. The location of hydroxyl groups is also an important factor. Pinnataip [115] and Janceva [77] indicate that the catechol structure present in most bark compounds (only reported in this article: astringin, flavan-3-ols, taxifolin, myricetin, caffeic acid, quercetin, procyanidin, gallic acid, protocatechuic acid and selected glycosides of the listed compounds) is sensitive to oxidation and therefore reactive in the form of orthoquinone with primary and secondary amine groups. It is an important attribute for the chemistry of epoxy resins and amine-based formaldehyde-free alternatives of PF resins (Figure 4).
Also, functional groups with double bonds (vinyl, allyl, etc.) discussed in the “Other Thermosets” Section [110,111] may be interesting for the creation of a polymer network. Most phytochemicals found in bark do not have multiple easily accessible double bonds, but a relatively large group has at least one such double bond (only reported here: stilbenoids and stilbene glycosides, caffeic acid, 4-vinyl guaiacol, syringin, triterpenoids). This makes it possible to use thiol-ene chemistry as a mechanism, which may support hydroxyl-based curing reactions. Also, the Prilezhaev reaction [105] is an interesting approach to the epoxidation of groups with double bonds, the advantage of which is the availability of hydroxyl groups along with epoxy groups in one compound (enabling homopolymerisation similar to magnolol). Phytochemicals are mainly aromatic or cycloaliphatic structures and provide greater stiffness of the polymer chain and better thermal properties but may be less reactive because of steric hindrance. Furthermore, the more complex architecture of compounds such as condensed tannins (CTs) must also be taken into account. As oligomeric structures, they are convenient as a resin backbone, but the main drawback of their application is alkaline degradation resulting from the opening of the C15 flavonoid unit [79].

7. Conclusions

All of the advantages and disadvantages mentioned above should be helpful in selecting the most suitable compounds for each specific application. However, selection usually requires separation processes, because compounds are generally components of the mixture. This approach should be considered as long as the separation is economically justified. The summary of the presented polymer solutions with phytochemicals reveals two main application methods. The first method—applying them as additives to commercial thermosets—may be convergent with a selective approach if the addition of phytochemical provides sufficient benefits. However, a small amount of phytochemicals does not change a petroleum-based resin to a bio-based one (even if marketing specialists say so). It is still only a palliative measure for the issues mentioned in the “Introduction” Section. The second method, the bio-based thermoset (or the main part of it), as an alternative to the commercial one, should provide a better solution for managing bio-waste and replacing fossil-based chemicals. Unfortunately, a selective approach may result in higher production costs for such alternatives and lower competitiveness than existing solutions (only significantly better properties will justify the “bio-choice”). To avoid expensive separation stages, a more holistic approach is needed. The selection of phytochemicals should be less rigorous and more focused on groups of compounds having common characteristics that fit the desired application. TPC can be a useful standardisation tool for such extracts as a relatively fast and simple method (but a unified expression of the TPC values should be used, as stated in Section 4). Exclusion should be taken into account only for compounds that significantly worsen the properties of the final product. However, the question about the competitiveness of such bio-resins compared to commercial solutions remains. Of course, none of these approaches and application methods is the most favourable. Each has advantages and disadvantages that must be considered when designing new thermoset materials.

Author Contributions

Conceptualisation, T.S.; writing—original draft preparation, T.S.; writing—review and editing, M.M.; supervision, M.M.; funding acquisition, M.M. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by the University of Lodz (Grant No. B2411100000050.01).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

The data presented in this study are mostly openly available. DOI addresses are provided in the References Section.

Conflicts of Interest

The authors declare no conflicts of interest.

References

  1. Forest Products 2020; FAO: Rome, Italy, 2022; ISBN 978-92-5-137395-8.
  2. Lu, W.; Sibley, J.L.; Gilliam, C.H.; Bannon, J.S.; Zhang, Y. Estimation of U.S. Bark Generation and Implications for Horticultural Industries 1. J. Environ. Hort. 2006, 24, 29–34. [Google Scholar] [CrossRef]
  3. Corral-Rivas, J.; Vega-Nieva, D.; Rodríguez-Soalleiro, R.; López-Sánchez, C.; Wehenkel, C.; Vargas-Larreta, B.; Álvarez-González, J.; Ruiz-González, A. Compatible System for Predicting Total and Merchantable Stem Volume over and under Bark, Branch Volume and Whole-Tree Volume of Pine Species. Forests 2017, 8, 417. [Google Scholar] [CrossRef]
  4. Neumann, M.; Lawes, M.J. Quantifying Carbon in Tree Bark: The Importance of Bark Morphology and Tree Size. Methods Ecol. Evol. 2021, 12, 646–654. [Google Scholar] [CrossRef] [PubMed]
  5. Kain, G.; Morandini, M.; Barbu, M.-C.; Petutschnigg, A.; Tippner, J. Specific Gravity of Inner and Outer Larch Bark. Forests 2020, 11, 1132. [Google Scholar] [CrossRef]
  6. Brennan, M.; Fritsch, C.; Cosgun, S.; Dumarcay, S.; Colin, F.; Gérardin, P. Yield and Compositions of Bark Phenolic Extractives from Three Commercially Significant Softwoods Show Intra- and Inter-Specific Variation. Plant Physiol. Biochem. 2020, 155, 346–356. [Google Scholar] [CrossRef] [PubMed]
  7. Truong, D.-H.; Nguyen, D.H.; Ta, N.T.A.; Bui, A.V.; Do, T.H.; Nguyen, H.C. Evaluation of the Use of Different Solvents for Phytochemical Constituents, Antioxidants, and In Vitro Anti-Inflammatory Activities of Severinia buxifolia. J. Food Qual. 2019, 2019, 8178294. [Google Scholar] [CrossRef]
  8. Haminiuk, C.W.I.; Plata-Oviedo, M.S.V.; de Mattos, G.; Carpes, S.T.; Branco, I.G. Extraction and Quantification of Phenolic Acids and Flavonols from Eugenia Pyriformis Using Different Solvents. J. Food Sci. Technol. 2014, 51, 2862–2866. [Google Scholar] [CrossRef] [PubMed]
  9. Ilek, A.; Kucza, J.; Morkisz, K. Hygroscopicity of the Bark of Selected Forest Tree Species. iForest-Biogeosci. For. 2017, 10, 220–226. [Google Scholar] [CrossRef]
  10. Feng, S.; Cheng, S.; Yuan, Z.; Leitch, M.; Xu, C. (Charles) Valorization of Bark for Chemicals and Materials: A Review. Renew. Sustain. Energy Rev. 2013, 26, 560–578. [Google Scholar] [CrossRef]
  11. Vieito, C.; Fernandes, É.; Vaz Velho, M.; Pires, P. The Effect of Different Solvents on Extraction Yield, Total Phenolic Content and Antioxidant Activity of Extracts from Pine Bark (Pinus pinaster Subsp. Atlantica). Chem. Eng. Trans. 2018, 64, 127–132. [Google Scholar]
  12. Petráš, R.; Mecko, J.; Krupová, D.; Pažitný, A. Aboveground Biomass Basic Density of Hardwoods Tree Species. Wood Res. 2020, 65, 1001–1012. [Google Scholar] [CrossRef]
