Modification of condensed tannins: from polyphenol chemistry to materials engineering

Danny E. García *a, Wolfgang G. Glasser b, Antonio Pizzi cd, Sebastian P. Paczkowski ef and Marie-Pierre Laborie ef
aÁrea Productos Químicos, Unidad de Desarrollo Tecnológico (UDT), Universidad de Concepción, Bio-Bio, Chile. E-mail: d.garcia@udt.cl
bDepartment of Sustainable Biomaterials, Virginia Polytechnic Institute and State University, Blacksburg, Virginia, USA
cLaboratoire d'Etudes et de Recherche sur le Matériau Bois (LERMAB), University of Lorraine, 27 rue du Merle Blanc, Epinal, France
dDepartment of Physics, King Abdulaziz University, Jeddah, Saudi Arabia
eFreiburger Materialforschungszentrum–FMF, Albert-Ludwigs University of Freiburg, Freiburg, Germany
fChair of Forest Biomaterials, University of Freiburg, Freiburg, Germany

Received (in Montpellier, France) 10th August 2015 , Accepted 2nd November 2015

First published on 6th November 2015


Abstract

Condensed tannins (CTs) are high molar mass polyphenolic bio-polymers based on flavonol units. CTs have been well-studied due to their respective biological activity. However, the application of CTs in several areas is limited because of their physicochemical properties. The objective of this review is to investigate the state of the art regarding the chemical modification of CTs, and to outline recent and potential applications of tannin derivatives. An overview of the most important reactions is given, and a comprehensive summary of the experimental parameters for modification (chemicals, time, temperature, solvent type, and yield) is presented. The impact of the modification on the physicochemical properties of derivatives in comparison to native tannin behavior is discussed. Finally, the applicability of modified or unmodified CTs is described, referring to academic articles and patents. Future research in terms of modification reaction type, as well as derivatizing agents, is discussed.


image file: c5nj02131f-p1.tif

Danny E. García

Danny Eugenio García Marrero was born in Cuba in 1977. He received his MSc degree from the Matanzas University (Cuba), working on Phytochemistry and Organic analysis. For his PhD studies, he moved to the Albert-Ludwigs University of Freiburg - (Germany) to work with Professor Marie-Pierre Laborie on bark polyphenols chemistry. He received his Dr. rer. Nat. degree (Summa Cum Laude) on condensed tannin chemical modification. He joined the Department of Chemical Products from the Technological Development Unit (UDT), Concepción University (Chile) in 2015. He has more than 80 publications and was awarded for his research work. The main research theme of his work is the valorization of tannin, and lignin from forestry biomass.

image file: c5nj02131f-p2.tif

Wolfgang G. Glasser

Glasser's research and teaching efforts have been focused on high-value uses of sustainable bioresources by means of chemistry since obtaining a doctorate in Hamburg, Germany. In 1986, at Virginia Tech, he founded the first Biobased Materials Center. His work with polysaccharides and lignin resulted in 20 patents and 200+ publications and was recognized with awards from pulp and paper, forestry and chemical societies.

image file: c5nj02131f-p3.tif

Antonio Pizzi

A. Pizzi is a full professor of industrial chemistry at the ENSTIB. Prof. Pizzi, who holds a DChem (Polymers, Rome, Italy), a PhD (Organic Chemistry, South Africa) and a DSc (Wood Chemistry, South Africa), is the author of more than 700 research and technical articles, patents, contract reports and international conference papers as well as 7 books on adhesion and adhesives published in New York, and is the recipient of numerous international prizes for new industrial developments in his fields of specialization. His best-known areas of specialization are wood and fiber glueing and wood adhesives chemistry, formulation and application, in particular composite products based on natural materials.

image file: c5nj02131f-p4.tif

Sebastian P. Paczkowski

Dr Sebastian Paczkowski studied Forestry and Wood Technology at the University of Göttingen, Germany, and holds a PhD in Chemical Ecology. After two years of Post-Doc experience at the Chair of Forest Biomaterials and the Chair of Forest Engineering at the University of Freiburg, Germany, he currently works as a Research Associate and a Lecturer for Bioenergy at the Chair of Forest Engineering, University of Freiburg, Germany. His major research activities are wood fuel modifications with polyphenols or clay minerals to produce low emission fuels and the detection of thermally or physiologically induced chemical processes via VOC/gas sensor measurements for industrial applications such as wood or wood flake drying or wood fuel storage.

image file: c5nj02131f-p5.tif

Marie-Pierre Laborie

Marie-Pierre Laborie received an engineering degree in Wood Science from ENSTIB (France) in 1996 and a PhD from the Department of Wood Science and Forest Products at Virginia Tech (USA) in 2002. She subsequently joined the Composite Materials and Engineering Center at Washington State University as a faculty member of the Civil and Environment Engineering Department. There she focused on adhesion and macromolecular design for bio-based polymers and natural fiber composites. In 2008, she received her Habilitation in Materials Sciences and Process Engineering from Grenoble Institute of Technology (France). She currently holds the Chair of Forest Biomaterials in the Faculty of Environment and Natural Resources at the University of Freiburg. Her research takes a molecular scale approach to lignocellulosic materials towards their value-adding into novel bio-based polymers and composites. She has co-authored over 60 peer-reviewed publications and is a fellow of the International Academy of Wood Science.


1. Condensed tannins (CTs): renewable building-block polymers

1.1. CTs as abundant polyphenols

Metabolites produced by living organisms are classified as either primary or secondary. Primary metabolites are essential, and their concentration within the organism depends only on their physiological state.1 Proteins, nucleic acids, carbohydrates, and lipids are prominent examples for primary metabolic products. In contrast, secondary metabolites are not strictly essential for the survival of the organism, but they play an important ecological role in nature.2 There are five major classes of secondary metabolites defined by their chemical structure: nitrogen compounds (alkaloids, non-proteinogenic aminoacids, lectins, and cyanogens), isoprenoids (terpenes and steroids), polyketides, fatty acids, and phenols are the main groups found in nature.3

Phenolic structures with more than one –OH group, which are called polyphenols, are the largest class of secondary metabolites in vascular plants and they are mainly constituted from phenyl propanoids (coumarins, flavonoids, tannins, and lignin) from the shikimate–cinnamate pathway.4 Polyphenols are synthesized exclusively in vascular plants and in a wide variety of species (e.g. monocotyledons, dicotyledons, ferns, and conifers), and plant tissues (e.g. leaves, cone, seedpot, wood, blade, stipe, and bark).5 However, this review will focus on CTs as the most studied polyphenols nowadays.

In vascular plants, tannins are the third most abundant group of secondary metabolites after the two main wood constituent carbohydrates (in the form of cellulose and hemicellulose) and lignin. Despite that lignin is the widest distributed aromatic biopolymer in nature, the peculiar reactivity, diverse biological activity, high –OH/monomer content, and versatility of CTs in terms of physicochemical properties are recognized advantages for developing CT-based biomaterials. In addition, in soft tissues of woody plants such as leaves, needles, and bark, tannins can exceed lignin in abundance.5

Tannins are divided into two different classes: hydrolizable tannins (HT) and CTs. The former are found in chestnut (Castanea sp.), myrabolans (Terminalia and Phyllantus sp.), dividivi (Caesalpinia sp.), oak (Quercus sp.), eucalyptus (Eucalyptus sp.), and extracts of Asian Myrica sp. The limited distribution of HTs comprising only some dicotyledonian species, as well as their molecular structure and their low nucleophilicity decreased the interest in their utilization beyond traditional uses.7 CTs constitute more than 90% of the total world production of commercial tannins (>220.000 tons per year), and are chemically and economically more interesting as a bio-polymer.6,7 CTs and their flavonoid precursors are substantially concentrated in some barks and woods. They are highly abundant in barks of legumes (Acacia sp. (wattle or mimosa), Lotus sp., and Sericea sp.), conifers (Tsuga sp. and Pinus sp.), birch (Betula sp.), gambier (Uncaria sp.), and mangrove (Rhizophora sp.) species, and in the wood of Schinopsis sp. (quebracho). Tannin extracts mainly from barks of Pinus pinaster or Acacia sp.,2–4 and from Schinopsis sp. wood are commercially exploited. The main use of CTs comprise alternative medicine, the leather industry, and for adhesive preparation. Table 1 shows the CT content of selected tanniferous species.