  13. Plastics Europe Plastics—The Facts 2022; Plastics Europe: Brussels, Belgium, 2022.
  14. Costa, A.; Oliveira, G. Cork Oak (Quercus suber L.): A Case of Sustainable Bark Harvesting in Southern Europe. In Ecological Sustainability for Non-Timber Forest Products; Routledge: London, UK, 2015; pp. 193–212. ISBN 9781315851587. [Google Scholar]
  15. Christenhusz, M.J.M.; Byng, J.W. The Number of Known Plants Species in the World and Its Annual Increase. Phytotaxa 2016, 261, 201–217. [Google Scholar] [CrossRef]
  16. Bajpai, P. Green Chemistry and Sustainability in Pulp and Paper Industry; Springer International Publishing: Cham, Switzerland, 2015. [Google Scholar]
  17. Routa, J.; Brännström, H.; Laitila, J. Effects of Storage on Dry Matter, Energy Content and Amount of Extractives in Norway Spruce Bark. Biomass Bioenergy 2020, 143, 105821. [Google Scholar] [CrossRef]
  18. Kemppainen, K.; Siika-aho, M.; Pattathil, S.; Giovando, S.; Kruus, K. Spruce Bark as an Industrial Source of Condensed Tannins and Non-Cellulosic Sugars. Ind. Crops Prod. 2014, 52, 158–168. [Google Scholar] [CrossRef]
  19. Peeters, K.; Esakkimuthu, E.S.; Tavzes, Č.; Kramberger, K.; Miklavčič Višnjevec, A. The Potential Value of Debarking Water as a Source of Polyphenolic Compounds for the Specialty Chemicals Sector. Molecules 2023, 28, 542–554. [Google Scholar] [CrossRef] [PubMed]
  20. Multia, E. Evgen Multia Potential and Utilization of Water Extracts from Spruce Bark. Master’s Thesis, Aalto University, Espoo, Finland, 2018. [Google Scholar]
  21. Kloch, M.; Toczyłowska-Mamińska, R. Toward Optimization of Wood Industry Wastewater Treatment in Microbial Fuel Cells—Mixed Wastewaters Approach. Energies 2020, 13, 263–273. [Google Scholar] [CrossRef]
  22. Beszterda, M.; Frański, R. Seasonal Qualitative Variations of Phenolic Content in the Stem Bark of Prunus Persica Var. Nucipersica—Implication for the Use of the Bark as a Source of Bioactive Compounds. ChemistrySelect 2022, 7, e202200418. [Google Scholar] [CrossRef]
  23. Medic, A.; Zamljen, T.; Hudina, M.; Solar, A.; Veberic, R. Seasonal Variations of Naphthoquinone Contents (Juglone and Hydrojuglone Glycosides) in Juglans regia L. Sci. Hortic. 2022, 300, 111065. [Google Scholar] [CrossRef]
  24. Dou, J.; Kögler, M.; Kesari, K.K.; Pitkänen, L.; Vuorinen, T. Seasonal Dynamics in Structural Characteristics within Bark Stems of Cultivated Willow (Salix sp.) by NMR and Time-Gated Raman Spectroscopy. Green. Chem. 2023, 25, 1908–1919. [Google Scholar] [CrossRef]
  25. Gabr, S.; Nikles, S.; Pferschy Wenzig, E.M.; Ardjomand-Woelkart, K.; Hathout, R.M.; El-Ahmady, S.; Motaal, A.A.; Singab, A.; Bauer, R. Characterization and Optimization of Phenolics Extracts from Acacia Species in Relevance to Their Anti-Inflammatory Activity. Biochem. Syst. Ecol. 2018, 78, 21–30. [Google Scholar] [CrossRef]
  26. Ketkar, P.; Nayak, S.; Pai, S.; Joshi, R. Monitoring Seasonal Variation of Epicatechin and Gallic Acid in the Bark of Saraca asoca Using Reverse Phase High Performance Liquid Chromatography (RP-HPLC) Method. J. Ayurveda Integr. Med. 2015, 6, 29–34. [Google Scholar] [CrossRef] [PubMed]
  27. Jyske, T.; Brännström, H.; Halmemies, E.; Laakso, T.; Kilpeläinen, P.; Hyvönen, J.; Kärkkäinen, K.; Saranpää, P. Stilbenoids of Norway Spruce Bark: Does the Variability Caused by Raw-Material Processing Offset the Biological Variability? Biomass Convers. Biorefin. 2024, 14, 5085–5099. [Google Scholar] [CrossRef]
  28. Jyske, T.M.; Suuronen, J.P.; Pranovich, A.V.; Laakso, T.; Watanabe, U.; Kuroda, K.; Abe, H. Seasonal Variation in Formation, Structure, and Chemical Properties of Phloem in Picea Abies as Studied by Novel Microtechniques. Planta 2015, 242, 613–629. [Google Scholar] [CrossRef] [PubMed]
  29. Bello, A.; Bergmann, U.; Vepsäläinen, J.; Leiviskä, T. Effects of Tree Harvesting Time and Tannin Cold/Hot-Water Extraction Procedures on the Performance of Spruce Tannin Biocoagulant for Water Treatment. Chem. Eng. J. 2022, 449, 137809. [Google Scholar] [CrossRef]
  30. Halmemies, E.S.; Brännström, H.E.; Nurmi, J.; Läspä, O.; Alén, R. Effect of Seasonal Storage on Single-Stem Bark Extractives of Norway Spruce (Picea abies). Forests 2021, 12, 736. [Google Scholar] [CrossRef]
  31. Routa, J.; Brännström, H.; Hellström, J.; Laitila, J. Influence of Storage on the Physical and Chemical Properties of Scots Pine Bark. Bioenergy Res. 2021, 14, 575–587. [Google Scholar] [CrossRef]
  32. Zhao, T.; Kandasamy, D.; Krokene, P.; Chen, J.; Gershenzon, J.; Hammerbacher, A. Fungal Associates of the Tree-Killing Bark Beetle, Ips Typographus, Vary in Virulence, Ability to Degrade Conifer Phenolics and Influence Bark Beetle Tunneling Behavior. Fungal Ecol. 2019, 38, 71–79. [Google Scholar] [CrossRef]
  33. Hammerbacher, A.; Kandasamy, D.; Ullah, C.; Schmidt, A.; Wright, L.P.; Gershenzon, J. Flavanone-3-Hydroxylase Plays an Important Role in the Biosynthesis of Spruce Phenolic Defenses against Bark Beetles and Their Fungal Associates. Front. Plant Sci. 2019, 10, 208. [Google Scholar] [CrossRef]
  34. Preminger, M. Changes in Bark Chemistry Across Beech Bark Disease Development. Ph.D. Thesis, State University of New York College of Environmental Science and Forestry, Syracuse, NY, USA, 2019. [Google Scholar]