Table 1 CT bark content of tropical and subtropical woody plants
Species Extractives (%) CTs (%) Ref.
a From wood.
Larix leptolepis 2.9–11.0 41.8–55.6 3
Rhizophora apiculata 12.8–20.2 43.5–50.8
Acacia mangium 18.4–37.9 46.7–52.9
Acacia auriculiformis 18.9–28.6 38.1–52.6
Acacia nilotica 49.5–55.0 11.5–13.0 5
Schinopsis balansae 23.0 21.0 6
Pinus radiata 16.0–20.0 12.0–16.0
Acacia mearnsii 42.0–51.0 35.0
Rhizophora mangle 45.0 36.0
Pinus pinaster 35.8–50.0 15–25 7
Pinus caribaea 17.6 9.2–15.4 9
Picea abies 33.0 8.0–11.0 11


1.2. CT chemistry

The monomer unit of CTs consists of flavonoid units (C15: flavan-3-ols or flavan-3,4-diols), which are condensed by 4 → 8 or 4 → 6 linkages. Their structure is a derivative of the basic C6–C3–C6 flavan unit, being two aromatic rings separated by three carbon atoms.12,13 The flavan rings are labelled A, C, and B, respectively, with a systematic numbering of carbon atoms (Fig. 1).
image file: c5nj02131f-f1.tif
Fig. 1 Chemical structure of the C15 unit in condensed tannins. C → C linkage between C15 units: C4 → C8 (A), and/or C4 → C6 (B).

The CT chemistry is dominated by the presence of multiple –OH groups with narrow but varying dissociation constants (pKa catechin-base CTs: pKa C-3′–OH: 9.02; pKa C-4′–OH: 9.12; pKa C-5–OH: 9.43; pKa C-7–OH: 9.58). In consequence, acid and base catalyzed rearrangements, nucleophilic substitutions, and electrophilic aromatic substitutions are common reactions.14,15

Nucleophilic substitution occurs at the –OH, introducing either acyl- or alkyl-groups. Acylation is commonly achieved by a reaction with acid chlorides or anhydrides,16 whereas alkylation is achieved with alkyl halides or epoxides.15 These reactions are the focus of this review and are described in detail in the sections below.

Activation towards electrophilic aromatic substitution is due to multiple phenol groups. The hydroxyl is a powerful electron donor (+M) especially under basic conditions (pH: 9–12) where the phenoxide ion is formed, which pushes electrons onto the A- and B-ring, thereby enhancing the nucleophilic potential in specific C15 reactive sites.

Under acidic conditions the reactivity of the rings towards electrophilic compounds (e.g. aldehydes) increases. Phenol groups are ortho and para directing. Hence, the meta phenol groups, which are common on the A-ring, both activate the C-6 and C-8 positions. In contrast, the vicinal di-phenol groups (catechol- or pyrogallol-type B-ring) cause a general activation on the B-ring rather than a specific activation of their corresponding C-ring atoms. Therefore, the B-ring is relatively unreactive during electrophilic substitutions in comparison to the A-ring. It's reactivity can be enhanced by selecting specific pH and catalyst type. However, from a practical point of view this is not efficient as the reactivity of the A-ring is increased as well, which increases the polymerization kinetics dramatically.

Further studies on the reactivity of CTs based on model compounds (resorcinol, phloroglucinol, and flavonoids) have been undertaken by selective reactions.15 These studies showed that a derivatization of the C-8 site or a linkage to another C15 unit at this position is ubiquitous for a derivatization or an advancing polymerization at the C-6 position. These model studies allowed an understanding of the general order of reactivity of the CT-monomer unit carbon atoms (8 > 6 ≫ 3′ > 4′ > 5′). However, the molar mass (Mw) and the steric hindrances conditioned by the conformation of CTs are factors that must be considered to explain the varied and sometimes unpredictable reactivity.17

Under acidic conditions two competing reactions occur: (1) the polymeric or oligomeric chain can be degraded to their C15 monomers and (2) the flavonoid units can condense. The first reaction is thought to be induced by the hydrolysis of the heterocyclic C-ring and its electrophilic attack upon a nucleophilic center of the A-ring.15

Under alkaline conditions, epimerization and rearrangements occur. Epimerization of catechin to epicatechin occurs in hot water or diluted alkaline solutions. This epimerization involves only a change in stereochemistry at the C-2 of the C-ring. Under strong basic conditions this epimerization reaches equilibrium in several hours.15–18 Rearrangement and loss of aromaticity of the C15 unit in alkaline medium is also reported.19

1.3. Physicochemical properties of CTs

There are few studies on the physical properties of CTs. In general, CTs are low-density amorphous non-crystalline pale yellow-slightly brown solid oligomers/polymers. Numerous tannins are optically active, while their stereochemistry is highly influenced by the C15 sequence. Dried tannin extracts are hygroscopic and exhibit a strong absorption in the UV-Vis region with several local maxima (λmax: 250 and 290 nm). At room temperature, CT-extracts are mainly soluble in water, short-chain alcohols (methanol, ethanol), and acetone aqueous solutions, and yield an acid environment with a sharp “puckering” and “astringent” flavor.

They yield colloids in water where they form viscous solutions (e.g. 40–45 wt%, 5 mPa s−1 at 20 °C) due to the strong hydrogen bridge formed between the –OH groups and the water molecules.7,20 The viscosity depends on the CT concentration, the dispersity, the topology and the ratio of hydrophilic/hydrophobic domains. The viscosity of some tannin extracts is increasing the difficulty to formulate CT-based materials or to spray CT-based resins, especially in the wood-based composites industry.7,10

In addition, CTs are thermo- and UV-labile compounds prone to oxidation. The residual weight at 600 °C oscillates between 40–50% as a consequence of decomposition processes between 220–350 °C.22 However, the thermal behavior is strongly associated with the Mw and the chemical structure.23,24 In general, CTs decompose in three main degradation steps according to their specific chemical structure and their linkages. The B-ring stripping process at temperatures between 200–250 °C opens up the C15 structure, which enables sequential degradations.26

Considering that CTs are bio-polymers, the glass transition temperature (Tg) is an important variable to be considered for CT-applications in materials engineering. Their Tg is strongly associated with Mw, the purity grade, and the moisture content. The Tg of native tannin oscillates between 120 and 180 °C. In general, it is difficult to detect Tg in CT samples. In addition, the reproducibility of the values is very low, regardless of the experimental parameter variation using DSC (Differential Scanning Calorimetry).37

2. Chemical modification of CTs

2.1. Main derivatization pathways

Derivatization is the process of chemically modifying a compound to produce a new structure. The new properties of these structures can be suitable for selected purposes. Thus, by selectively changing the properties through chemical modification, the range of potential utilization of the original compound can be increased.8,15 In case of CTs, modification of the native tannin extracts can overcome drawbacks like low solubility, high viscosity, and too high or too low reactivity. Modification allows further enhancement of these properties under consideration of the requirements for the selected application. According to the recent literature, derivatization seems to be an appropriate way to use this renewable product as a building-block for bio-polymer engineering and chemical synthesis.15,25–27

Several CTs have been subjected to a large variety of derivatization reactions. This section has been restricted to the identification and the discussion of the most relevant work according to the impact on physicochemical properties.

Tables 2 and 3 show the main modification reactions reported in polyphenols and the common derivatizing agents used during modification, respectively. According to the reactivity of the CT monomers discussed above, functionalization on the –OH groups, as well as the nucleophilic positions on the rings, are the most important sites for derivatization (see Fig. 1).

Table 2 Main modification reactions in CTs and chemically-related polyphenols
Reaction-type Derivatizing agent Ref.
Functionalization of the –OH group(s)
Esterification Acid chlorides, anhydrides, or carboxylic acids 15
Alkylation Acid chlorides or alkylene oxides 32 and 33
Carboxymethylation Chloroacetic acid 46
Carbamate formation Isocyanates 38 and 39
Dealkylation HCl 37
Oxidation/reduction Redox agent (K2Cr2O7, KMnO4), enzymes 40
Chelate formation Transitional metals (Fe3+, Cu2+, Zn2+) 41
Benzoylation Benzoic acid/H+ 7
H/deuterium exchange Deutherated solvents (D2O, EtOD) 42
Functionalization of the aromatic ring(s)
Alkylation Alkyl halides/K2CO3 or OH 43
Acetylation CH3COO 25
Methylolation/condensation Aldehydes/H+ or OH 44
Halogenation Halogens (Cl2, Br2, I2) 45
Sulphonation or sulphitation H2SO4 or Na2SO3 30
Amination NH3/H2O 47
Nitration HNO3/H2O or HNO3/AcOH 48
Nitrosation NaNO2/H+ 49
Phenolation Phenol 50
Depolymerization Mercaptan/H+ 34


Table 3 Common derivatizing agents used for chemical modification of CTs and chemical-related polyphenols
Reaction type Chemical Structure
TCA: tri-chloro acetic acid, TCP-TMA: N-(3-chloro-2-hydroxypropyl) trimethyl ammonium chloride, MDI: methyl di-isocyanate.
Esterification Lauroyl chloride image file: c5nj02131f-u1.tif
Acetic anhydride image file: c5nj02131f-u2.tif
Benzoic acid image file: c5nj02131f-u3.tif
trans-Esterification Vinyl octanoate image file: c5nj02131f-u4.tif
Alkoxidation Alkylene oxides image file: c5nj02131f-u5.tif image file: c5nj02131f-u6.tif
Alkylation Organic esters image file: c5nj02131f-u7.tif image file: c5nj02131f-u8.tif
Benzyl halides (R: Cl, Br, I), image file: c5nj02131f-u9.tif image file: c5nj02131f-u10.tif
Functional halide image file: c5nj02131f-u11.tif image file: c5nj02131f-u12.tif
Carbamate formation MDI image file: c5nj02131f-u13.tif


The degree of chemical modification (DS) depends on the polyphenol type, the derivatizing agent, and the experimental conditions. A wide range of strategies to get tailored derivatives as additives,28 fiber reinforcements,29 biologically-active compounds,30 thermosetting systems,31 foams,32 and copolymers33 are documented based on either one-step synthesis, or multiple-stage reactions.

In CT monomers there are three sites for derivatization. There are reactions that are selective towards one of these three derivatization sites.34 The A-ring can be derivatized at the C-5 or C-7 –OH group or at the C-6 position (see Fig. 1). On the B-ring, derivatization takes place mainly at the C-3′ or C-4′ –OH group (or additionally at the C-5′ –OH group in case of delphinidins or gallo-pyrogallols). The C-ring can be modified at the C-3 –OH and at the C-4 inter-C15 linkage. In C15 terminal units of CTs, derivatization may occur at the C-8 position.35

O-Acylation. Among all the reactions involving –OH groups of polyphenols, esterification is probably the easiest to carry out considering the reaction parameters and the reagents used.36 The O-acylation (esterification and trans-esterification) can introduce a wide range of functionalities by several different methods (Table 2), and derivatizing agents (Table 3). Reactions with acid chlorides and anhydrides are typical for CT modification. Several theses, papers, and patents outline methods for the preparation of functional derivatives.15,36,58

The synthesis of CT-ester derivatives can be achieved by several methods such as Fischer esterification, alcoholysis of acid halides or anhydrides,15 novel approaches with catalysts,51,52 microwave irradiation,53 and replacing halides by less toxic organic moieties are also reported.54 Fischer esterification is the classic route to prepare CT-esters. In the course of this reaction an acid reacts with –OH groups via nucleophilic acyl substitution. Aliphatic alcohols are relatively reactive in Fischer esterification, while in the case of hydroxylated phenols the reaction has to be enhanced by a dehydrating agent or by other strategies.

For instance, acid halides offer a higher reactivity in comparison to their corresponding carboxylic acids or anhydrides. Therefore, CTs can be easily acylated this way.55

The reaction is often carried out in the presence of hydroxides or organic (pyridine, ionic liquids, and amines) bases. The base-catalysis assists the activation of the –OH for an electrophilic attack (alcoholysis) of anhydrides in a comparable way as acid halides. Anhydrides are less reactive than acid halides, but the first one is often used for modification in order to gain structural information of products and intermediates.25,56 In comparison to low Mw polyphenols, CT derivatives are often synthesized under less drastic conditions according to temperature and reaction time and this is independent of the derivatizing agent.57

This is caused by the fact that alcoholysis reactions are often affected by steric hindrances. Bulky groups slow down the reaction, resulting in a defined –OH reactivity order, where the A- and B-ring aromatic –OH groups show a higher reactivity than the C-ring aliphatic –OH groups.15

However, beyond the noticeable differences between –OH reactivities, the conformational structure of CTs influences reactivity even at high temperatures and pressures. Tannin from mimosa and quebracho tannin are considered less-reactive CTs upon certain modification reactions (oxybutylation and oxypropylation) than pine tannin.82,83 The C4 → C6 linkages confer a peculiar “angular” configuration, which maximizes the steric hindrance upon alkylation. Such results are in contrast to the high reactivity of pine tannins (PRBT and PPBT), described a so-called “linear” conformational oligomer-chain. Anyway, it has to be taken into account that apart from steric hindrances, all the mentioned tannins exhibit very similar pKa values of their OH-groups.

On the other hand, specific CT reactivity is conditioned by the reaction type. For instance, hydroxypropylation reaction takes place on the A-ring at low derivatizing agent content, while isocyanates react preferably on the B-ring. So, the selectivity seems not to be related exclusively to the acidity values of –OH groups.15,27

Direct acylation of CTs with short-chain chemicals is strongly dependent on the experimental conditions. Pyridine and aliphatic amines are frequently used as catalysts, and chloroform, acetone, or aqueous solutions are the best choices as solvents, considering the low solubility of native CTs in organic liquids. O-Acylation of CTs with alkyl halide and long-chain esters in chloroform and polar solvents, respectively, provided low yields regardless of the stoichiometry of the derivatizing agent. Therefore, a high Mw of CTs requires a higher reactivity of the reaction agents in order to favor high yields (Table 5).49

While ionic liquids have been used for the derivatization of flavonoids and other chemically-related polyphenols15,29,59 the utilization of catalysts with polar functionalities such as carbodiimides, triazoles, and organic salts (chloroformates) has not been used in CT derivatizations. Table 4 shows a wide range of experimental conditions used for modification via O-acylation.60–67

Table 4 Acylation (esterification and trans-esterification) reactions in CT-monomer units and CTs
CT DA T (°C) t (h) Solv. Cat. Yield (%) Ref.
CT: condensed tannin, DA: derivatizing agent, T: temperature, t: time, Solv.: solvent, Cat.: catalyst, PRBT: radiata pine bark tannin, QT: quebracho tannin, MEI: methyl imidazole, n.i.: no information, VA: vinyl acetate, AH: anhydrides, Hld.: halides, THF: tetrahydrofuran, DMSO: dimethyl sulfoxide, Pen: pentanone, Acet: acetone, Chlor.: chloroform, Py: pyridine, DOX: dioxane, TEA: trimethylamine.
PRBT, QT VA 20 and 70 48 H2O, THF, DMSO KOH, TEA n.i. 16
PRBT AH 65 16 MEI n.i. 29
65 4 and 6 MEI n.i.
70 4 2-Pen/Acet. MEI 0–134 59
PRBT, QT Hld. 75 24 Chlor. Py 8–24 15
Nutgall tannin 22 2 Acet./Py TEA n.i 66
Wattle tannin 20 n.i. 1,4-DOX Py n.i 67


Table 5 O-Acylation (esterification and trans-esterification) of CTsa
CT DA (equiv.) T (°C) t (h) Solv. Cat. Yield (%)
CT: condensed tannin, PRBT: radiata pine bark tannin, QT: quebracho tannin, DA: derivatizing agent, T: temperature, t: time, Solv.: solvent, Cat.: catalyst, Chlor.: chloroform, THF: tetrahydrofuran, DMSO: dimethylsulfoxide, TEA: trimethylamine, Py: pyridine.a According to Hadley.15
Reaction with lauroylCl (esterification)
PRBT 3 75 24 Chlor. Py 10
Excess 24
QT 3 19
Excess 8
Reaction with vinyl laureate (trans-esterification)
PRBT Excess 20 48 THF KOH 7
QT 12
PRBT Excess 70 DMSO TEA 21
QT 8


Alkylation (etherification). CTs were etherified by several methods including the Williamson ether synthesis and alcoholysis of epoxides.68 The Williamson ether synthesis, where a tannin-alkoxide reacts with an alkyl halide based on the differences in nucleophilicity,69 is the more feasible method to prepare CT-ether derivatives. Considering the acidity of –OH in CTs, organic or inorganic catalysts can be utilized to increase the yield of this reaction (e.g. NaOH and K2CO3).15 A major drawback of C-alkylation is the possible side reaction for natural phenols and CTs. This reaction can be minimized by using polar aprotic solvents (DMSO, THF, pyridine, 1,4-dioxane, and acetone) instead of short-chain alcohols (MeOH and EtOH) and water. Polar aprotic solvents are able to increase the bonding strength between the proton and the oxygen of the phenolic hydroxyl group, which decreases the nucleophilicity of this group. In consequence, the carbon atoms ortho and para to the phenoxide compete in the nucleophilic displacement.69

Alkyl bromides are more reactive than alkyl chlorides in CT-etherification reactions, which lead to higher yields. As the reaction is induced by an SN2 mechanism, the primary alkyl halides are ideal for supporting E2 elimination on secondary substrates.15

Asymmetric ethers should be synthesized by reactions between sterically hindered alkoxides and the less sterically hindered halides.43,70 Other strategies comprising alkylation with carbonates using phosphines71 and nitrogenous buffers72 instead of alkali hydroxides and amines73 led to reasonable yields. However, the above-mentioned strategies have not been used for high Mw polyphenol modifications.