  35. Haz, A.; Jablonsky, M.; Majova, V.; Skulcova, A.; Strizincova, P. Comparison of Different Extraction Methods for the Extraction of Total Phenolic Compounds from Spruce Bark. J. Hyg. Eng. Des. 2018, 22, 72–75. [Google Scholar]
  36. Ferreira-Santos, P.; Genisheva, Z.; Botelho, C.; Santos, J.; Ramos, C.; Teixeira, J.A.; Rocha, C.M.R. Unravelling the Biological Potential of Pinus Pinaster Bark Extracts. Antioxidants 2020, 9, 334. [Google Scholar] [CrossRef]
  37. Venkatesan, T.; Choi, Y.-W.; Kim, Y.-K. Impact of Different Extraction Solvents on Phenolic Content and Antioxidant Potential of Pinus densiflora Bark Extract. Biomed. Res. Int. 2019, 2019, 3520675. [Google Scholar] [CrossRef] [PubMed]
  38. Tanase, C.; Mocan, A.; Coșarcă, S.; Gavan, A.; Nicolescu, A.; Gheldiu, A.-M.; Vodnar, D.C.; Muntean, D.-L.; Crișan, O. Biological and Chemical Insights of Beech (Fagus sylvatica L.) Bark: A Source of Bioactive Compounds with Functional Properties. Antioxidants 2019, 8, 417. [Google Scholar] [CrossRef] [PubMed]
  39. Brglez Mojzer, E.; Knez Hrnčič, M.; Škerget, M.; Knez, Ž.; Bren, U. Polyphenols: Extraction Methods, Antioxidative Action, Bioavailability and Anticarcinogenic Effects. Molecules 2016, 21, 901. [Google Scholar] [CrossRef]
  40. Jablonsky, M.; Majova, V.; Strizincova, P.; Sima, J.; Jablonsky, J. Investigation of Total Phenolic Content and Antioxidant Activities of Spruce Bark Extracts Isolated by Deep Eutectic Solvents. Crystals 2020, 10, 402. [Google Scholar] [CrossRef]
  41. Liu, Q.; Wang, Y.; Bian, J.; Li, M.-F.; Ren, J.-L.; Hao, X.; Peng, F. Sustainable Polar Aprotic/Poly-Deep Eutectic Solvent Systems for Highly Efficient Dissolution of Lignin. Green. Chem. 2023, 25, 4808–4817. [Google Scholar] [CrossRef]
  42. Hamad, A.M.A.; Ates, S.; Olgun, Ç.; Gür, M. Chemical Composition and Antioxidant Properties of Some Industrial Tree Bark Extracts. Bioresources 2019, 14, 5657–5671. [Google Scholar] [CrossRef]
  43. Sillero, L.; Prado, R.; Labidi, J. Simultaneous Microwave-Ultrasound Assisted Extraction of Bioactive Compounds from Bark. Chem. Eng. Process.-Process Intensif. 2020, 156, 108100. [Google Scholar] [CrossRef]
  44. Švarc-Gajić, J.; Cerdà, V.; Delerue-Matos, C.; Mašković, P.; Clavijo, S.; Suarez, R.; Cvetanović, A.; Ramalhosa, M.J.; Barroso, M.F.; Moreira, M.; et al. Chemical Characterization and In Vitro Bioactivity of Apple Bark Extracts Obtained by Subcritical Water. Waste Biomass Valorization 2021, 12, 6781–6794. [Google Scholar] [CrossRef]
  45. Singleton, V.L.; Rossi, J.A. Colorimetry of Total Phenolics with Phosphomolybdic-Phosphotungstic Acid Reagents. Am. J. Enol. Vitic. 1965, 16, 144–158. [Google Scholar] [CrossRef]
  46. Spinelli, S.; Costa, C.; Conte, A.; La Porta, N.; Padalino, L.; Del Nobile, M.A. Bioactive Compounds from Norway Spruce Bark: Comparison among Sustainable Extraction Techniques for Potential Food Applications. Foods 2019, 8, 524–532. [Google Scholar] [CrossRef]
  47. Sillero, L.; Prado, R.; Andrés, M.A.; Labidi, J. Characterisation of Bark of Six Species from Mixed Atlantic Forest. Ind. Crops Prod. 2019, 137, 276–284. [Google Scholar] [CrossRef]
  48. Sillero, L.; Prado, R.; Labidi, J. Optimization of Different Extraction Methods to Obtaining Bioactive Compounds from Larix Decidua Bark. Chem. Eng. Trans. 2018, 70, 1369–1374. [Google Scholar]
  49. Skrypnik, L.; Grigorev, N.; Michailov, D.; Antipina, M.; Danilova, M.; Pungin, A. Comparative Study on Radical Scavenging Activity and Phenolic Compounds Content in Water Bark Extracts of Alder (Alnus glutinosa (L.) Gaertn.), Oak (Quercus robur L.) and Pine (Pinus sylvestris L.). Eur. J. Wood Wood Prod. 2019, 77, 879–890. [Google Scholar] [CrossRef]
  50. Dróżdż, P.; Pyrzynska, K. Extracts from Pine and Oak Barks: Phenolics, Minerals and Antioxidant Potential. Int. J. Environ. Anal. Chem. 2021, 101, 464–472. [Google Scholar] [CrossRef]
  51. Dróżdż, P.; Pyrzynska, K. Assessment of Polyphenol Content and Antioxidant Activity of Oak Bark Extracts. Eur. J. Wood Wood Prod. 2018, 76, 793–795. [Google Scholar] [CrossRef]
  52. Okselni, T.; Santoni, A.; Dharma, A.; Efdi, M. Determination of antioxidant activity, total phenolic content, and total flavonoid content of roots, stem bark, and leaves of elaeocarpus mastersii king. Rasayan J. Chem. 2018, 11, 1211–1216. [Google Scholar] [CrossRef]
  53. Wairata, J.; Fadlan, A.; Setyo Purnomo, A.; Taher, M.; Ersam, T. Total Phenolic and Flavonoid Contents, Antioxidant, Antidiabetic and Antiplasmodial Activities of Garcinia Forbesii King: A Correlation Study. Arab. J. Chem. 2022, 15, 103541. [Google Scholar] [CrossRef]
  54. Phuyal, N.; Jha, P.K.; Raturi, P.P.; Rajbhandary, S. Total Phenolic, Flavonoid Contents, and Antioxidant Activities of Fruit, Seed, and Bark Extracts of Zanthoxylum Armatum DC. Sci. World J. 2020, 2020, 8780704. [Google Scholar] [CrossRef] [PubMed]
  55. Tanase, C.; Nicolescu, A.; Nisca, A.; Ștefănescu, R.; Babotă, M.; Mare, A.D.; Ciurea, C.N.; Man, A. Biological Activity of Bark Extracts from Northern Red Oak (Quercus rubra L.): An Antioxidant, Antimicrobial and Enzymatic Inhibitory Evaluation. Plants 2022, 11, 2357. [Google Scholar] [CrossRef]