On the other hand, studies on the selective/non-selective alkylation of CTs described the use of ethylene oxide, alkyl sulphates, and halides as derivatizing agents.15 Other commercial CTs have been alkylated using agents like oxides, organic sulphates, acid chlorides, and halides. These alkylation strategies led to mono-alkylation of the C15 unit as the major reaction product.

Double-functionalization using mixtures of mercaptans and electrophilic agents have been used to increase the biological activity of the derivatives.75 This strategy yields a highly homogeneous product, because of the multiple reactive sites in polyphenols where bulky-halides induce the O-acylation instead of the alkylation.43 Therefore, short-chain derivatizing agents are preferable in comparison to bulky derivatizing agents.

Experimental conditions for the alkylation of CTs are listed in Table 6.

Table 6 Experimental conditions for CT alkylation
CT DA T (°C) t (h) Solv. Cat. Ref.
CT: condensed tannin, PRBT: radiata pine bark tannin, MBT: mimosa bark tannin, *: unspecified, DA: derivatizing agent, T: temperature, t: time, Solv.: solvent, Cat.: catalyst, DGD: diglycidyl di-epoxide, PGDE: polyglycidyl di-epoxide, TCP-TMA: N-(3-chloro-2-hydroxypropyl) trimethyl ammonium chloride, BO: 1-bromo octane, TCAA: trichloroacetic acid, C/F: condensation/formaldehyde, Resorc.: resorcinol, Catech.: catechol, GT: gelation time, OA: 1-octyl alcohol, MDF: dimethylformamide, n.u.: not utilized.
PRBT DGD and PGDE 80–90 GT H2O NaOH 74
Tannin* 1-BO 60 2 DMF K2CO3 66
Tannin* TCP-TMA C/F 48 1.5 H2O NaOH 75
MBT TCAA, Resorc., Catech., Na2SO3 86 and 90 2, 1.5 and 1.7 H2O, OA, H2O n.u. 30


Epoxides are used for the alkylation of CTs in a wide range of temperatures (20–200 °C). High pressure and the combination of less polar aprotic solvents (toluene and acetone) or aqueous solutions are used to enhance the derivatization reaction with alkylene oxides.15,37,38,76,77 The chemical structure of CTs, the Mw distribution, and the –OH pattern significantly affect the degree of substitution (DS).25–27,37,38

2.2. Alkylation of CTs with alkylene oxides

Alcoholysis of epoxides is a simple method to prepare hydroxyalkyl ethers. CT-ethers are used as chain-extenders in polyurethanes or as a co-monomer for polyester co-polymerization.77,78 The ring opening reaction of an epoxide is an efficient route to prepare ethers from high Mw polyphenols.79 On the other hand, alkylene oxides can be used for the preparation of simple phenol- and flavonoid-derivatives.15

The “oxypropylation” reaction and related short-chain oxyalkylation is an alternative pathway for the functionalization of CTs and lignin. This reaction has been extensively studied also for the modification of non-phenolic biopolymers and is one of the most attractive etherification alternatives considering the dramatic impact on properties.80,81

The general reaction mechanism of short-chain alkylene oxides, like propylene-, ethylene-, and butylene oxides, is determined by the high strain within the epoxide rings. This leads to their high reactivity compared to larger cyclic ethers. Epoxide ring scission can occur under neutral, basic, or acidic conditions, at any of the ring carbons. This high reactivity and flexibility in ring scission are the two major reasons for the feasibility of these short chain alkylene oxides for high Mw polyphenol modification.

Under basic or neutral conditions monoalkyl-substituted epoxides react predominantly at the less substituted carbon, due to steric effects (SN2 type mechanism). Under acidic conditions the phenate attack at the more substituted propylene oxide-carbon increases, but the attack at the less substituted carbon is still predominant.15,81

Acid-catalyzed ring opening occurs with a partial carbenium ion character on the reaction center. This is more in agreement with a SN1 type mechanism. The rate constant of alkoxylation (k) highly depends on the pKa of the –OH group.15 The order of reactivity is as follows: kphenolic(A-[thin space (1/6-em)]and[thin space (1/6-em)]B[thin space (1/6-em)]ring) –OH > kprimary –OH > ksecondary (C-ring) –OH > ktertiary –OH. However, CT-modification via acid-catalysis has not been reported yet.

Base-catalyzed ring opening reaction is the most common pathway in order to modify CTs with alkylene oxides. Typical catalysts are KOH or NaOH. Polyphosphazenium, aluminum tetraphenyl porphine, caesium hydroxide, and tertiary amines show excellent results for simple phenol and flavonoid modification, as well,50 while mineral bases still being the most useful catalysts for CTs modification.

The oxyalkylation of CTs has been used recently.15,81–83 Modification was achieved mainly under drastic experimental conditions (pressure and temperature),82 as well as at room temperature.81 The most comprehensive research concerning CT derivatization with PO and butylene oxide using a Parr reactor was conducted by Hadley15 and Arbenz and Avérous,82,83 respectively (Table 7). A positive linear relation between the PO molarity and the yield of oxypropylated CTs with an optimum ratio at 1[thin space (1/6-em)]:[thin space (1/6-em)]3 (OH[thin space (1/6-em)]:[thin space (1/6-em)]PO) was reported.

Table 7 Oxypropylation of CTs with alkylene oxides
Polyphenol PO (equiv.) Yield (%) Conditions
CT: condensed tannin, PRBT: radiata pine bark tannin, PPBT: pinaster pine bark tannin, QT: quebracho tannin, DA: derivatizing agent, PO: propylene oxide, BO: butylene oxide, MBT: mimosa bark tannin, *(CT/alkylene oxide ratio, w/w), CE: chain extended derivatives, THF: tetrahydrofuran, DMSO: dimethylsulfoxide, TEA: trimethylamine, n.i.: no information.
QT15 1 93 T: 110 °C t: 24 h

Catalyst: TEA

Solvent: toluene, alcohols

DA: propylene oxide (PO) Parr reactor

3 93
5 79
10 47 (CE)
20 28 (CE)
PRBT15 1 99
3 100
5 96
10 84 (CE)
20 58 (CE)
PPBT81 1 91 T: 22 °C t: 24 h

Catalyst: NaOH

Solvent: water

DA: PO (p: 1 atm)

2 86
3 89
4 92
5 94
6 85
Gambier CT,82,83 MBT,83 QT,83 PPBT83 40/60* n.i. (CE) T: ≥150 °C t: 1–24 h

Catalyst: KOH

Solvent: water

DA: PO,82 BO83 Parr reactor

30/70*
20/80*
10/90*


In addition, a strong relationship between the DS, and selected physico-chemical properties was established when CT-monomer units (catechin) were used.

Alternative strategies based on less drastic conditions (22 °C, 1 atm) have recently been described in the literature.25–27P. pinaster tannin oxypropylation improves the solubility of derivatives in several organic solvents, as well as the thermal resistance. Under the mentioned conditions the derivatives show non-oligomerization of the chain,81 while pressure conditions yield chain extended CT-derivatives.82,83

More complex epoxides such as di-epoxides of diglycidyl ether and polyglycidyl ether type have been used to cross-link bark tannins. The variation of the pH showed that the opening of the epoxide moiety was favored in basic media.74

The oxypropylation reaction is a common strategy to modify the physicochemical properties of natural polyol sources and this is particularly noticeable in the case of PPBT.81 The effect of internal plasticization of PO-modified CTs was extensively studied.32–36 The functionalization or chain-grafting degree improves the thermal properties and solubility.77,81

The solubility in non-polar solvents increases due to the derivatization and this extends the utilization potential of CTs in materials science, mainly for thermoplastic engineering. The type of alkylene influences the solubility of the derivatives, e.g. propoxylation leads to a better solubility than butoxylation.15,82

Some systematic assays have been performed regarding the effect of chemical modification on CT-solubility trends. Hydroxypropylation of highly purified P. pinaster bark tannin improved the solubility in short-chain alcohols, as well as THF, dioxane, and acetic acid, while in consequence the solubility in polar solvents (H2O, acetone, and acetonitrile) was reduced. The solubility trend of modified CTs follows a non-linear behavior for protic and a linear trend for aprotic solvents at the entire range of experimental modification degree (DS: 0.1–4.7). The trend suggests a cross-over point associated with specific interaction of the solvents proton with specific groups (–OH, grafting) as a function of the aromatic ring-type and conformation.

However the influence of the extraction method of the neat CT on CT-derivative solubility behavior has not been well-clarified. Hydroxypropyl derivatives from PRBT, obtained under pilot-plant conditions, exhibited a peculiar behavior. As expected, the modification changed the derivatives' solubility in several solvents significantly. However, the difference in solubility as a function of the DS was negligible. Concomitants such as carbohydrate moieties and terpene traces in the extract might affect the solubility trends in a high extent.

On the other hand, the Tg of CTs decreases due to oxyalkylation. The Tg-reduction is determined by two factors: (i) the intramolecular hydrogen bond rupture and (ii) the increasing free volume.56 Because of the advantage in solubility and thermal resistance CT-hydroxyalkyl derivatives were used for polyurethane foams,77 and for resin formulations.73,81

Hadley15 reported a significant decrease of the Tg value (ΔTg: ∼70 °C) when catechin was oxypropylated. This decrease followed a linear trend with increasing DS (1 → 5). However, it was difficult to establish a correlation between DS and Tg values when PRBT and QT were used.

In addition, PPBT modification decreased the Tg by 0.9 °C per wt% upon oxypropylation, while PRBT derivatives showed a reduction of 0.7–0.8 °C per wt%. The Tg rate reduction in both cases was similar to the plasticization rate reported for oxypropyl-lignin (1.0 °C per wt%). The aromatic content, the rigidity, as well as the Mw of the oligomers seemed to be the main factors that affect the Tg reduction in modified CTs.

2.3. Other derivatization reactions

Besides acylation and alkylation as traditional strategies to modify CTs, several other derivatization methods were used. Nucleophilic depolymerization,84 chelate formation,85 sulphitation86 and the Mannich reaction87 are four routes for CT functionalization that are described here (Table 8).
Table 8 Other derivatization reactions for CTs
CT DA T (°C) t (h) Solv. Cat. Yield (%) Ref.
CT: condensed tannin, DA: derivatizing agent, T: temperature, t: time, Solv.: solvent, Cat.: catalyst, *min, PAs: procyanidins, PRBT: radiata pine bark tannin, MBT: mimosa bark tannin, BMC: benzylmepcantan, PG: phloroglucinolysis, F: formaldehyde, n.i.: not informed. n.u.: not utilized.
PAs BMC 90 2* MeOH HCl 88 90
BMC and PG 80 and 20 50 and 48 EtOH and DOX/H2O AcOH and HCl 25–38 91
MBT CuCl2/NH3 20 n.i. H2O n.u. n.i. 92
PRBT F 80 0.3 HNO3 93


Nucleophilic depolymerization. This is a well-known derivatization reaction utilized for structural characterization of CTs. There are two derivatization reaction types: thiolysis and phloroglucinolysis. The former utilizes benzylmercaptans (thiol, R-SH) and the second utilizes phloroglucinol as a derivatizing agent. In both cases the C15 → C15 interflavanol linkage is cleaved and, thereby, the polymer chain is fractured. These reactions yield C-4 thioether or C-4 phloroglucinolate derivatives, respectively.34 The low Mw derivatives are analyzed by chromatography in order to gain insights regarding the chain-length and the C15-unit composition.88 The thermal stability of these derivatives dramatically changes after chemical modification. However, such reactions have not been explored as a strategy for developing novel bio-based-derivative materials.
Chelate formation. This is a complexation reaction between transition metal (d-electron acceptors) and vicinal –OH groups, which is typical for catechol based B-rings, 4,5-OR based C15 units, or flavonol-3,4-ol-based CTs. The chelation of CTs with cations can lead to a strong alteration of the biological properties of the modified CTs.41,105 The chelation reaction is often irreversible and can yield solid derivatives with enhanced biological properties.89 Beyond the bio-active property characterization of CT-complexes, factors affecting chelation stability such as (1) charge on the metal ion, (2) nature of the A- and B-ring as ligands, (3) stoichiometry of chelation, and (4) supramolecular structure have not been systematically studied.
Sulfonation. This is a common reaction that alters the solubility of CTs during the extraction process.7 This reaction was applied in the quebracho industry to improve the solubility of high Mw oligomers and hence to complete solubilization from the hardwood.9 This reaction affects the particle size of the tannin and enhances the selective solubility according to the lability of specific linkages in the A and B-ring. Sulphitation, which is generally accomplished by adding sodium sulphite into aqueous sodium hydroxide or sodium carbonate for CT extraction can change the solubility and viscosity of tannin.81,86 The effects of sulphitation on viscosity, acidity, and reactivity of CTs depend on the bark species.10 However, CT-sulphite derivatives that exhibit enhanced solubility in high polar solvents have not been used as a block polymer for biomaterial design.
The Mannich reaction. The Mannich reaction is an alternative route for CT functionalization. In many cases such derivatization is combined with a condensation reaction between CTs and formaldehyde. Carboxylic acids, cationic salts, and keto-acid functional groups are often graft-functionalities. This reaction comprises a nucleophilic addition of an amine to a carbonyl group followed by dehydration to the Schiff base.87 CTs are modified into derivatives with a wide range of applicability in the chemical and cosmetic industry.89 However, research has been focused on CT-derived material characterization instead of the understanding of key points, such as the mechanism involved when CTs are used as a block co-polymer.

Despite the lack of systematic studies providing valuable insight into the influence of chemical modification in a wide range of experimental conditions, Fig. 2 illustrates the most common applications of modified CTs as a function of the chemical structure of the grafted-chain.


image file: c5nj02131f-f2.tif
Fig. 2 Main chemical groups for tannin derivatives in function of the main applications.

The physicochemical properties of CT-derivatives are correlated to the DS, the type of the derivatized polyphenol, the derivatizing agent, and the specific sites where the derivatization occurs. In general, the derivatives exhibit considerable changes in the physicochemical properties in comparison to the corresponding native polyphenols.21,77

2.4. Solubility and viscoelastic properties

Acetylation and alkylation of CTs can alter the solubility,81 reactivity,81 the Mw distribution,56 and the thermal, and rheological properties15 to a high extent. Hydrophobic modifications improve the interactions of CT with less polar solvents, bio-polymers,28,29,59 and membranes,94 and reduce the water binding capacity of processed bio-polymers.25,26 In the context of CT-based polyurethane resin synthesis,33 partially benzoylated CTs alter the solubility of the highly polar tannins in order to match the solubility of the diisocyanates and thus to allow the polymerization reaction.38

2.5. Reactivity

CT reactivity was severely affected by modification. The influence of the grafting on the electronic configuration of aromatic rings (specifically the nucleophilic centers) was the reason for the dramatic reduction in reactivity due to the replacement of the aromatic –OH by secondary aliphatic moieties.

2.6. Thermal properties and reaction kinetics

A higher DS usually leads to an increased thermal stability and is associated with a higher degradation onset temperature.16 Long-chain grafting on P. radiata bark tannin and quebracho tannin retarded the decomposition.16 In detail, the substitution pattern on both the B-ring and the C15 → C15 linkage type affects the stability and the degradation onset. By benzoylating tannins that were used in polyurethane chemistry, the –OH/C15 ratio was reduced, which decreased the kinetics of the reaction with the isocyanates and led, in consequence, to a better network formation and enhanced thermal stability. While mechanisms that dominate the thermal decomposition of certain derivatives are known, the thermo-protective role of derivatives used as a block co-polymer in materials such as polyurethane foams and selected thermoplastics is unclear.