  56. Nisca, A.; Ștefănescu, R.; Moldovan, C.; Mocan, A.; Mare, A.D.; Ciurea, C.N.; Man, A.; Muntean, D.-L.; Tanase, C. Optimization of Microwave Assisted Extraction Conditions to Improve Phenolic Content and In Vitro Antioxidant and Anti-Microbial Activity in Quercus Cerris Bark Extracts. Plants 2022, 11, 240. [Google Scholar] [CrossRef]
  57. Tanase, C.; Babotă, M.; Nișca, A.; Nicolescu, A.; Ștefănescu, R.; Mocan, A.; Farczadi, L.; Mare, A.D.; Ciurea, C.N.; Man, A. Potential Use of Quercus Dalechampii Ten. and Q. Frainetto Ten. Barks Extracts as Antimicrobial, Enzyme Inhibitory, Antioxidant and Cytotoxic Agents. Pharmaceutics 2023, 15, 343. [Google Scholar] [CrossRef] [PubMed]
  58. Agarwal, C.; Hofmann, T.; Vršanská, M.; Schlosserová, N.; Visi-Rajczi, E.; Voběrková, S.; Pásztory, Z. In Vitro Antioxidant and Antibacterial Activities with Polyphenolic Profiling of Wild Cherry, the European Larch and Sweet Chestnut Tree Bark. Eur. Food Res. Technol. 2021, 247, 2355–2370. [Google Scholar] [CrossRef]
  59. Munira, S.; Islam, A.; Islam, S.; Koly, S.F.; Nesa, L.; Muhit, A. Phytochemical Screening and Comparative Antioxidant Activities of Fractions Isolated from Sonneratia Caseolaris (Linn.) Bark Extracts. Eur. J. Med. Plants 2019, 28, 1–9. [Google Scholar] [CrossRef]
  60. Itam, A.; Wati, M.S.; Agustin, V.; Sabri, N.; Jumanah, R.A.; Efdi, M. Comparative Study of Phytochemical, Antioxidant, and Cytotoxic Activities and Phenolic Content of Syzygium Aqueum (Burm. f. Alston f.) Extracts Growing in West Sumatera Indonesia. Sci. World J. 2021, 2021, 5537597. [Google Scholar] [CrossRef] [PubMed]
  61. Aktaş Karaçelik, A.; Şeker, M.E.; Karakose, M. Determination of Antioxidant Activity of Different Extracts from Bark of Pinus Spp. Grown in Giresun (Turkey) Province—Phenolic Analysis by RP-HPLC-DAD. KSU J. Agric. Nat. 2022, 25, 10–18. [Google Scholar] [CrossRef]
  62. Piątczak, E.; Dybowska, M.; Płuciennik, E.; Kośla, K.; Kolniak-Ostek, J.; Kalinowska-Lis, U. Identification and Accumulation of Phenolic Compounds in the Leaves and Bark of Salix alba (L.) and Their Biological Potential. Biomolecules 2020, 10, 1391. [Google Scholar] [CrossRef]
  63. Salem, M.Z.M.; Mansour, M.M.A.; Elansary, H.O. Evaluation of the Effect of Inner and Outer Bark Extracts of Sugar Maple (Acer Saccharum Var. Saccharum) in Combination with Citric Acid against the Growth of Three Common Molds. J. Wood Chem. Technol. 2019, 39, 136–147. [Google Scholar] [CrossRef]
  64. Yongram, C.; Sungthong, B.; Puthongking, P.; Weerapreeyakul, N. Chemical Composition, Antioxidant and Cytotoxicity Activities of Leaves, Bark, Twigs and Oleo-Resin of Dipterocarpus Alatus. Molecules 2019, 24, 3083. [Google Scholar] [CrossRef]
  65. Indirayati, N.; Nisa, K.; Kurang, R.Y.; Tarmo, N.C.; Adang, K.T.P. Radical Scavenging Activity and Total Phenolic Content of Seven Tropical Plants. IOP Conf. Ser. Earth Environ. Sci. 2020, 462, 012043. [Google Scholar] [CrossRef]
  66. Safaeian, L.; Haghighatian, Z.; Zolfaghari, B.; Amindeldar, M. Cardioprotective Effects of Pinus Eldarica Bark Extract on Adrenaline-Induced Myocardial Infarction in Rats. Asian Pac. J. Trop. Biomed. 2023, 13, 148. [Google Scholar] [CrossRef]
  67. Tanase, C.; Cosarca, S.; Toma, F.; Mare, A.; Man, A.; Miklos, A.; Imre, S.; Boz, I. Antibacterial activities of beech bark (Fagus sylvatica L.) polyphenolic extract. Environ. Eng. Manag. J. 2018, 17, 877–884. [Google Scholar] [CrossRef]
  68. Hendrik, J.; Hadi, Y.S.; Massijaya, M.Y.; Santoso, A.; Pizzi, A. Properties of Glued Laminated Timber Made from Fast-Growing Species with Mangium Tannin and Phenol Resorcinol Formaldehyde Adhesives. J. Korean Wood Sci. Technol. 2019, 47, 253–264. [Google Scholar] [CrossRef]
  69. Hajriani, S.; Yunianti, A.D.; Suhasman, S. Determination of Optimum Formulation of Tusam (Pinus merkusii) Tannin Bark with Resorcinol and Formaldehyde. IOP Conf. Ser. Mater. Sci. Eng. 2020, 935, 012032. [Google Scholar] [CrossRef]
  70. Bilgin, U.; Colakoglu, G. Effect of Using Urea Formaldehyde Modified with Extracts in Plywood on Formaldehyde Emission. Drv. Ind. 2021, 72, 237–244. [Google Scholar] [CrossRef]
  71. Vangeel, T.; Neiva, D.M.; Quilhó, T.; Costa, R.A.; Sousa, V.; Sels, B.F.; Pereira, H. Tree Bark Characterization Envisioning an Integrated Use in a Biorefinery. Biomass Convers. Biorefinery 2023, 13, 2029–2043. [Google Scholar] [CrossRef]
  72. Peng, Y.; Wang, Y.; Chen, P.; Wang, W.; Cao, J. Enhancing Weathering Resistance of Wood by Using Bark Extractives as Natural Photostabilizers in Polyurethane-Acrylate Coating. Prog. Org. Coat. 2020, 145, 105665. [Google Scholar] [CrossRef]
  73. Gruber, L.; Seidl, L.; Zanetti, M.; Schnabel, T. Calorific Value and Ash Content of Extracted Birch Bark. Forests 2021, 12, 1480. [Google Scholar] [CrossRef]
  74. Nath, S.K.; Islam, M.N.; Rahman, K.-S.; Rana, M.N. Tannin-Based Adhesive from Ceriops decandra (Griff.) Bark for the Production of Particleboard. J. Indian Acad. Wood Sci. 2018, 15, 21–27. [Google Scholar] [CrossRef]