2.7. Biological properties

The effects of chemical modifications on the biological properties of flavonoid derivatives have been studied by antioxidant and anticancerogenic screening.17 Derivatization patterns in specific sites improved the therapeutic activity50 and the derivatives' feasibility for wood preservation against termites and fungi.81 CT derivatization by sulphonation, catecholation, and phloroglucinolation enhanced the bacteriostatic power and the biodegradability of CT-based rigid foams.30

3. Utilization of CTs with an emphasis on materials science

The most representative patents according to the impact of the content on databases (United States Patent and Trademark Office and the European Patent Office) and additional bibliographic sources were surveyed. This overview shows several patents (granted or pending) concerning applications of CT. Fig. 3 shows the evolution of published patents related to CT- and modified CT (MCT)-applications (Tables 9 and 10).
image file: c5nj02131f-f3.tif
Fig. 3 Evolution of published patents related to CT- and MCT-applications. Patents were surveyed based on the United States Patent and Trademark Office and the European Patent Office databases. CT: condensed tannin, MCT: modified condensed tannin.
Table 9 Most cited patents on CT/CT-derivative applications (1944–1999)
Patent (year) Assignee CTs-type/key word
CT: condensed tannin, MCT: modified condensed tannin, PU: polyurethane.
US2354672 (1944) Resinous Prod & Chem. Co. CT/resin
US2590760 (1952) Da Veiga Vinicio
US3254038 (1966) Borden Co. CT/adhesive
US3515556 (1970) Eastman Kodak Co. CT/photography film
US3856845 (1974) ITT MCT/bark etherification
US3886066 (1975) E. I. Du Pont De Nemours & Co. CT/non-porous material
US4090919 (1978) Tanabe Seiyaku Co., Ltd CT/protein immobilization
US4558080 (1985) Dearborn Chemical Co. CT/polymer
US4595736 (1986) Betzdearborn Inc. MCT/aqueous systems
US4944812 (1988) Henkel Corporation MCT/corrosion resistance
US5134215 (1992) Nalco Chemical Co. MCT/modified cement
US5158711 (1992) Mitsubishi Nuclear Fuel Co. MCT/adsorption
US5270083 (1993) Cecco Trading, Inc. MCT/preservative
US5659002 (1997) Nalco Chemical Co. MCT/Mannich polymers
US5698601 (1997) Bayer AG CT/PU foams
US5830315 (1998) Betzdearborn Inc. MCT/water purification
US5912037 (1999) W. R. Grace & Co., Conn. MCT/beverages
US5977287 (1999) Betzdearborn Inc. MCT/aqueous systems


Table 10 Recent most cited patents on CT/CT-derivative applications
Patent (year) Assignee CTs-type/key word
CT: condensed tannin, MCT: modified condensed tannin.
US6478986 (2002) Tanac S.A. MCT/coagulant
WO2004058843 (2004) Borden Chemical Australia Pty CT/binding agent
EP1437419A1 (2004) Tanac S.A. CT/bioactive
US20060084718 (2006) Stancliffe Mark R. CT/binder
WO2008079652A1 (2008) Gen Electric Co. CT/polymeric coagulants
US7611632 (2009) CT/co-polymer
US7998351 (2011) Vinod Kumar Rai CT/coagulants
US7998351B2 (2011)
US20120148740 (2012) Yang Chia-Wei CT/epoxy resins
WO2012162684A2 (2013) E.I. Du Pont De Nemours & Co. CT/rigid foam
WO2013010668 (2013) Silvachimica S.R.L., Universite' De Lorraine – Institut Enstib-Lermab
WO2012162653A3 (2013) E.I. Du Pont De Nemours & Co.
US20130252909A1 (2013) University of Iowa Research Foundation CT/HIV inhibitors
EP2805 943 A1 (2013) Albert-Ludwigs-Universität, Freiburg MCT/propylene oxide
US8642088 (2013) Wisconsin Alumni Res. Foundation CT/chitosan-composites
US20140093720 (2014) E.I. Du Pont de Nemours & Co. CT/rigid foam
EP 2617792 A4 (2015) Korea Advanced Inst Sci & Tech, Inno Therapy Inc. CT/adhesives


Before ca. 1980, CTs and CT-derivatives were used for several commercial applications like traditional leather tanning, adhesive additives, vinification, for pharmaceutical purposes, asphalt additives, dispersants, in medicine, as well as additives for tiles, floorings, and drilling muds.15 However, the most cited patents between 1940 and 1980 describe the use of native tannin for conventional applications.

As dispersants, CTs can alter the viscosity of various solids. It reduces the viscosity in clays to improve the manufacturing of bricks and pottery.18 A tannin based concrete additive provides improved plasticity, uniformity, flowability, and reduced shrinkage of the concrete.

The CT capacity to immobilize metals and macromolecules has been used for the design of tannin-based products (Tanfloc® and Silvafloc®) to clean water by flocculation,95–98 and to restore ecosystems. Collagen-tannin fiber membranes, which are selective to UO22+,99 tannin hydrogels,100 and tannin-rigid foams101 have been utilized to remove harmful chemicals from aqueous solutions, as well. CTs were also immobilized on substrates for the selective complexation with proteins. Such immobilization on inert and/or active matrices was also used to capture chemical compounds for further analysis,102–104 and in general to manufacture absorbent composites.105 Complexed CTs are used to inhibit fouling on vessel surfaces,106 and as polyphenol containing algae they are an efficient bio-based strategy to control harmful marine populations. In addition, applications were invented for purification technologies,40,75,92,93 antioxidant films,107 and cosmetics.55 Although the pharmaceutical use of CTs is well-known and applied, several investigations reported new properties and applications in alternative medicine in the last 60 years. Potential applications include prevention of chronic diseases, antimicrobial uses, antitumor activity,108 and cardio-protective effects.109

On the other hand, CT-derivatives were often used after 1980. Modified CTs play a key role in application areas where properties such as high miscibility and solubility in a non-polar medium, enhanced/suppressed reactivity, biological activity, and high thermal resistance are critical factors to consider. Derivatives instead of native CTs are useful in order to improve physicochemical properties and durability of thermoplastics. Functionalization enables high miscibility in biodegradable polymer blends such as poly(lactic) acid (PLA),16,18 and improves flexibility in polyurethane-based materials.81 Despite this specific chemical modification, oxyalkylation and partial acetylation are strategies to slow down the kinetics of complex reactions. There are no reports describing the understanding of the kinetics of the polymerization mechanisms, the polymerization stages, and the rheological behavior of these modified CTs. In addition, CT-modification might diversify properties of CT-based foams. Grafting of CT with aliphatic moieties may affect the ratio of hard/soft segments, which confers a dramatic impact on properties.

4. Concluding remarks and perspectives for the future

Considering the wide distribution of condensed tannins in vascular plants, these renewable natural products exhibit greater prospects in materials science beyond their traditional use in the leather industry and as components of adhesive resins.

The applicability of condensed tannins is restricted by high or low reactivity and poor solubility in organic solvents. A strategy to overcome this limitation is to modify the chemical structure by derivatization reactions. Among many possibilities, acetylation and alkylation, and here especially O-acylation using acid chlorides or anhydrides and alkyl halide, are the most frequent reactions. They change the physicochemical properties of the derivatives for their effective use in the industry and for purification technologies. Condensed tannin modification with conventional chemicals (e.g. anhydrides and halides) and propylene oxide affects drastically the physicochemical properties of the derivatives and provides high performance as macro building-blocks for materials engineering. The utilization of derivatives instead of native tannins has been described in several scientific journal articles and patents. The new trend points to the use of native and modified tannins as components of novel materials that are essentially based on: (i) the condensed tannin's properties as phenolic biopolymer with high molar mass, and (ii) the versatility conferred by the grafting functionalization. In the future, it is expected that tannin will be used in other research fields beyond medicine, adhesives, and the leather industry. Modified tannin as an additive and a block co-polymer for thermoplastics (ultraviolet-filter), tannin-based co-polymer with enhanced biological activity, and as macro building-blocks for hybrid resins preparation are promising research areas. Acetylation using specific derivatizing agents (non-linear alkyl halide and cyclic anhydrides) might yield useful alternative chemicals in order to tailor properties and diversify applications.

In the future, it is expected that tannins and tannin-derivatives may play a key role as nutraceutical products, as well as controlled release promoters, volatile organic compounds (VOCs) sequesters, and moisture retainers in promising research areas such as food science and agriculture.