  75. Hafiz, N.L.M.; Tahir, P.M.; Hua, L.S.; Abidin, Z.Z.; Sabaruddin, F.A.; Yunus, N.M.; Abdullah, U.H.; Abdul Khalil, H.P.S. Curing and Thermal Properties of Co-Polymerized Tannin Phenol–Formaldehyde Resin for Bonding Wood Veneers. J. Mater. Res. Technol. 2020, 9, 6994–7001. [Google Scholar] [CrossRef]
  76. Reh, R.; Kristak, L.; Sedliacik, J.; Bekhta, P.; Wronka, A.; Kowaluk, G. Molded Plywood with Proportions of Beech Bark in Adhesive Mixtures: Production on an Industrial Scale. Polymers 2024, 16, 966. [Google Scholar] [CrossRef]
  77. Janceva, S.; Andersone, A.; Spulle, U.; Tupciauskas, R.; Papadopoulou, E.; Bikovens, O.; Andzs, M.; Zaharova, N.; Rieksts, G.; Telysheva, G. Eco-Friendly Adhesives Based on the Oligomeric Condensed Tannins-Rich Extract from Alder Bark for Particleboard and Plywood Production. Materials 2022, 15, 3894. [Google Scholar] [CrossRef]
  78. Santos, J.; Delgado, N.; Fuentes, J.; Fuentealba, C.; Vega-Lara, J.; García, D.E. Exterior Grade Plywood Adhesives Based on Pine Bark Polyphenols and Hexamine. Ind. Crops Prod. 2018, 122, 340–348. [Google Scholar] [CrossRef]
  79. García, D.E.; Glasser, W.G.; Pizzi, A.; Lacoste, C.; Laborie, M.-P. Polyphenolic Resins Prepared with Maritime Pine Bark Tannin and Bulky-Aldehydes. Ind. Crops Prod. 2014, 62, 84–93. [Google Scholar] [CrossRef]
  80. Wibowo, E.S.; Park, B.-D. Surface Adhesion of PMDI Resin on Wood Biopolymer Model Films. Eur. J. Wood Wood Prod. 2023, 81, 1305–1312. [Google Scholar] [CrossRef]
  81. Aristri, M.A.; Lubis, M.A.R.; Laksana, R.P.B.; Sari, R.K.; Iswanto, A.H.; Kristak, L.; Antov, P.; Pizzi, A. Thermal and Mechanical Performance of Ramie Fibers Modified with Polyurethane Resins Derived from Acacia Mangium Bark Tannin. J. Mater. Res. Technol. 2022, 18, 2413–2427. [Google Scholar] [CrossRef]
  82. Aristri, M.A.; Sari, R.K.; Lubis, M.A.R.; Laksana, R.P.B.; Antov, P.; Iswanto, A.H.; Mardawati, E.; Lee, S.H.; Savov, V.; Kristak, L.; et al. Eco-Friendly Tannin-Based Non-Isocyanate Polyurethane Resins for the Modification of Ramie (Boehmeria nivea L.) Fibers. Polymers 2023, 15, 1492. [Google Scholar] [CrossRef]
  83. Lubis, M.A.R.; Aristri, M.A.; Sari, R.K.; Iswanto, A.H.; Al-Edrus, S.S.O.; Sutiawan, J.; Lee, S.H.; Antov, P.; Kristak, L. Isocyanate–Free Tannin–Based Polyurethane Resins for Enhancing Thermo-Mechanical Properties of Ramie (Boehmeria nivea L.) Fibers. Alex. Eng. J. 2024, 90, 54–64. [Google Scholar] [CrossRef]
  84. Hussain, I.; Sanglard, M.; Bridson, J.H.; Parker, K. Preparation and Physicochemical Characterisation of Polyurethane Foams Prepared Using Hydroxybutylated Condensed Tannins as a Polyol Source. Ind. Crops Prod. 2020, 154, 112636. [Google Scholar] [CrossRef]
  85. D’Souza, J.; Camargo, R.; Yan, N. Polyurethane Foams Made from Liquefied Bark-based Polyols. J. Appl. Polym. Sci. 2014, 131, 40599. [Google Scholar] [CrossRef]
  86. Yalcin, M. Surface Glossiness Properties of Wood Impregnated with Some Plant Extracts. Forestist 2018, 68, 61–69. [Google Scholar] [CrossRef]
  87. Lemouzy, S.; Delavarde, A.; Lamaty, F.; Bantreil, X.; Pinaud, J.; Caillol, S. Lignin-Based Bisguaiacol Diisocyanate: A Green Route for the Synthesis of Biobased Polyurethanes. Green. Chem. 2023, 25, 4833–4839. [Google Scholar] [CrossRef]
  88. Han, S.; Yao, S.; Meng, W.; Yang, J. Rapid, Controlled Ring-Opening Polymerization of Salicylic Acid o-Carboxyanhydride for Poly(Salicylate) Synthesis. Polym. Chem. 2021, 12, 6465–6471. [Google Scholar] [CrossRef]
  89. Lang, J.; Dörrstein, J.; Zollfrank, C. Archaeo-Inspired Material Synthesis: Sustainable Tackifiers and Adhesives from Birch Bark. Green. Mater. 2018, 6, 157–164. [Google Scholar] [CrossRef]
  90. Okada, M.; Suzuki, K.; Mawatari, Y.; Tabata, M. Biopolyester Prepared Using Unsaturated Betulin (Betulinol) Extracted from Outer Birch Bark and Dicarboxylic Acid Dichlorides and Its Thermal-Induced Crosslinking. Eur. Polym. J. 2019, 113, 12–17. [Google Scholar] [CrossRef]
  91. Kumar, A.; Korpinen, R.; Möttönen, V.; Verkasalo, E. Suberin Fatty Acid Hydrolysates from Outer Birch Bark for Hydrophobic Coating on Aspen Wood Surface. Polymers 2022, 14, 832. [Google Scholar] [CrossRef] [PubMed]
  92. Gosecki, M.; Urbaniak, M.; Makarewicz, C.; Gosecka, M. Converting Unrefined Birch Suberin Monomers into Vitrimer. ACS Sustain. Chem. Eng. 2024, 12, 3841–3850. [Google Scholar] [CrossRef]
  93. Cho, S.-H.; Yoon, B.; Lee, S.K.; Nam, J.-D.; Suhr, J. Natural Cork Suberin-Originated Ecofriendly Biopolyester Syntactic Foam. ACS Sustain. Chem. Eng. 2022, 10, 7508–7514. [Google Scholar] [CrossRef]
  94. McMichael, P.; Schultze, X.; Cramail, H.; Peruch, F. Sourcing, Thermodynamics, and Ring-Opening (Co)Polymerization of Substituted δ-Lactones: A Review. Polym. Chem. 2023, 14, 3783–3812. [Google Scholar] [CrossRef]
  95. Rali, T.; Wossa, S.; Leach, D. Comparative Chemical Analysis of the Essential Oil Constituents in the Bark, Heartwood and Fruits of Cryptocarya massoy (Oken) Kosterm. (Lauraceae) from Papua New Guinea. Molecules 2007, 12, 149–154. [Google Scholar] [CrossRef]
  96. Garcia, D.E.; Carrasco, J.C.; Salazar, J.P.; Perez, M.A.; Cancino, R.A.; Riquelme, S. Bark Polyflavonoids from Pinus Radiata as Functional Building-Blocks for Polylactic Acid (PLA)-Based Green Composites. Express Polym. Lett. 2016, 10, 835–848. [Google Scholar] [CrossRef]