Acknowledgements

D. E. García thanks the Basal Project (PFB-27) from the Concepción University, Chile.

References

  1. D. E. García, Pastos y Forrajes, 2004, 27, 1 Search PubMed.
  2. R. W. Hemingway, in Organic Chemicals and Biomass, ed. I. S. Goldstein, CRC, Boca Raton, Florida, 1981, p. 190 Search PubMed.
  3. R. Makino, S. Ohara and K. Hashida, J. Trop. Agric. Sci., 2009, 21, 45 Search PubMed.
  4. E. Haslam, The shikimate pathway, John Wiley & Sons, New York, 1974, p. 316 Search PubMed.
  5. P. J. Hernes and J. I. Hedges, Geochim. Cosmochim. Acta, 2004, 68, 1293 CrossRef CAS.
  6. E. Roffael, B. Dix and J. Okum, Holz- und Werktoff, 2000, 58, 301 CrossRef CAS.
  7. A. Pizzi, J. Macromol. Sci., Polym. Rev., 1980, 18, 247 CrossRef.
  8. A. Pizzi, in Monomers, polymers and composites from renewable resources, ed. N. Belgacem and A. Gandini, Elsevier Ltd., 2008, p. 179 Search PubMed.
  9. N. S. A. Derkyi, D. Sekyere, N. A. Darkwa and N. B. Boakye, Int. J. Chem. React. Eng., 2011, 9, 1 Search PubMed.
  10. M. C. Vieira, R. C. C. Lelis, B. C. D. Silva and G. D. L. Oliveira, Floresta e Ambiente, 2011, 18, 1 CrossRef.
  11. K. Kemppainen, M. Siika-aho, S. Pattathil, S. Giovando and K. Kruus, Ind. Crops Prod., 2014, 52, 158 CrossRef CAS.
  12. H. A. Stafford, Flavanoid Metabolism, CRC Press, Boca Raton, Florida, USA, 1990, p. 63 Search PubMed.
  13. M. N. Nurulhuda, L. T. Chew, N. M. Y. Mohd and R. M. A. Abdul, Holz Roh- Werkst., 1990, 48, 381 CrossRef CAS.
  14. C. Cren-Olivé, J. M. Wieruszeski, E. Maesc and C. Rolando, Tetrahedron Lett., 2002, 43, 4545 CrossRef.
  15. J. Hadley, Derivatisation of polyphenols, MSc thesis, The University of Waikato, N.Z. Hamilton, 2007, p. 157 Search PubMed.
  16. J. H. Bridson, W. J. Grigsby and L. Main, J. Appl. Polym. Sci., 2013, 129, 181 CrossRef CAS.
  17. S. Burda and W. Oleszek, J. Agric. Food Chem., 2001, 49, 2774 CrossRef CAS PubMed.
  18. K. Hashida, R. Makino and S. Ohara, Holzforschung, 2009, 63, 319 CrossRef CAS.
  19. K. Hashida and S. Ohara, J. Wood Chem. Technol., 2002, 22, 11 CrossRef CAS.
  20. L. J. Porter, in Plant Polyphenols, ed. R. W. Hemingway and P. E. Laks, Plenum Press, New York, 1992, p. 245 Search PubMed.
  21. I. Volf, I. Ignat, M. Neamtu and V. I. Popa, Chem. Pap., 2014, 68, 121 CAS.
  22. M. Pantoja-Castro and H. González-Rodríguez, Rev. Latinoam. Quím., 2011, 39, 107 CAS.
  23. R. W. Hemingway and J. J. Karchesy, Chemistry and significance of condensed tannins, Springer, US, 1989, p. 553 Search PubMed.
  24. M. Gaugler and W. J. Grigsby, J. Wood Chem. Technol., 2009, 29, 305 CrossRef CAS.
  25. D. E. García, W. G. Glasser, A. Pizzi, A. Osorio-Madrazo and M.-P. Laborie, Ind. Crops Prod., 2013, 49, 730 CrossRef.
  26. D. E. García, W. G. Glasser, A. Pizzi, A. Osorio-Madrazo and M.-P. Laborie, Holzforschung, 2014, 68, 411 CrossRef.
  27. D. E. García, W. G. Glasser, A. Pizzi, S. Paczkowski and M.-P. Laborie, J. Mass Spectrom., 2014, 49, 1050 CrossRef PubMed.
  28. C. Luo, W. Grigsby, N. Edmonds, A. Easteal and J. Al-Hakkak, J. Appl. Polym. Sci., 2010, 117, 352 CAS.
  29. W. J. Grigsby and J. F. Kadla, Macromol. Mater. Eng., 2014, 299, 368 CrossRef CAS.
  30. J. Ge, X. Shi, M. Cai, R. Wu and M. Wang, J. Appl. Polym. Sci., 2003, 90, 2756 CrossRef CAS.
  31. J.-M. Raquez, M. Deléglise, M.-F. Lacrampe and P. Krawczak, Prog. Polym. Sci., 2010, 35, 487 CrossRef CAS.
  32. J.-J. Ge, Bull. Kyushu Univ. For., 1998, 79, 21 Search PubMed.
  33. K. Hofmann and W. G. Glasser, J. Wood Chem. Technol., 1993, 13, 73 CrossRef CAS.
  34. P. Schofield, D. M. Mbugua and A. N. Pell, Anim. Feed Sci. Technol., 2001, 91, 21 CrossRef CAS.
  35. D. E. García, W. G. Glasser, A. Pizzi, C. Lacoste and M.-P. Laborie, Ind. Crops Prod., 2014, 62, 84 CrossRef.
  36. G. Sartori and R. Maggi, Advances in Friedel–Crafts acylation reactions: catalytic and green processes, CRC Press, 2010, p. 216 Search PubMed.
  37. M. Kessler, G. Ubeaud and L. Jung, J. Pharm. Pharmacol., 2003, 55, 131 CrossRef CAS PubMed.
  38. J.-J. Ge, CN98110709, 1998.
  39. J.-J. Ge, CN98100914, 1998.
  40. G. Alexandru, T. Lupascu, N. Timbaliuc and A. Meghea, Cent. Eur. J. Chem., 2013, 18, 1 Search PubMed.
  41. N. Bellotti, B. del Amo and R. Romagnoli, Procedia Mater. Sci., 2012, 1, 259 CrossRef CAS.
  42. E. Obreque-Slier, C. Mateluna, Á. Peña-Neira and R. López-Solís, J. Agric. Food Chem., 2010, 58, 8375 CrossRef CAS PubMed.
  43. K. S. Babu, T. H. Babu, P. V. Srinivas, B. S. Sastry, K. H. Kishore, U. S. N. Murty and J. M. Rao, Bioorg. Med. Chem. Lett., 2005, 15, 3953 CrossRef CAS PubMed.
  44. N. E. El Mansouri, Q. Yuan and F. Huang, BioResources, 2011, 6, 2647 CAS.
  45. W. R. Lotz, US5270083, 1993.
  46. Z. H. Huang, G. Z. Fang and B. Zhang, Chem. Ind. Forest Prod., 2007, 27, 27 CAS.
  47. F. Braghiroli, V. Fierro, A. Pizzi, K. Rode, W. Radke, L. Delmotte, J. Parmentier and A. Celzard, Ind. Crops Prod., 2013, 44, 330 CrossRef CAS.
  48. A. R. Pourali and A. Goli, J. Chem. Sci., 2011, 123, 63 CrossRef CAS.
  49. S. González-Mancebo, M. P. García-Santos, J. Hernández-Benito, E. Calle and J. Casado, J. Agric. Food Chem., 1999, 47, 2235 CrossRef.
  50. S. Laurichesse and L. Avérous, Prog. Polym. Sci., 2014, 39, 1266 CrossRef CAS.
  51. M.-Y. Chen and A. S.-Y. Lee, J. Chin. Chem. Soc., 2003, 50, 103 CrossRef CAS.
  52. M. Ikejiri, K. Miyashita, T. Tsunemi and T. Imanishi, Tetrahedron Lett., 2004, 45, 1243 CrossRef CAS.
  53. Z. Duan, Y. Gu and Y. Deng, J. Mol. Catal. A: Chem., 2006, 246, 70 CrossRef CAS.
  54. X. Yu, R. Zhao and G. Q. Liu, Se Pu, 2000, 18, 25 CAS.
  55. E. Perrier, A. Mariotte, A. Boumendjel and D. Bresson-Rival, Fr. Pat., 6235294, 2001 Search PubMed.
  56. S. S. Kelley, W. G. Glasser and T. C. Ward, J. Wood Chem. Technol., 1988, 8, 341 CrossRef CAS.
  57. J. Barry, G. Bram and A. Petit, Tetrahedron Lett., 1988, 29, 4567 CrossRef CAS.
  58. S.-C. Lin and M.-H. Rei, China Taipei Pat.,5136084, 1992 Search PubMed.
  59. W. J. Grigsby, J. H. Bridson, C. Lomas and J.-A. Elliot, Polymers, 1013, 5, 344 CrossRef.
  60. C. Wu, Transesterification, phase transition, rheology, and mechanical properties of blends of thermoplastic polyester and thermotropic polyester, PhD thesis, University of Akron, Department of Polymer Engineering, 2006, p. 330 Search PubMed.
  61. T. Weidlich, L. Dušek, B. Vystrčilová, A. Eisner, P. Švec and A. Růžička, Appl. Organomet. Chem., 2012, 26, 293 CrossRef CAS.
  62. B. Freedman, R. O. Butterfield and E. H. Pryde, J. Am. Oil Chem. Soc., 1986, 63, 1375 CrossRef CAS.
  63. M. I. Martins, R. F. Pires, M. J. Alves, C. E. Hori, M. H. M. Reis and V. L. Cardoso, Chem. Eng. Trans., 2013, 32, 817 Search PubMed.
  64. J. Sherma and B. Fried, Handbook of Thin-Layer Chromatography, CRC Press, 2003, p. 1048 Search PubMed.
  65. M. Sakai, M. Suzuki, N. Fumio and H. Yukihiko, Jp. Pat., 0618203A1, 2005 Search PubMed.
  66. E. Nomura, A. Hosoda, T. Hashizume, K. Kimura, K. Asahi, M. Yamazaki and H. Taniguchi, Jp. Pat., 2004307362, 2004 Search PubMed.
  67. A. P. Barbosa, E. B. Mano and C. T. Andrade, For. Prod. J., 2000, 50, 89 CAS.
  68. R. M. Roberts and A. A. Khalaf, Friedel–Crafts alkylation chemistry: a century of discovery, M. Dekker, 1984, p. 790 Search PubMed.
  69. L. G. Wade, Organic chemistry, Pearson Prentice Hall, 2010, p. 1262 Search PubMed.
  70. C. L. Wang, X. Zheng, W. D. Meng, H. Q. Li and F. L. Qing, Tetrahedron Lett., 2005, 46, 5399 CrossRef CAS.
  71. C. E. Summer, Jr., B. J. Hitch and B. L. Bernard, WO Pat., 1991016292A1, 1991 Search PubMed.
  72. A. Gissot, A. Wagner and C. Mioskowski, Tetrahedron, 2004, 60, 6807 CrossRef CAS.
  73. H. Dressler, Hydroxyalkylation of phenols or thiophenols with cyclic organic carbonates using triorganophosphine catalysts, US Pat., 5059723A, 1991 Search PubMed.
  74. R. Soto, J. Freer and J. Baeza, Bioresour. Technol., 2005, 96, 95 CrossRef CAS PubMed.
  75. D. B. Mitchell, R. L. Minnis, T. P. Curran, S. M. Deboo, J. A. Kelly, R. Patwardhan and W. T. Tai, US Pat., 5843337, 1998 Search PubMed.
  76. W. G. Glasser, C. Barnett, T. Rials and V. P. Saraf, J. Appl. Polym. Sci., 1984, 29, 1815 CrossRef CAS.
  77. W. G. Glasser, V. P. Saraf and W. H. Newman, J. Adhes., 1982, 14, 233 CrossRef CAS.
  78. W. G. Glasser, O. H.-H. Hsu, D. L. Reed, R. C. Forte and L. C.-F. Wu, in Urethane chemistry and applications, ed. K. Edwards, R. S. Graff, F. E. Bailey, S. Stinson, D. Klempner and W. G. Glasser, ACS Symposium No. 172, 1981, pp. 311–338 Search PubMed.
  79. G. Constantinescu, G. Cazacu and V. I. Popa, Cellul. Chem. Technol., 2005, 39, 201 CAS.
  80. A. de Menezes, D. Pasquini, A. Curvelo and A. Gandini, Cellulose, 2009, 16, 239 CrossRef CAS.
  81. D. E. García, Pinus pinaster (Ait.) bark tannin and its hydroxypropyl derivatives as building-blocks for bio-material design, PhD thesis, Freiburg University, Freiburg, Germany, 2014, p. 215 Search PubMed.
  82. A. Arbenz and L. Avérous, Ind. Crops Prod., 2015, 67, 295 CrossRef CAS.
  83. A. Arbenz and L. Avérous, RSC Adv., 2014, 4, 61564 RSC.
  84. S. Matthews, I. Mila, A. Scalbert, B. Pollet, C. Lapierre, C. L. M. Hervé du Penhoat, C. Rolando and D. M. X. Donnelly, J. Agric. Food Chem., 1997, 45, 1195 CrossRef CAS.
  85. M. Pérez, M. García, G. Blustein and M. Stupak, Biofouling, 2007, 23, 151 CrossRef PubMed.
  86. C. Amen-Chen, H. Pakdel and C. Roy, Bioresour. Technol., 2001, 79, 277 CrossRef CAS PubMed.
  87. M. R. Finck and P. E. Reed, US Pat., 5659002, 1997 Search PubMed.
  88. E. M. Kuskoski, J. J. Rios, J. M. Bueno, R. Fett, A. M. Troncoso and A. G. Asuero, Open Anal. Chem. J., 2012, 6, 1 CrossRef CAS.
  89. N. Bellotti, C. Deyá, B. del Amo and R. Romagnoli, Ind. Eng. Chem. Res., 2010, 49, 3386 CrossRef CAS.
  90. B. Labarbe, V. Cheynier, F. Brossaud, J.-M. Souquet and M. Moutounet, J. Agric. Food Chem., 1999, 47, 2719 CrossRef CAS PubMed.
  91. R. K. Gupta and E. Haslam, J. Chem. Soc., Perkins Trans. 1, 1978, 3, 892 RSC.
  92. H. Yamaguchi, M. Higuchi and I. Sakata, Mokuzai Gakkaishi, 1991, 37, 942 CAS.
  93. G. Palma, J. Freer and J. Baeza, Water Res., 2003, 37, 4974 CrossRef CAS PubMed.
  94. J. Li, Y. He and Y. Inoue, Polym. Int., 2003, 52, 949 CrossRef CAS.
  95. J. Sánchez-Martín, J. Bertrán-Heredia and G. Frutos-Blanco, Sep. Purif. Technol., 2009, 67, 295 CrossRef.
  96. J. Sánchez-Martín and J. Bertrán-Heredia, Appl. Water Sci., 2012, 2, 199 CrossRef.
  97. J. Beltrán-Heredia, J. Sánchez-Martín and M. C. Gómez-Munoz, Chem. Eng. J., 2010, 162, 1019 CrossRef.
  98. J. Beltrán-Heredia, J. Sánchez-Martín and M. Jiménez-Giles, Water, Air, Soil Pollut., 2011, 222, 53 CrossRef.
  99. L. Xuepin, M. Hewei, W. Ru and S. Bi, J. Membr. Sci., 2004, 243, 235 CrossRef.
  100. I. A. Sengil and M. Özacar, J. Hazard. Mater., 2008, 157, 277 CrossRef CAS PubMed.
  101. G. Tondi, C. W. Oo, A. Pizzi, A. Trosa and M.-F. Thevenon, Ind. Crops Prod., 2009, 29, 336 CrossRef CAS.
  102. T. Watanabe, T. Mori, T. Tosa and I. Chibata, J. Chromatogr. A, 1981, 207, 13 CrossRef CAS.
  103. T. Watanabe, Y. Matuo, T. Mori, R. Sano, T. Tosa and I. Chibata, J. Solid-Phase Biochem., 1978, 3, 161 CAS.
  104. J. Nkiliza, US Pat., 5808119, 1998 Search PubMed.
  105. S. González-Mancebo, M. P. García-Santos, J. Hernández-Benito, E. Calle and J. Casado, J. Agric. Food Chem., 1999, 47, 2235 CrossRef.
  106. A. Silkina, A. Bazes, J.-L. Mouget and N. Bourgougnon, Mar. Pollut. Bull., 2012, 64, 2039 CrossRef CAS PubMed.
  107. M. del M. Castro López, C. López de Dicastillo, J. M. López Vilariño and M. V. González Rodríguez, J. Agric. Food Chem., 2013, 61, 8462 CrossRef PubMed.
  108. S. Uesato, Y. Kitagawa, Y. Hara, H. Tokuda, M. Okuda, X. Y. Mou, T. Mukainaka and H. Nishino, Bioorg. Med. Chem. Lett., 2000, 10, 1673 CrossRef CAS PubMed.
  109. S. S.-M. Kim, S.-W. Kang, J.-S. Jeon and B.-H. Um, J. Med. Plants Res., 2012, 6, 2839 Search PubMed.

This journal is © The Royal Society of Chemistry and the Centre National de la Recherche Scientifique 2016