  97. Bridson, J.H.; Grigsby, W.J.; Main, L. One-Pot Solvent-Free Synthesis and Characterisation of Hydroxypropylated Polyflavonoid Compounds. Ind. Crops Prod. 2018, 111, 529–535. [Google Scholar] [CrossRef]
  98. Cortés-Guzmán, K.P.; Parikh, A.R.; Sparacin, M.L.; Johnson, R.M.; Adegoke, L.; Ecker, M.; Voit, W.E.; Smaldone, R.A. Thermal Annealing Effects on the Mechanical Properties of Bio-Based 3D Printed Thermosets. Polym. Chem. 2023, 14, 2697–2707. [Google Scholar] [CrossRef]
  99. Bridson, J.H.; Sanglard, M.; Hussain, I.; Bouad, V.; Patel, M.; Parker, K. Hydroxyalkylation of Condensed Tannins: Comparison of Proanthocyanidin Extraction Process and Epoxide Chain Length on Physicochemical Properties. Ind. Crops Prod. 2019, 140, 111618. [Google Scholar] [CrossRef]
  100. Montoya, L.F.; Contreras, D.; Jaramillo, A.F.; Carrasco, C.; Fernández, K.; Schwederski, B.; Rojas, D.; Melendrez, M.F. Study of Anticorrosive Coatings Based on High and Low Molecular Weight Polyphenols Extracted from the Pine Radiata Bark. Prog. Org. Coat. 2019, 127, 100–109. [Google Scholar] [CrossRef]
  101. Shnawa, H.A. Curing and Thermal Properties of Tannin-Based Epoxy and Its Blends with Commercial Epoxy Resin. Polym. Bull. 2021, 78, 1925–1940. [Google Scholar] [CrossRef]
  102. Zhu, Q.; Tan, J.; Li, D.; Zhang, T.; Liu, Z.; Cao, Y. Cross-Linked Chitosan/Tannin Extract as a Biodegradable and Repulpable Coating for Paper with Excellent Oil-Resistance, Gas Barrier and UV-Shielding. Prog. Org. Coat. 2023, 176, 107399. [Google Scholar] [CrossRef]
  103. Zhang, T.; Yu, C.; Yu, M.; Huang, Y.; Tan, J.; Zhang, M.; Zhu, X. Multifunctional Tannin Extract-Based Epoxy Derived from Waste Bark as a Highly Toughening and Strengthening Agent for Epoxy Resin. Ind. Crops Prod. 2022, 176, 114255. [Google Scholar] [CrossRef]
  104. Chen, B.; Wu, Q.; Li, J.; Lin, K.; Chen, D.; Zhou, C.; Wu, T.; Luo, X.; Liu, Y. A Novel and Green Method to Synthesize a Epoxidized Biomass Eucommia Gum as the Nanofiller in the Epoxy Composite Coating with Excellent Anticorrosive Performance. Chem. Eng. J. 2020, 379, 122323. [Google Scholar] [CrossRef]
  105. Cao, Q.; Weng, Z.; Qi, Y.; Li, J.; Liu, W.; Liu, C.; Zhang, S.; Wei, Z.; Chen, Y.; Jian, X. Achieving Higher Performances without an External Curing Agent in Natural Magnolol-Based Epoxy Resin. Chin. Chem. Lett. 2022, 33, 2195–2199. [Google Scholar] [CrossRef]
  106. Cao, Q.; Li, J.; Qi, Y.; Zhang, S.; Wang, J.; Wei, Z.; Pang, H.; Jian, X.; Weng, Z. Engineering Double Load-Sharing Network in Thermosetting: Much More than Just Toughening. Macromolecules 2022, 55, 9502–9512. [Google Scholar] [CrossRef]
  107. Zhang, C.; Song, S.; Cao, Q.; Li, J.; Liu, Q.; Zhang, S.; Jian, X.; Weng, Z. Improving the Comprehensive Properties of Chitosan-Based Thermal Insulation Aerogels by Introducing a Biobased Epoxy Thermoset to Form an Anisotropic Honeycomb-Layered Structure. Int. J. Biol. Macromol. 2023, 246, 125616. [Google Scholar] [CrossRef] [PubMed]
  108. Wei, C.; Gao, T.; Xu, Y.; Yang, W.; Dai, G.; Li, R.; Zhu, S.E.; Yuen, R.K.K.; Yang, W.; Lu, H. Synthesis of Bio-Based Epoxy Containing Phosphine Oxide as a Reactive Additive toward Highly Toughened and Fire-Retarded Epoxy Resins. Chin. J. Polym. Sci. 2023, 41, 1733–1746. [Google Scholar] [CrossRef]
  109. Weng, Z.; Song, L.; Qi, Y.; Li, J.; Cao, Q.; Liu, C.; Zhang, S.; Wang, J.; Jian, X. Natural Magnolol Derivatives as Platform Chemicals for Bio-Based Phthalonitrile Thermoset: Achieving High Performances without an External Curing Agent. Polymer 2021, 226, 123814. [Google Scholar] [CrossRef]
  110. Wang, Z.; Wang, Y.; Wang, H.; Gang, H.; Zhang, N.; Zhou, Y.; Gu, S.; Zhuang, Y.; Xu, W.; Ke, G.; et al. Bioinspired Natural Magnolol-Based Adhesive with Strong Adhesion and Antibacterial Properties for Application in Wet and Dry Environments. ACS Appl. Mater. Interfaces 2023, 15, 24846–24857. [Google Scholar] [CrossRef] [PubMed]
  111. Weems, A.C.; Delle Chiaie, K.R.; Worch, J.C.; Stubbs, C.J.; Dove, A.P. Terpene- and Terpenoid-Based Polymeric Resins for Stereolithography 3D Printing. Polym. Chem. 2019, 10, 5959–5966. [Google Scholar] [CrossRef]
  112. Li, A.Y.; Melendez-Zamudio, M.; Yepremyan, A.; Brook, M.A. Learning from the Trees: Biomimetic Crosslinking of Silicones by Phenolic Coupling. Green. Chem. 2023, 25, 5267–5275. [Google Scholar] [CrossRef]
  113. Şen, A.U.; Esteves, B.; Lemos, F.; Pereira, H. Insights into the Combustion Behavior of Cork and Phloem: Effect of Chemical Components and Biomass Morphology. Eur. J. Wood Wood Prod. 2023, 81, 999–1010. [Google Scholar] [CrossRef]
  114. Warlo, H.; Windeisen-Holzhauser, E.; Brüchert, F.; Sauter, U.H.; Richter, K. Extractives in Douglas Firs (Pseudotsuga menziesii (Mirb.) Franco) from Three Sites in South-West Germany and Potential Opportunities for Valorization. Eur. J. Wood Wood Prod. 2023, 81, 1093–1108. [Google Scholar] [CrossRef]
  115. Pinnataip, R.; Lee, B.P. Oxidation Chemistry of Catechol Utilized in Designing Stimuli-Responsive Adhesives and Antipathogenic Biomaterials. ACS Omega 2021, 6, 5113–5118. [Google Scholar] [CrossRef]
Figure 1. Simplified sourcing process of transforming bark into thermosetting resin.
Figure 1. Simplified sourcing process of transforming bark into thermosetting resin.
Materials 17 02123 g001
Figure 2. Two representatives of trees with exfoliating bark. The plane tree (left) has irregularly shaped flakes of bark (marked with green circles), which can be easily removed. Birch (right) has a thin paper-like bark with horizontal lenticels (marked with red circles), which are more fibrous.
Figure 2. Two representatives of trees with exfoliating bark. The plane tree (left) has irregularly shaped flakes of bark (marked with green circles), which can be easily removed. Birch (right) has a thin paper-like bark with horizontal lenticels (marked with red circles), which are more fibrous.
Materials 17 02123 g002
Figure 3. The locations of the tree species used to obtain all reported bark extracts. The sketch of the world map (author: OpenClipart-Vectors) was obtained as a free graphic from Pixabay GmbH (https://pixabay.com/vectors/map-world-geography-continents-117174/ (accessed on 15 January 2024)).
Figure 3. The locations of the tree species used to obtain all reported bark extracts. The sketch of the world map (author: OpenClipart-Vectors) was obtained as a free graphic from Pixabay GmbH (https://pixabay.com/vectors/map-world-geography-continents-117174/ (accessed on 15 January 2024)).
Materials 17 02123 g003
Scheme 1. Structures of identified abundant compounds in bark of Norway spruce (Picea abies): (1) stilbenoids and stilbene glycosides (astringin: R1—beta-D-glucosyl, R2—OH; trans-resveratrol: R1—OH, R2—H; trans-isorhapontin: R1—β-D-glucosyl, R2—OCH3); (2) D-glucopyranose (glucose); (3) gluconic acid; (4) dehydroabietic acid.
Scheme 1. Structures of identified abundant compounds in bark of Norway spruce (Picea abies): (1) stilbenoids and stilbene glycosides (astringin: R1—beta-D-glucosyl, R2—OH; trans-resveratrol: R1—OH, R2—H; trans-isorhapontin: R1—β-D-glucosyl, R2—OCH3); (2) D-glucopyranose (glucose); (3) gluconic acid; (4) dehydroabietic acid.
Materials 17 02123 sch001
Scheme 2. Structures of identified abundant compounds in bark of different pine species (Pinus sp.): (1) flavan-3-ols (catechin: R1—H, a, b—S, R (−)/R, S (+); epicatechin: R1—H, a, b—S, S (+)/R, R (−); gallocatechin: R1—OH, a, b—S, R (−)/R, S (+)); (2) other flavonoids (taxifolin: R3—OH, R4—H, R5—OH, c—single bond; myricetin: R3—OH, R4—OH, c—double bond, R5—OH; naringenin: R3—H, R4—H, R5—H, c—single bond; (3) ellagic acid; (4) caffeic acid: R6—OH; ferulic acid: R6—OCH3; (5) guaiacol: R7—H; 4-methylguaiacol: R7—CH3; 4-vinylguaiacol: R7—CH=CH2; (6) protocatechuic acid.
Scheme 2. Structures of identified abundant compounds in bark of different pine species (Pinus sp.): (1) flavan-3-ols (catechin: R1—H, a, b—S, R (−)/R, S (+); epicatechin: R1—H, a, b—S, S (+)/R, R (−); gallocatechin: R1—OH, a, b—S, R (−)/R, S (+)); (2) other flavonoids (taxifolin: R3—OH, R4—H, R5—OH, c—single bond; myricetin: R3—OH, R4—OH, c—double bond, R5—OH; naringenin: R3—H, R4—H, R5—H, c—single bond; (3) ellagic acid; (4) caffeic acid: R6—OH; ferulic acid: R6—OCH3; (5) guaiacol: R7—H; 4-methylguaiacol: R7—CH3; 4-vinylguaiacol: R7—CH=CH2; (6) protocatechuic acid.
Materials 17 02123 sch002
Scheme 3. Structures of identified abundant compounds in bark of other conifers: (1) guaiacol: R1—H; 4-methylguaiacol: R1—CH3; 4-vinylguaiacol: R1—CH=CH2; (2) astringin; (3) (+)-catechin; (4) flavones (myricetin: R2—OH, R3—OH, R4—OH; isorhamnetin: R2—OH, R3—H, R4—OCH3; isorhamnetin glucoside: R2—β-D-glucosyl, R3—OH, R4—OCH3; quercetin glycoside: R2—β-D-glucosyl acetate, R3—H, R4—OH).
Scheme 3. Structures of identified abundant compounds in bark of other conifers: (1) guaiacol: R1—H; 4-methylguaiacol: R1—CH3; 4-vinylguaiacol: R1—CH=CH2; (2) astringin; (3) (+)-catechin; (4) flavones (myricetin: R2—OH, R3—OH, R4—OH; isorhamnetin: R2—OH, R3—H, R4—OCH3; isorhamnetin glucoside: R2—β-D-glucosyl, R3—OH, R4—OCH3; quercetin glycoside: R2—β-D-glucosyl acetate, R3—H, R4—OH).
Materials 17 02123 sch003
Scheme 4. Structures of identified abundant compounds in bark of different oak species: (1) flavan-3-ols (catechin: R1—H, a, b—S, R (−)/R, S (+); gallocatechin: R1—OH, a, b—S, R (−)/R, S (+)); (2) 4-vinylguaiacol; (3) myricetin; (4) caffeic acid; (5) p-hydroxybenzoic acid: R2—H, R3 —H; vanillic acid: R2—OCH3, R3—H; gallic acid: R2—OH, R3—OH.
Scheme 4. Structures of identified abundant compounds in bark of different oak species: (1) flavan-3-ols (catechin: R1—H, a, b—S, R (−)/R, S (+); gallocatechin: R1—OH, a, b—S, R (−)/R, S (+)); (2) 4-vinylguaiacol; (3) myricetin; (4) caffeic acid; (5) p-hydroxybenzoic acid: R2—H, R3 —H; vanillic acid: R2—OCH3, R3—H; gallic acid: R2—OH, R3—OH.
Materials 17 02123 sch004
Scheme 5. Structures of identified abundant compounds in bark of other deciduous trees of north temperate zone: (1) phenolic acids (benzoic acid: R1—H, R2—H, R3—H, R4—H; gallic acid: R1—OH, R2—OH, R3—OH, R4—H; vanillic acid: R1—OCH3, R2—OH, R3—H, R4—H; p-hydroxybenzoic acid: R1—H, R2—OH, R3—H, R4—H; phthalic acid: R1—H, R2—H, R3—H, R4—COOH); (2) guaiacol: R5—H; 4-methylguaiacol: R5—CH3; 4-vinylguaiacol: R5—CH=CH2; (3) quinic acid; (4) caffeine; (5) trigalloyl-HHDP-glucose; (6) vescalagin; (7) palmitic acid; (8) flavan-3-ols (catechin: a, b—S, R (−)/R, S (+); (-)-epicatechin: R1—H, a, b—R, R); (9) other flavonoids (myricetin: R6—OH, R7—OH, R8—OH, R9—OH, R10—OH, c—double bond; quercetin: R6—OH, R7—H, R8—OH, R9—OH, R10—OH, c—double bond; luteolin-O-hexoside: R6—H, R7–OH, R8—β-D-glucosyl, R9—H, R10—OH, c—double bond; apigenin-O-hexoside: R6—H, R7—H, R8—OH, R9—H, R10—β-D-glucosyl, c—double bond; taxifolin: R6—OH, R7—H, R8—OH, R9—OH, R10—OH, c—single bond; taxifolin-3-glucoside: R6–β-D-glucosyl (S), R7—H, R8—OH, R9—OH, R10—OH, c—single bond; taxifolin-7-glucoside: R6—OH (S), R7—H, R8—OH, R9—OH, R10—β-D-glucosyl (R), c—single bond; kaempferol-O-hexoside: R6—β-D-glucosyl, R7—H, R8—OH, R9—H, R10—OH, c—double bond; (10) scopolin; (11) syringin; (12) daidzein-O-hexoside; (13) caffeoyl hexose: R11—OH; caffeoyl hexose deoxyhexoside: R11—deoxyhexosyl; (14) procyanidin dimer type A (A1: e—S; A2: d—S, e—R).
Scheme 5. Structures of identified abundant compounds in bark of other deciduous trees of north temperate zone: (1) phenolic acids (benzoic acid: R1—H, R2—H, R3—H, R4—H; gallic acid: R1—OH, R2—OH, R3—OH, R4—H; vanillic acid: R1—OCH3, R2—OH, R3—H, R4—H; p-hydroxybenzoic acid: R1—H, R2—OH, R3—H, R4—H; phthalic acid: R1—H, R2—H, R3—H, R4—COOH); (2) guaiacol: R5—H; 4-methylguaiacol: R5—CH3; 4-vinylguaiacol: R5—CH=CH2; (3) quinic acid; (4) caffeine; (5) trigalloyl-HHDP-glucose; (6) vescalagin; (7) palmitic acid; (8) flavan-3-ols (catechin: a, b—S, R (−)/R, S (+); (-)-epicatechin: R1—H, a, b—R, R); (9) other flavonoids (myricetin: R6—OH, R7—OH, R8—OH, R9—OH, R10—OH, c—double bond; quercetin: R6—OH, R7—H, R8—OH, R9—OH, R10—OH, c—double bond; luteolin-O-hexoside: R6—H, R7–OH, R8—β-D-glucosyl, R9—H, R10—OH, c—double bond; apigenin-O-hexoside: R6—H, R7—H, R8—OH, R9—H, R10—β-D-glucosyl, c—double bond; taxifolin: R6—OH, R7—H, R8—OH, R9—OH, R10—OH, c—single bond; taxifolin-3-glucoside: R6–β-D-glucosyl (S), R7—H, R8—OH, R9—OH, R10—OH, c—single bond; taxifolin-7-glucoside: R6—OH (S), R7—H, R8—OH, R9—OH, R10—β-D-glucosyl (R), c—single bond; kaempferol-O-hexoside: R6—β-D-glucosyl, R7—H, R8—OH, R9—H, R10—OH, c—double bond; (10) scopolin; (11) syringin; (12) daidzein-O-hexoside; (13) caffeoyl hexose: R11—OH; caffeoyl hexose deoxyhexoside: R11—deoxyhexosyl; (14) procyanidin dimer type A (A1: e—S; A2: d—S, e—R).
Materials 17 02123 sch005
Figure 4. Functional groups of bark phytochemicals and their applicability in the synthesis of thermosetting polymers.
Figure 4. Functional groups of bark phytochemicals and their applicability in the synthesis of thermosetting polymers.
Materials 17 02123 g004
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Szmechtyk, T.; Małecka, M. Phytochemicals from Bark Extracts and Their Applicability in the Synthesis of Thermosetting Polymers: An Overview. Materials 2024, 17, 2123. https://doi.org/10.3390/ma17092123

AMA Style

Szmechtyk T, Małecka M. Phytochemicals from Bark Extracts and Their Applicability in the Synthesis of Thermosetting Polymers: An Overview. Materials. 2024; 17(9):2123. https://doi.org/10.3390/ma17092123

Chicago/Turabian Style

Szmechtyk, Tomasz, and Magdalena Małecka. 2024. "Phytochemicals from Bark Extracts and Their Applicability in the Synthesis of Thermosetting Polymers: An Overview" Materials 17, no. 9: 2123. https://doi.org/10.3390/ma17092123

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop