Biobased Amines: From Synthesis to Polymers; Present and Future

Nov 3, 2016 - Citation data is made available by participants in Crossref's Cited-by Linking service. For a more comprehensive list of citations to th...
29 downloads 0 Views 4MB Size
Review pubs.acs.org/CR

Biobased Amines: From Synthesis to Polymers; Present and Future Vincent Froidevaux,† Claire Negrell,† Sylvain Caillol,*,† Jean-Pierre Pascault,‡,§ and Bernard Boutevin† †

Institut Charles Gerhardt UMR 5253−CNRS, UM, ENSCM, 8 rue de l’Ecole Normale, F-34296 Montpellier Cedex 5, France INSA-Lyon, IMP, UMR5223, F-69621 Villeurbanne, France § Université de Lyon, F-69622 Lyon, France ‡

ABSTRACT: Amines are key intermediates in the chemical industry due to their nucleophilic characteristic which confers a high reactivity to them. Thus, they are key monomers for the synthesis of polyamides, polyureas, polyepoxydes, which are all of growing interest in automotive, aerospace, building, or health applications. Despite a growing interest for biobased monomers and polymers, and particularly polyamides, it should be noticed that very few natural amines are available. Actually, there is only chitosan and poly(lysine). In this review we present both fundamental and applied research on the synthesis of biobased primary and secondary amines with current available biobased resources. Their use is described as a building block for material chemistry. Hence, we first recall some background on the synthesis of amines, including the reactivity of amines. Second we focus on the synthesis of biobased amines from all sorts of biomass, from carbohydrate, terpenes, or oleochemical sources. Third, because they need optimization and technological developments, we discuss some examples of their use for the creation of biobased polymers. We conclude with the future of the synthesis of biobased amines and their use in different applications.

CONTENTS 1. Introduction 2. Background 2.1. General Characteristics of the Amine Function 2.1.1. Basicity 2.1.2. Nucleophilicity 2.1.3. Conclusion 2.2. Different Routes to Amine Synthesis 2.2.1. From Functions Bearing Nitrogen 2.2.2. From Alkyl/Aryl-Alcohol: Nucleophile Substitution and Addition/Reduction 2.2.3. From Alkyl/Aryl- and -Acyl Halide: Nucleophile Substitution and Reduction 2.2.4. From Aldehyde and Ketone: Addition/ Reduction Reactions 2.2.5. Short Summary on Amine Synthesis 2.3. Use of Amines for the Synthesis of Polymers 2.3.1. Reaction of Amines with Acids: Polyamides and Polyimides 2.3.2. Polymaleimides and Michael Addition 2.3.3. Polyepoxide Polymers 2.3.4. Polyureas and Polyhydroxyurethanes 2.3.5. Phenalkamide and Benzoxazines: Synthesis and Polymerization 2.3.6. Miscellaneous Reactions 2.3.7. Short Summary Concerning the Use of Amines for Making Polymers 3. Synthesis of Bio-Based Amines

© 2016 American Chemical Society

3.1. From Chitosan 3.2. From Amino Acids 3.2.1. Why Amino Acids? 3.2.2. Lysine as a Platform for Chemicals 3.2.3. Glutamic Acid 3.3. From Vegetable Oils and Derivatives 3.3.1. Amino-Acids/Esters Synthesis 3.3.2. Diamines 3.3.3. Miscellaneous Amines 3.4. From Sugar Derivatives 3.4.1. From Bio-Based Diacid 3.4.2. From Bio-Based Polyols 3.5. From Terpenes 3.6. From Cardanol 3.7. From Lignin Derivatives 4. Polymers Synthesis from Bio-Based Amines 4.1. Polymer from Modified Chitosan 4.2. Polyamides (PA) 4.2.1. Polypeptides 4.2.2. Polyamide 6,6, Polyamide 6, and Substitutes 4.2.3. Polyamides and Derivatives from Long Aliphatic Chains 4.2.4. Semiaromatic and Rigid Polyamides and Substitutes 4.2.5. Thermoplastic Elastomers with Polyamide Blocks 4.3. Epoxy Networks

14182 14182 14182 14182 14183 14183 14183 14183 14185

14186 14186 14187 14187 14187 14189 14190 14191 14191 14192 14192 14192

14192 14194 14194 14195 14197 14197 14197 14198 14199 14199 14199 14199 14200 14201 14203 14205 14205 14205 14205 14206 14207 14207 14209 14209

Received: July 26, 2016 Published: November 3, 2016 14181

DOI: 10.1021/acs.chemrev.6b00486 Chem. Rev. 2016, 116, 14181−14224

Chemical Reviews 4.3.1. Diamines as Hardeners for Epoxies 4.3.2. Hardeners from Chitosan 4.3.3. Hardeners from Amino-Acids 4.3.4. Hardeners from Vegetables Oils 4.3.5. From Terpenes 4.3.6. From Cardanol 4.3.7. Aromatic Amine Hardeners 4.4. Miscellaneous Polymers 4.4.1. Other Thermoplastic Polymers 4.4.2. Miscellaneous Thermosets Polymers 5. Conclusions and Future Trends Author Information Corresponding Author Notes Biographies Acknowledgments References

Review

There are many opportunities to develop biobased amine monomers for preparation of renewable polymers and materials. Nevertheless, the biggest development in the near future will certainly come from the regulation changes such as the environmental directives for reducing volatile organic constituents (VOC) or the “Registration, Evaluation and Authorization of Chemicals” (REACh). These regulations will oblige the replacement of some products and more specifically amines such as methylene dianiline. Biobased compounds will not be the only solution for these policy pressures, but they will be an opportunity. To our knowledge, even if there are some specific reviews on the use of a few biobased amines in materials,5 there is no review depicting the various routes for the synthesis of biobased amines from available renewable resources and their advantageous use in polymer chemistry. However, some authors already reviewed amines from vegetable oils but did not take into account some interesting natural aromatic resources such as lignin or cardanol. More generally, no review reports the synthesis of biobased amine curing agents or material properties to discuss the envisaged applications.6 Therefore, our review is essential and complementary to previous ones. Thus, the aim of this review is to present both fundamental and applied research (including industrial patents) on the synthesis of biobased primary and secondary amines (we do not study tertiary or quaternary amines) with current available biobased resources, their use will be described as a building block for material chemistry. Biorefinery processes and retreatment of biomass will be out of our scope. First, we will recall some background on the synthesis of amines, including the reactivity of amines. Second we will focus on the synthesis of biobased amines from all sorts of biomass, from carbohydrate, from terpenes, or from oleochemical sources. Third, because they need optimization and technological developments, we will try to discuss some examples of their use for the creation of biobased polymers. Finally, we will give our opinion on what could be the future of the synthesis of biobased amines and their use in different applications.

14209 14209 14209 14209 14210 14210 14210 14210 14210 14211 14212 14212 14212 14212 14212 14213 14213

1. INTRODUCTION Amines are key intermediates in chemical industry due to their nucleophilic characteristic which confers a high reactivity to them. Amines could be used in various applications such as agrochemicals, drugs, detergents, lubricants, food-additives, and also polymers. Indeed, amines are of crucial interest in polymers and composites fields. Thus, they are key-monomers for the synthesis of polyamides, polyureas, polyepoxydes, which are all of growing production in automotive, aerospace, building, or health applications. As an example, the global production of polyamides was around 7 Mt/y in 2015 and is expected to increase by 3%/y within 2020, and the production of epoxy polymers (4 Mt/y in 2015) is following the same tendency. Moreover, the development of polyhydroxyurethanes will also entail an increasing demand of amines as hardeners of polycyclic carbonates. The most classical syntheses are a direct reaction between ammonia and halogenoalkanes or reductions of nitro- or nitrile functions. Other routes were developed such as hydroamination followed by a reductive amination. Due to their high reactivity, amines are generally available as amides or salts to avoid any carbonation with CO2.1 Ammonia plays a key-role in the synthesis of amines, and most of the commercial amines are synthesized directly or indirectly from ammonia. However, amination of alcohols is also a route of increasing interest since water is the main byproduct with some traces of secondary amine. Despite a growing interest for biobased monomers and polymers, and particularly polyamides, it should be noticed that very few natural amines are available. Actually, there are mainly two polyamines: poly(ε-L-lysine) produced by Streptomyces microorganism2 and chitosan, issued from deacetylation of crustacean or insect chitin. However, as for other monomers, both scientific and industrial communities are looking for either biobased monomers or polymers and especially aliphatic or aromatic polyamines which are very useful in polymer synthesis. Indeed, utilization of biomass represents an outstanding opportunity for our next generation feedstocks for the chemical industry. According to Pike Research, green chemistry markets represented a market opportunity of $2.8 billion in 2011 and are expected to reach $98.5 billion by 2020.3 Judicious use of the 12 principles of green chemistry4 (waste as a resource, biobased or renewable materials) is expected to save the chemical industry $65.5 billion by 2020.

2. BACKGROUND 2.1. General Characteristics of the Amine Function

The reactivity of an amine depends both on its basicity and nucleophilicity. These two concepts are fundamentally different, but generally it is possible to assess that nucleophilicity increases with basicity. However, and it is often the case, experimental conditions such as the nature of solvent and substituents close to the amine function could modify this rule. 2.1.1. Basicity. First, amines are amphoteric compounds. Hence they are able to exchange a proton both with a base or an acid. These notions of basicity or acidity are based on equilibrium systems that are thermodynamically controlled. Proton exchanges are so fast that the kinetic aspect is not involved in acido-basic equilibrium. In amines, the basicity increases with the number of electro-donating substituents linked to the amine function. Indeed, this effect increases the availability of the doublet. Thus, aliphatic amines are more basic compared to ammonia. Moreover, aromatic amines are less basic than aliphatic amines due to the mesomeric effect of aromatic ring which decreases the availability of the doublet. The case of aromatic amines is more complex because the basicity depends also on substituents attached to the aromatic ring. In a nutshell, the more electro-donating are the 14182

DOI: 10.1021/acs.chemrev.6b00486 Chem. Rev. 2016, 116, 14181−14224

Chemical Reviews

Review

could react twice (with each proton of the amine) and could easily lead to cross-linked polymers. Hence, amines present great interest, due to their reactivity, for the synthesis of polymers. First, the main reactions used for the synthesis of primary and secondary amines will be reported (tertiary amines are excluded from this review because of their lack of interest for the structure design of polymer). Then, the use of amines as monomers for polymers and materials synthesis will be reported, i.e., their reactivity with some other functional groups.

substituent on the aromatic cycle, the less the doublet is contributing to the mesomeric form and, as a consequence, the pKa is rising. Hence, the pKb determines the basicity of amines, and the pKb increases with the basicity of amines. 2.1.2. Nucleophilicity. The nucleophilicity is a kinetic concept that defines the nucleophilic strength of a nucleophile. A strong nucleophile presents an electronic doublet which can easily form chemical bonds. Nevertheless, nucleophilicity of an amine is complex to determine since it depends on several factors such as its charge, the nature of the chemical groups present in the substituent attached to the amine or close to the amine function, and the nature of the reaction solvent. The charge is one of the easiest factors to evaluate. Amines with a negative charge are more nucleophilic than their neutral forms. Conversely, amines with a positive charge are less nucleophilic than their neutral form. Solvents have also a huge influence on the nucleophilicity of amines7,8 and have also an influence on basicity.9 Indeed, protic solvents (such as alcohols, water, m-cresol, carboxylic acids, ...) could form hydrogen bonds and/or give a proton to the amine. In some cases, the amine is surrounded with molecules of solvent and has some difficulty to diffuse and attack a nucleophile site. In that case the nucleophilic character of the amine is reduced. Hence, the influence of solvent on the nucleophilicity of amines depends on the following order: aprotic polar solvent > apolar solvent > protic polar solvent. The last parameter to take into account is the nature of the substituents attached to the amine or close to it, which modify the nucleophilic character of the amine. Aliphatic amines exhibit often the most nucleophilic character. As said before for the basicity, the availability of the doublet increases the nucleophilic character. Thus, aromatic amines are less nucleophilic than aliphatic ones. Some studies have been performed to determine the influence between basicity and nucleophilicity of aliphatic and aromatic amines.10−12 Thus, triethylamine (pKa = 10.75), with a third ethyl substituent, presents a lower nucleophilic character than diethylamine (pKa = 11.05); therefore, triethylamine is often preferred as a base. Indeed, it is more difficult for a tertiary amine to create strong bonds. Indeed, in the case of primary or secondary amines, the loss of a proton after reaction allows stabilizing the nitrogen atom and the chemical bond at the same time, which is not the case for tertiary amines. After reaction a tertiary amine has to return into its initial state or form a strong pair of ions. The overloading could also be far from the amine function. For instance, if there is a methyl substituent on the carbon in α position relative to the nitrogen, the reactivity considerably decreases. A study on amines’ reactivity for the ring-opening reaction of cyclic carbonate allowed again to demonstrate these results. Finally an α effect has been discovered in 1962 by Pearson and Edwards.13 This effect corresponds to the increase of nucleophilicity when the nucleophilic center’s adjacent atom has a free doublet. 2.1.3. Conclusion. The objective of this part was to shed light on the importance of the choice of amine for the reactivity. Indeed, reactivity of amines has strong influence on the formation of polymers/materials. For instance in the case of polyamides it is more interesting and easier to use primary amines, which are more nucleophilic and could then create hydrogen bonding, compared to secondary amines which react more slowly and would not participate to hydrogen bonding, which is a significant characteristic of polyamides. Moreover, for the reaction of epoxide or acrylate with amines, primary amines

2.2. Different Routes to Amine Synthesis

Several routes are available to synthesize amine functions and consist mainly of nucleophilic (addition or substitution) and reductive reactions. The following paragraphs present different routes to amine synthesis. The first one reports the synthesis routes from functions with nitrogen atom(s). The three following paragraphs report the synthesis routes from function without nitrogen atom (alcohol, halide, and aldehyde or ketones). Indeed, it has to be noticed that some reaction pathways from function without nitrogen lead to intermediate that bears nitrogen atom(s). Therefore, it is more convenient to focus first on functions with nitrogen. All routes are not described according to their relevance; however, for each one the utility is mentioned. Some of them use hazardous reactants or conditions and are not applicable to industry. For each route the hazards and the use at industrial scale is, if possible, discussed. 2.2.1. From Functions Bearing Nitrogen. 2.2.1.1. From Amide: Hydrolysis, Reduction or Hofman Degradation. Amide function could be transformed easily into amine14 by amide hydrolysis reaction with water in acidic15,16 or basic conditions17 (Scheme 1(1)). This reaction is an equilibrium in some conditions. Scheme 1. Hydrolysis (1), Reduction (2), or Hofmann Rearrangement (3) of Amides

The reduction of an amide with metal halide (Scheme 1(2)) leads to an amine. For instance Micovic et al.18 used lithium aluminum hydride as reducing agent. Depending on the substituents, a primary (R1 = R2 = H) or a secondary amine (R1, R2 = H) is obtained. In the case of nonsubstituted amide, a Hofmann rearrangement could occur in the presence of bromide and sodium hydroxide (Scheme 1(3)). The substituent of carbonyl is transferred onto the nitrogen atom and gives the corresponding isocyanate, which is then hydrolyzed into amine.19 The particularity of this way is the disappearance of the carbon of carbonyl at the end of the mechanism, contrary to the amide reduction (2). 14183

DOI: 10.1021/acs.chemrev.6b00486 Chem. Rev. 2016, 116, 14181−14224

Chemical Reviews

Review

aldehyde and one equivalent of ammonia.34 Eight years later, he described a method for producing 2-aminopyridine derivatives from the reaction between pyridine and sodium amide called Chichibabin reaction.35,36 This reaction was then extended to quinoline and isoquinoline derivatives (Scheme 4).37

The isocyanate hydrolysis into amine is a well-known reaction, usually considered as a side-reaction (Scheme 2).20 Scheme 2. Mechanism of Isocyanate Hydrolysis

Scheme 4. Chichibabin Reaction Isocyanates could easily react with nucleophiles as alcohol, amine, water, and sodium hydroxide.21 Carbon dioxide is released when isocyanates react with water or bases. The hydrolysis could be assimilated to the amide hydrolysis because the mechanism goes through a carbamic acid or amidic acid. However, the reaction does not stop, and some ureas are obtained as byproduct of amine synthesis. The main limitation of this reaction, for industrial use, is the formation of isocyanate groups, which are often toxic and CMR (carcinogenic, mutagenic, and reprotoxic). Moreover, the amine formed during the hydrolysis reacts faster than water with isocyanate. Thus, this reaction will also generate urea coproduct in addition to the expected amine. Rasshofer et al.22−24 proposed a catalytic route that allowed us to reach 86% conversion of NCO into amine. Similar, uncatalyzed reaction in dioxane gives only 29% conversion. The used catalysts were bicarbonates, tertiary amines, arylaliphatic or cycloaliphatic amines. Also, it has to be noticed that amides are often synthesized from amines and acids; therefore, this route to amines synthesis is useless, since corresponding amine is generally commercially available. Some specific and rare mechanisms exist and will be discussed in future parts. 2.2.1.2. From Nitro: Reduction. The nitro reduction is another easy route to synthesize a primary amine. However, this reaction is often performed on aromatic compounds and thus gives aromatic amines. Indeed, nitration of aromatic compound is easier, for instance with nitric acid.25,26 The reduction of this nitro function yields the corresponding amine. Reduction is generally performed with reducing agent such as solid metals in the presence of acidic compound27 or oxide metal,28−31 but hydrate hydrazine as substitutive agent (Scheme 3) could be used also. Only few examples on aliphatic nitro compounds have been studied.32,33

This reaction is most of the time used for the modification of pyridine derivatives which are difficult to functionalize. The main difficulty of this amination consists of the use of sodium amide which is very reactive with water. The use at industrial scale has not been reported even if some patents have been published on this reaction.38 2.2.1.4. From Nitrile: Hydrogenation, Ritter Reaction, and Reduction. The transformation of nitriles into amines is usually performed by a two-step hydrogenation: the first consists of the hydrogenation of nitrile into a primary aldimine intermediate and the second one leads to the amine (Scheme 5(1)). Sometimes, during the first hydrogenation reaction, a release of ammonia occurs which entails the formation of a secondary aldimine. In that case, the second hydrogenation is also performed on this intermediate, yielding a secondary amine (Scheme 5(2)).39,40 Several catalytic systems based on metals have been tested.41−47 These catalysts could be used under heterogeneous48,49 or complex50,51 forms. Kukula et al.52 transformed unsaturated nitriles into corresponding amines by catalyzing the reaction with cobalt boride (CoB) or nickel boride (NiB) catalysts. A doping with chromium allows limiting the hydrogenation of the double bond even in the presence of nitrile functions. The nitrile reduction could also be carried out with hydride such as lithium aluminum hydride.53 Saavedra et al.54 used a system of indium trichloride/borohydride to reduce aromatic and aliphatic nitriles. Another remarkable reaction is the transformation of nitrile derivatives into specific amide (which could be transformed into amine cf. section 2.2.1.1). This is the Ritter reaction (Scheme 5(3)),55 which needs electrophilic alkylating reagents in the presence of a strong acid. The first Ritter reaction was performed with alkene as the alkylating agent.56 Later, the reaction has been extended to alkyl (primary, secondary, and tertiary) alcohol.55,57 Acetate t-butyl has been used as alkylating reagents for such an amine synthesis pathway.58 More recently some advances have been obtained on catalytic systems for the Ritter reaction by using sulfonamide derivatives or metal halides.59 In order to convert a nitrile into a secondary amine, Werkmeister et al.60 reported the N-monoalkylation of the primary amine formed (Scheme 5(1)) with a system alcohol/ NaOH. This reaction was performed by varying the initial nitrile, the catalytic system, and the alcohol used. The best yield (99%) was obtained for the reduction of benzonitrile with the complex RuCl2(PPh3)3 followed by the reduction with system isopropanol/NaOH at 120 °C for 6 h. All these routes to the synthesis of mono- or disubstituted amines are performed on functions which already have a nitrogen atom, such as amide, nitrile, isocyanate, nitro, or

Scheme 3. Amine Synthesis by Reduction of Nitro Group

In industry, these aromatic monoamines are reacted with formaldehyde in the presence of an acid catalyst to give aromatic diamines. A harder hydrogenation leads to cycloaliphatic diamines. 4,4′-Methylene dianiline (MDA), the most used aromatic diamines, is suspected to be highly toxic (CMR); therefore, a lot of aromatic diamines with different substituents have been proposed, such as 4,4′-methylenebis(2,6-diisopropylaniline), 4,4′-methylenebis(2-isopropyl-6-methylaniline), or 4,4′-methylenebis(3-chloro 2,6-diethylaniline). However, these diamines are solids. For this reason a commercial liquid diamine, diethyltoluenediamine (DETDA), has been prepared from the corresponding dinitro compounds. 2.2.1.3. From Pyridine Derivatives: Chichibabin Reaction. Chichibabin discovered in 1906 a pseudofour-component pyridine synthesis from three equivalent of an enolizable 14184

DOI: 10.1021/acs.chemrev.6b00486 Chem. Rev. 2016, 116, 14181−14224

Chemical Reviews

Review

Scheme 5. Reduction of Nitrile Compounds, Leading to a Primary (1) or to a Side-Reaction Giving a Secondary (2) Amine and Ritter Reaction with Alkene (3)

of three carbons between the oxygen and the nitrogen (Scheme 6).

pyridine. However, it is possible to obtain amines from other reactive functions such as alcohol, halide, ketone, or aldehyde. Such transformations have been carried out to propose amines from various reactive functions, with a high availability, including in natural products. 2.2.2. From Alkyl/Aryl-Alcohol: Nucleophile Substitution and Addition/Reduction. 2.2.2.1. Nucleophile Substitution. Alcohol amination was reviewed by Hamid et al.61 and by Guillena et al.62 This amination is already performed at industrial scale with ammonia and a heterogeneous catalyst. Millions of tons of amines are produced according to this method which could also be carried out on with primary amines instead of ammonia in order to obtain secondary amines.63 This route is really attractive since water is the only byproduct.64 However, because heterogeneous catalysts have a limited activity, reaction conditions are tough65 (T > 200 °C) and the selectivity is hard to control.66 Thus, other heterogeneous catalysts have been already studied.67−92 Moreover, alcohol aminations with homogeneous systems were also studied in the 1980s by Grigg et al.93 and by Watanabe et al.94 Some examples used ruthenium or iridium complex95−99 and even biocatalysts.100 Nevertheless, primary amines formed during the reaction are often more reactive than ammonia itself; thus, they react on alcohol (nucleophile substitution and release of water) to yield secondary, and sometimes, tertiary amines.66 Therefore, some methods have been reported to give selectively only primary amines. Thus, Milstein et al.101 described the selective amination of a primary alcohol into primary amine thanks to a ruthenium pincer catalyst. This reaction could be carried out in soft conditions, in toluene, in water, or even without solvent. It gives excellent yields (between 60 and 95%). Shimizu et al.102 did this amination on primary alcohol with a heterogeneous catalysts (Al2O3 supported on nickel). To synthesize ethylenediamine, an industrial route involves the reaction of ethanolamine and ammonia with Ni-catalyst. Since 2010, few teams succeeded in synthesizing primary amines from secondary alcohol with good yields.103 For instance, Imm et al.,66 used a commercially available catalyst based on rhutenium metal. The synthesis of ethanolamines is specific. Monoethanolamine (MEA), diethanolamine (DEA), and triethanolamine (TEA) are produced by the reaction between ammonia and ethylene oxide in the presence of water. Based on operational conditions MEA, DEA, or TEA are produced through three parallel-consecutive competitive reactions.104 2.2.2.2. Addition/Reduction. Another method consists of the addition of acrylonitrile on alcohol followed by the reduction of the obtained nitrile into amine (cf. section 2.2.1.4). This reaction has been patented on trimethylol propane by Akzo Nobel.105 Such a method alters the chemical environment of the alcohol because it consists of the addition

Scheme 6. Alcohol Amination by Acrylonitrile Addition Followed by Reduction/Substitution Mechanism

These two last methods (nucleophile substitution and addition/reduction) are most often used to create amine industrially and could be illustrated with the synthesis of poly(propylene oxide) (PPO) diamines (commercially available from Huntsman). Amines are obtained by cyanoacrylation reaction of PPO to give reactive H2N-(CH3)3-O-PPO-O(CH3)3-NH2 structures which could react on themselves, as described before, to give secondary amines. However, some companies modified this process since acrylonitrile is a very harmful reactant (CMR) which explains the low commercial availability of amine nonsubstituted in α position. Thus, the method based on the selective amination of alcohols with ammonia is preferred to create PPO diamines, which leads to primary amines α-substituted with a methyl group. However, this method reduces considerably the reactivity of the amine for other reaction but it avoids the intermolecular reaction giving secondary or tertiary amine. This is precisely one of the most significant problems for the synthesis of amines from alcohols (which is, industrially, the most used method). Indeed chemical industry has to choose between either the synthesis of a non-αsubstituted highly reactive primary amine which entails several purifications to remove secondary amines by-produced during the process or the easier synthesis of α-substituted primary amines with reduced reactivity. Hence industry choses naturally the cheapest routes which is the synthesis of α-substituted amines at the cost of the amine’s reactivity, creating a lack of primary reactive amines for polymer synthesis. 2.2.2.3. Bucherer Reaction on Phenol. The transformation of aromatic alcohol has already been carried out on naphtol derivatives. The Bucherer reaction provides a relatively simple method for the conversion of hydroxyaromatic compounds into primary and secondary aromatic amines.106 It proceeds in an aqueous phase in the presence of sulfurous acid or its salts, and is used for the preparation of many naphthalene derivatives (Scheme 7). Scheme 7. General Ritter Reaction Mechanism on Naphtol

14185

DOI: 10.1021/acs.chemrev.6b00486 Chem. Rev. 2016, 116, 14181−14224

Chemical Reviews

Review

Lastly, Smiles rearrangement,107,108 an intramolecular nucleophilic aromatic substitution, could be used to transform a phenol function into aniline in two steps. 2.2.3. From Alkyl/Aryl- and -Acyl Halide: Nucleophile Substitution and Reduction. 2.2.3.1. Nucleophile Substitution with Ammonia. This amination reaction is performed by reacting alkyl halide with ammonia (Scheme 8).109

catalysts, followed by a hydrolysis (Scheme 10). The formation of the phthalimide derivative is often carried out to protect the primary amine. Thus, phthalic anhydride is used with amines and then the acid-amic is cyclized in phthalimide. Then, hydrolysis enables the release of the amine.117−119 If the reaction is performed with NH3, the phthalimide (NH) is obtained and could react as nucleophilic agent on alkyl halide in basic condition; this is the Gabriel synthesis. Then, the hydrolysis of the product gives the corresponding primary amine.120,121 Also a particular reaction could be carried-out on acyl halide to give an amine through a singular rearrangement. The last three methods, using alkyl or aryl halide present a significant issue, which is the same as for alcohol amination. Indeed, the created amine is also nucleophilic and could react on the halide reactant as well and lead to secondary or tertiary amines byproducts. 2.2.3.4. Curtius Rearrangement. As usually isocyanates are the result of a reaction of an amine with phosgene, the isocyanate hydrolysis to obtain the initial amine is nonsense. However, if another route for the synthesis of the isocyanate is used, it could present some interest. Acyl azides compounds are synthesized by azidification of an acyl chloride with sodium azide122 or, more easily, by azidification of a carboxylic acid with diphenylphosphoryl azide as it has been done by Shioiri et al.123 Both ways are performed in the presence of triethylamine. Once synthesized, this acyl type compound goes through a rearrangement called Curtius or Hoffman rearrangement under high temperature.124 It gives the corresponding isocyanate which is easily transformed in amine by water hydrolysis. This rearrangement has been widely discussed in the literature, and it has been proved that this is a nucleophilic rearrangement which occurs with C−C bond cleavage and displacement from an electron rich group to a center of electron deficiency (Scheme 11).125

Scheme 8. Amination of Alkylhalide by Nucleophile Substitution

Unfortunately, this reaction goes on after the amination. Indeed, the synthesized amine could also react on the alkyl halide and give secondary, tertiary, or quaternary amines, as for the reaction on alcohol. This last side-reaction is a significant limitation for the use of such method for the synthesis of primary amines. For instance, polyamines like the ethyleneamine family (ethylenediamine (EDA), diethylenetriamine (DETA), triethylenetriamine (TETA), etc.) are prepared from ethylene dichloride. This process consists of the reaction of ethylene dichloride with ammonia, followed by neutralization with sodium hydroxide to produce a mixture of ethyleneamines and sodium chloride. The salt is removed from the amine mixture, and the different amines are separated by fractional distillation. Additionally, TETA, in the presence of hydrogen, ammonia, and a catalyst (like Raney cobalt catalyst) can be used to prepare 1-(2-aminoethyl)piperazine.110 2.2.3.2. Nucleophile Substitution and Reduction with Azide. This method is performed in two steps: a substitution with sodium azide (NaN3) followed by a reduction of the azide function. The first step (alkyl halide substitution) was studied by Malik et al.111 and was performed in a DMSO (dimethyl sulfoxide)/water mixture at 50 °C with an excess of NaN3. Walkowiak-Kulikowska et al.112 also used this methodology to functionalize a fluorinated copolymer by copper azide/alkyne click chemistry. The second step consisted of the azide reduction with triphenylphosphine and an ammonia solution in N-methylpyrrolidone (Scheme 9).113,114

Scheme 11. General Reaction of the Curtius/Hofmann Rearrangement

The limitations are the same as beforehand for the use of isocyanates as precursors for amine synthesis (cf. section 2.2.1.1). However, to avoid any side-reactions, t-butanol or ditert-butyl dicarbonate (Boc2O) are added in the process to block the amine at the end of the rearrangement with a Boc group, which is easily released in acidic conditions.126,127 2.2.4. From Aldehyde and Ketone: Addition/Reduction Reactions. 2.2.4.1. Direct Reaction with Ammonia or Amines. Aldehydes and ketones can react with amino hydrogen from ammonia or primary amines. It is easier to obtain secondary or tertiary amines, but it is more difficult to obtain primary amines with good yield. The methods using molecular hydrogen as reductive agent,40 with a catalyst,128,129 are very

Scheme 9. Amine Synthesis from Azoture

Such RN3 compounds have to respect a molar ratio N3/CH lower than six to be stable and nonexplosives.115,116 This rule represents a huge limitation for the synthesis of amines with a high functionality. Indeed, the synthesis of small polyfunctional amines becomes unreachable with this methodology. 2.2.3.3. Gabriel Reaction. The Gabriel reaction allows obtaining primary amines from alkyl halide by nucleophile substitution with phthalimide (NH) in the presence of basic Scheme 10. Amination after a Gabriel Reaction

14186

DOI: 10.1021/acs.chemrev.6b00486 Chem. Rev. 2016, 116, 14181−14224

Chemical Reviews

Review

effective. However, this technic could not be used with some functional groups such as double bonds or aromatic rings. In these cases, metallic hydrides are used as reductive agents. This methodology traditionally uses a protective group to avoid an overalkylation (Scheme 12).130−132 Indeed, in the reaction

Scheme 14. Leuckart Reaction on Ketones Using Ammonium Formate

Scheme 12. Aldehyde Reductive Amination and OverAlkylation Undesired Reaction

lots of applications (coatings, varnishes, or composites, plus its use for the synthesis of the corresponding diisocyanate) it has been a recurrent subject. The first synthesis was performed by amino reduction of isophorone nitrile (IPN). Even if the conditions have been constantly improved, the hydrogenation is still carried out under severe temperatures and pressures (typically 120−150 °C and 150−270 bar) leading to a lot of byproducts: aminoalcohol, azabicyclic compound, secondary amine, etc.142,143 More recently, BASF patented a new two-step process which minimized the side-production of aminoalcohol (Scheme 15).144

between an aldehyde and an amine, an imine is formed which is then reduced into the desired amine (often sodium cyanoborohydride, NaCNBH3133). However, if the amine is not protected (if R2 = H in the below scheme), the synthesized amine could react again with the aldehyde and yield a secondary amine.40 Miriyala et al.134 proposed a reductive amination protocol from ketones, ammonia, and titanium isopropoxide-NaBH4 as reductive agent to obtain exclusively the primary amine without protection step. Unfortunately this methodology cannot be used on aldehydes. Dangerfield et al. 135 reported the preparation of primary amines from methylated glycosides thank to the Vasella’s amination without protection step. Then the same team136 synthesized primary amines, still without protection step, from aldehydes and metal hydrides as reductive agent. 2.2.4.2. Addition/Reduction Reactions: Beckmann Rearrangement (Hydroxylamine) and Hydrazine. The reaction of hydroxylamine (NH2OH) with aldehyde or ketone produces an oxime (Scheme 13). It leads, under acidic condition, to an

Scheme 15. BASF Process for Synthesizing IPDA

The first step consisted in synthesizing the intermediate imine with acid catalyst and the second one in hydrogenating the latter, without intermediate isolation. The announced yields were 95−97% but the conditions remained very severe (120 °C and 250 bar). However, for these reasons, another process has been proposed. It involves two consecutive reactions through the formation of an azine during a first step, which follows the same scheme as the addition/reduction of ethanol amine (Scheme 16). It has to be noted that azine formation can be considered as a new route to primary amines from ketones.145 Here, the problem is the use of nitrile derivatives which are generally highly toxic and hazardous for human being. 2.2.5. Short Summary on Amine Synthesis. Figure 1 summarizes all pathways to create primary or secondary amines.

Scheme 13. Oxime Preparation from Aldehydes or Ketones

2.3. Use of Amines for the Synthesis of Polymers

oxime rearrangement. This one, also called Beckmann rearrangement, allows amides to be created from ketone or aldehyde and thus amine through the hydrolysis or reduction of the amide.137 For instance Ayedi et al.138 used a system zinc/ HCl to hydrolyze the oxime intermediate. Jones et al.139 studied the mechanism and kinetics. Recently, Yamabe et al.140 discussed more precisely the mechanism (concerted or stepwised) and discovered that three mechanisms were possible for the Beckmann rearrangement. Indeed, they performed calculations on three different oxime substrates. The same principle is used for the Leuckart reaction which uses ammonium formate in the presence of ketone (Scheme 14).141 A formamide intermediate is formed which is reduced in acidic conditions. The inconvenience of this route is the temperature which is high, between 120 and 170 °C, preventing any use of thermosensitive substrates. An interesting example, which involves different amination reactions, concerns the synthesis of isophorone diamine (IPDA). Due to the importance of IPDA which is used in

The use of amines for the synthesis of polymers is generally ruled by the nucleophilic characteristics of these compounds. Thus, all the possible reactions of amines with various functions could lead to polymers/materials. Here a summary of all the different ways for preparing polymers from amines, either from amine or modified amine (maleimides, benzoxazines, ...), is reported. 2.3.1. Reaction of Amines with Acids: Polyamides and Polyimides. 2.3.1.1. Polyamides. Two methods allow the synthesis of polyamides by a step-growth polymerization mechanism including the reactions below (Scheme 17). The first method is the oldest one and goes through the preparation of an ammonium salt, formed by the acido-basic equilibrium between a carboxylic acid and an amine. Dehydration is performed at high temperature (generally higher than 180 °C).146 Phosphoric acid (H3PO4), phosphorous acid (H3PO3), or hypophosphorus acid (H3PO2) can be used as catalysts. There is some recent publications about the use of other catalysts, such as sulfated tungstate.147 The 14187

DOI: 10.1021/acs.chemrev.6b00486 Chem. Rev. 2016, 116, 14181−14224

Chemical Reviews

Review

Scheme 16. Synthesis of IPDA from Isophorone (with Azine Intermediary)

Figure 1. Summary of pathways to synthesize amine function from other common functions.

Scheme 17. Two Main Ways for Polyamide Synthesis

Scheme 18. Synthesis of Polyamide from Lactam Derivative

Scheme 19. Ring Opening Reaction of an Aromatic Dioxazoline by an Aminea advantage of this methodology is that the initial preparation of the salt allows respecting the stoichiometry. The second method which uses an aminolysis reaction was proposed by Jouffret et al.148 for the synthesis of polyamide 6,6 from hexane-1,6-diamine and dimethyl adipate as monomers in the presence of resorcinol as catalyst of the first aminolysis step. More recently Han et al.149 proposed several new efficient catalysts for the synthesis of amides from versatile esters. Thus, the association of zirconate V with hydroxypyridine or 4methyl-2-hydroxyquinoleine allows doing the reaction at a lower temperature (80 °C). Phenols (including resorcinol) are used as catalysts for the first step. Another reaction for the synthesis of polyamide is the ringopening polymerization (ROP) of lactams, mainly caprolactam and lauryl lactam (Scheme 18). Moreover, the nucleophilic reaction of aromatic amines with oxazolines at high temperature in the presence of catalyst leads to particular amides called poly(amino-amides). This reaction is often used for the chain-extension of thermoplastic polyamides (Scheme 19).150

a

Reprinted with permission from ref 150. Copyright 2002 Elsevier.

2.3.1.2. Polyimides: Reactions with Anhydrides of Diacids. The reaction between anhydride and/or naphthalic compounds and amines leads to the corresponding amido-acid. This particular function can be ring-closed to give an imide group (Scheme 20).151,152 The reaction of a diamine with a dianhydride leads to a polyimide. The reaction can be very fast, even at low temperature without catalyst. There is two ways for the cyclization, a chemical one and a thermic one. The first one uses chemical catalyst such as the system acetic anhydride and trimethylamine which the most common one.153 Nevertheless, the second one is the easiest way for preparing polyimides even if it needs high temperatures (250−400 °C) 14188

DOI: 10.1021/acs.chemrev.6b00486 Chem. Rev. 2016, 116, 14181−14224

Chemical Reviews

Review

The maleimide double bond (CC) can also undergo a Diels−Alder (DA) reaction which is widely used in thermally responsive systems because it is possible to do the opposite retro-DA (rDA) reaction and control it with temperature (Scheme 21). This reactivity is well-known and some studies have been done with various dienes, even if the most used dienes are furan derivatives.178−180 Always due to its high electron-acceptor character, the double bond of the maleimide function could be involved in radical addition reactions, including for polymerization. Indeed, BMI could thus copolymerize with other monomers. Some works have been reported on the system vinyl ether/maleimide, which is the most famous one because vinyl ethers are very good electro-donating monomers. This type of mechanism could be initiated thermally or by UV irradiation quite easily, which makes this system really interesting for industry.181−183 However, the most interesting reactivity of maleimide for this review remains the Michael addition and especially with amines. 2.3.2.2. Poly(aspartamide): Michael Addition. It is difficult to add a nucleophile, such as an amine, on a double bound because these are both nucleophile functions. However, if one or two electrophilic functions are in α position from the double bond, the addition becomes possible. Michael has described conjugated addition of stabilized carbanions on α, βunsaturated compounds. This nucleophile addition reaction (AN) became one of the most convenient method for the creation of carbon−carbon bonds. This reaction is performed under smooth conditions and could be used with various common nucleophiles such as thiols and principally amines to create respectively sulfur−carbon and nitrogen−carbon bonds (Scheme 22).184,185 Thiols are generally more nucleophilic than

Scheme 20. Synthesis of Imide from an Amine and a Cyclic Anhydride Such as Phtaleic, Maleic, or Naphtaleic Anhydride

for the cyclization step because it is done in bulk to avoid purification steps. The reaction of two moles of maleic anhydride with one mole of a diamine leads to bismaleimides (BMI). The reaction is carried out in chloroform, acetone, or toluene as solvent. The dehydration is the limiting step and is often performed with the use of catalysts.154,155 The catalysis nature allows controlling the proportion of imide and isoimide during the reaction. The intramolecular reaction is easier with aromatic diamines than with aliphatic ones. This method consists of transforming the acidic function of maleamic compound which is easy after the amide attack.156−164 The creation of such maleimide allows various thermosetting materials and polyimides to be obtained. 2.3.2. Polymaleimides and Michael Addition. The reactivity of maleimide functions has to be discussed. Indeed, even if it is not really an amine, it is derived from amine. It is thus significant to resume the possible reactions of maleimides. Moreover, in a second part, maleimide will be discussed as electron acceptor because it could react with amine in the Michael addition mechanism. Maleimides can be polymerized owing to four different routes: by radical or anionic homopolymerization, by Diels− Alder with polydiene, by radical copolymerization, and finally by Michael addition with amines.165 2.3.2.1. Maleimide Polymerization. Linear or cross-linked polymers derived from BMIs exhibit excellent thermal and mechanical properties. BMI can self-polymerize by radical or anionic chain-growth polymerization through their reactive maleic double bonds to give highly cross-linked materials. The anionic pathway is mostly considered as a parasite reaction and only few studies were conducted on the subject.166−171 In the same way their high electro-accepting character could also be used in thermic172 or UV27,172−177 radical mechanisms.

Scheme 22. Michael Addition with Amine

amines, leading to a higher reactivity, although bases are often used to deprotonate them due to their comparatively higher acidity.186 Moreover, amines could also be protonated to protect them and favor the thiol reaction. This method was especially used to achieve graft polymers from MMA and 2aminoethanethiol by Boyer et al.187 The advantages of such a reaction are a high conversion rate (close to 100%) and the rare

Scheme 21. Pathways to Polymerize Maleimide Derivatives

14189

DOI: 10.1021/acs.chemrev.6b00486 Chem. Rev. 2016, 116, 14181−14224

Chemical Reviews

Review

Figure 2. Ranking of some Michael acceptors in function of their reactivity with a nucleophile.

Scheme 23. Polyaspartimide Synthesis via Michael Addition between a Maleimide and Diaminea

a

A similar reaction could be done with dithiols.

polymerization reaction. One epoxide ring reacts with each amino proton of a primary amine (Scheme 24).

formation of byproducts. This description allows it to be classified among click chemistry reactions owing to the definition of Kolb and Sharpless.116 An equilibrium between thiol and tertiary amine was also highlighted by Boyer et al.,188 leading to a thiolate. In Michael addition, a side-reaction could lead to the addition of a carbanion intermediate directly on another Michael acceptor. This side-reaction has been used for the synthesis of linear polymers by anionic step-growth polymerization of methacrylates and maleimides189 compounds. This polymerization could also lead to branched polymers and networks.186 The amine addition could be performed on different acceptor groups.189−193 The reaction kinetics of the nucleophilic addition by ring opening or by Michael reaction have been studied by Billiet et al.194 and Gokmen et al.195 The conclusions of these works showed that maleimides compounds are very effective for fast Michael addition reactions (Figure 2); therefore, they are used in polymer/material applications. Moreover, Michael addition of amines to itaconic acid was also recently renewed by Ali et al.196 such as an important system to form pyrrolidone ring. The use of an amine coreactant allows polymerizing bismaleimides (BMI) through a Michael addition (Scheme 23).197 BMI step growth polymerization with aromatic diamines is generally carried out in m-cresol, and the polymers obtained can reach high molar masses. Crivello et al.198 reported a detailed investigation of the condensation of aromatic amines and maleimide. White et al.164 prepared and compared the reactions of secondary diamines with BMIs via Michael step growth polymerization. Sometimes, in the case of aliphatic amines a prevailing side-reaction could happen. Indeed, the amines can react with the maleimide ring and open it to give the corresponding polyamide. This reactivity has already been used at industrial scale for the preparation of polymers, such as Kerimid.199 Also polyaspartic products from Bayer are prepared by addition of aliphatic long chain diamine and fumarate (or maleate) esters to obtain secondary diamines which can further react with isocyanates. 2.3.3. Polyepoxide Polymers. Reaction of amino hydrogen with epoxide is a typical example of a step-growth

Scheme 24. Reactions between Primary and Secondary Amines and Epoxide Groups

If a monoamine reacts with a diepoxide, a linear polymer is formed. But usually di- and polyamines are used as curing agents/hardeners to build up polyepoxide networks. The reactivity of the amine increases with its nucleophilic character: aliphatic > cycloaliphatic > aromatic.200 While for aliphatic and cycloaliphatic amines the first and the second amine’s hydrogen exhibit similar reactivities, for aromatic amines the reactivity of the second hydrogen is 2 to 5 times lower than the reactivity of the first hydrogen. This means that, once the primary amine reacted, the created secondary amine exhibits a lower reactivity. Hydroxyl groups catalyze the reaction through the formation of a trimolecular complex that favors the nucleophilic attack of the amino group. Apart from species containing OH groups that may be added as catalysts, the epoxide−amine reaction generates OH groups. Therefore, the reaction is self-catalyzed by reaction products. Aliphatic amines which rapidly react with epoxide monomers are representative room-temperature curing agents for epoxide formulations for adhesives or coating applications. Aromatic amines are used as hardeners for epoxide-based matrices for composites. Sometimes, polyamide-amine oligomers (also called amidoamines), formed by the condensation reaction between diacids and polyamines, contain reactive primary and secondary amines in their structure that are used to react with epoxide monomers at room temperature (or even below) with moderate heat generation and a long pot life. They are used in primers or tie coats for steel and concrete. Very often for the synthesis of these polyamide-amines, the aliphatic amine is reacted with a 14190

DOI: 10.1021/acs.chemrev.6b00486 Chem. Rev. 2016, 116, 14181−14224

Chemical Reviews

Review

tall oil fatty acid which increases the hydrophobicity of the final network. 2.3.4. Polyureas and Polyhydroxyurethanes. 2.3.4.1. Polyureas. Isocyanate/amine reactions are very suited for polymerization since reactions are very fast and very exothermic, do not give any byproducts, reach high degree of conversion, and involve generally liquids: both diisocyanate and diamines or a mixture of diamines (Scheme 25). Depending on

reactive amines are linear aliphatic amines and linear aliphatic amines bearing N or O atoms in the β position, followed by cycloaliphatic amines, aromatic amines, and amines bearing a methyl group in the α position. Despite a higher reactivity, linear primary amines, which exhibit the higher nucleophilicity, allow complete reaction at room temperature in only 4 days.215 Therefore, the use of a catalyst is generally required for promoting the nucleophilic attack of the amine on the carbonyl group of the carbonate.216−218 In this context, the authors have decided to reinvestigate the long-known catalyzed aminolysis of carbonates and some of them reported complete reaction at room temperature with triazabiscyclodecene (TBD) catalyst.218 A lot of side reactions could explain the low molar masses of the formed PHU. Clements et al.219 described different intra- or intermolecular reactions likely to occur with a hydroxyurethane function in the absence of a catalyst. The formation of cyclic compounds (i.e., oxazolidinones or urea derivatives) depends on the reaction temperature, as well as elimination reactions (meaning formation of alcohols and water), which allow the recovery of an oxazolidinone and a diol. Another route for NIPU synthesis is the transurethane step growth polymerization. First a diurethane monomer is easily synthesized by reaction between a diamine (preferably an aliphatic one) with dimethylcarbonate using sodium methoxide as a base or through a direct reaction of a diamine with ethylene carbonate without any catalyst (Scheme 27). Then, the synthesis of the polymer is carried out in a melt transurethane process in which equimolar amounts of the diurethane monomer are reacted in bulk with a diol in the presence of Ti-catalyst at high temperature. As for the synthesis of polyesters or polyamides, the melt transurethane reaction is an equilibrium process, and therefore, the efficient and continuous removal of the byproduct (methanol or ethane diol) is crucial to drive the equilibrium to the right (Scheme 28).220,221 Granules of linear thermoplastic polyurethanes (TPU) can be prepared through this process, mostly aliphatic TPU due to the higher reactivity of aliphatic diamines compared to aromatic ones during the first step. Some side reactions have also been described.222 2.3.5. Phenalkamide and Benzoxazines: Synthesis and Polymerization. Phenalkamide or Mannich bases are prepared by reaction of an amine with a phenol and formaldehyde (Scheme 29).223,224 Usually, Mannich type hardeners for epoxide also include residual free phenol and diamine.225 They are designed for low temperature reaction, high humidity applications. Although the condensation decreases the functionality of the amine, the presence of the phenolic hydroxyl functionality on the aromatic ring produces a substantial accelerating effect on the epoxide/ amine reaction. Mannich bases rapidly react at low temperatures (0−30 °C). They have better compatibility with epoxide monomers than unmodified alkylene amines and are more resistant to blush and water spotting. These three molecules, one amine plus one phenol and two formaldehydes (instead of one for the Mannich reaction), are also used to synthesize benzoxazines (Scheme 30).226,227 Benzoxazines are thermally polymerized by ring opening polymerization (ROP) through an assumed cationic mechanism. By the use of dibenzoxazine monomers, cross-linked polymers are formed with near-zero shrinkage and high stiffness, excellent thermal properties, lower moisture absorption, and better flammability resistance (high char yield) than

Scheme 25. Formation of Polyurea Polymers

the functionality of the components, linear or cross-linked polymers are formed. However, linear polyureas or polyurethanes-ureas are insoluble polymers; therefore, the only processing route consists of cast molding or RIM (reaction injection molding). The main difference between cast molding and RIM is the reactivity. For cast molding, the pot-life is usually higher than 5 min, whereas for RIM it has to be lower than 1 min. As presented previously (cf. section 2.1.1), aliphatic amines react faster than aromatic amines. Hence, for cast molding, only hindered aromatic amines are used in order to control the reaction. Due to the high reactivity of isocyanate with amines, alcohols, and water, blocked isocyanates are often used for coating applications. In this case, reaction with amines is generally preferred for unblocking and curing reaction.201 2.3.4.2. Polyhydroxyurethanes: Non-Isocyanate Polyurethanes (NIPU). The chemistry of polyurethanes (PUs) has been extensively developed for several decades. Their properties can be modulated to use them in many fields of applications and numerous researchers paid a lot of attention to these versatile compounds.202,203 Typically, PUs are obtained by the reaction of diisocyanate or polyisocyanate with compounds having at least two reactive hydrogen atoms such as hydroxyl terminated oligomer (polyol). As diisocyanates are harmful reactants for human health, the old reaction of the synthesis of nonisocyanate PUs has recently gained an increasing interest in both the academic community and the chemical industry.204,205 Step growth polymerization of dicyclocarbonates and diamines can be an alternative route for the synthesis of PUs. This reaction, already studied in the past by Whelan et al.206 and Mikheev et al.,207 avoids the use of isocyanates and diamines and permits the formation of poly(hydroxyurethane)s (PHUs; Scheme 26). This method was tremendously studied and has recently attracted much attention, particularly by Endo’s team204,208−211 and Leykin et al.212,213 However, the aminolysis of carbonate presents a lower reactivity compared to isocyanates.201,214 Authors such as Diakoumakov and Kotzev have already studied the reactivity of amines owing to their structures. They reported that the more Scheme 26. Synthesis of Polyhydroxyurethane from Diamine and Dicyclocarbonate

14191

DOI: 10.1021/acs.chemrev.6b00486 Chem. Rev. 2016, 116, 14181−14224

Chemical Reviews

Review

Scheme 27. Synthesis of Diurethane Monomer

Scheme 28. Polymerization via Transurethane Reaction

Scheme 29. Synthesis of a Mannich Base

Scheme 31. Synthesis of Polyimines or Polyhydrazones from Dialdehyde and Diamine

polyepoxides. Commercial benzoxazine monomers based on bisphenol A or F, thiophenol, dicyclopentadiene are used in composites (prepregs or infusion processes) and adhesives applications and are ranked between polyepoxide and BMI in both performance and cost. 2.3.6. Miscellaneous Reactions. 2.3.6.1. Reaction with Aldehydes and Ketones. Polyimines and polyhydrazones are prepared by action of a dialdehyde on diamine or dihydrazides (Scheme 31). Those polymers are not thermally stable, but they present the advantage to undergo reversible reaction with other dialdehydes as showed by Lehn and Skene.228 Ketimine, which is attracting attention as a curing agent for high solid paints, is formed by the reaction between an aliphatic polyamine and a ketone such as methyl ethyl ketone or isobutyl ketone.229 Ketimine reacts very slowly when mixed with epoxide monomer, but it works as a kind of latent curing agent for coating applications by absorbing moisture in the air and regenerating amines to cure at room temperature. 2.3.6.2. Reaction with Aceto−Acetic Esters. Amines could easily react at room temperature with aceto−acetic esters derivatives to give particular vinylogous urethane functions. This reaction is often realized to cross-link polyaceto-acetic ester polymers. Amines react on the enol form of the aceto− acetic esters.230 This reaction is industrially used for coating or adhesive applications (Scheme 32).231 2.3.6.3. Reaction with Silanes. Silanes are highly reactive with nucleophiles and were thus used with amines to create silazanes functions (Scheme 33).232 The reaction between multifunctional monomers could lead to polysilazanes. Actually, the reactivity of amines in this mechanism involves only one proton as the reactivity with epoxides. Therefore, a monoamine could be used to form a polymer. The presence of catalysts (metal complexes) is necessary for this type of polymerization and a release of dihydrogen is observed. For instance Bellini et al.233 used a complex of barium and obtained between 2% and

99% of conversion for the reaction of secondary amines and hydrosilanes. 2.3.7. Short Summary Concerning the Use of Amines for Making Polymers. All the reactivities of amine allow large possibilities for the synthesis of polymer and materials with various properties (Figure 3). This shows the importance of amines in polymer chemistry and proves that it is compulsory, for the future, to synthesize specific biobased amines and use them to replace petro-based ones. Already some biobased amines are commercially available. The following paragraphs will aim, in a first part, at highlight and explain plenty of ways for obtaining or synthesizing amines from biobased compound. In a second part, their uses as precursors in material chemistry, if already reported, will be depicted.

3. SYNTHESIS OF BIO-BASED AMINES Two types of biobased amines could be highlighted; the first one is constituted from bioresources containing amines such as polylysine and amino acids or having functions with nitrogen as chitosan and the second one which is synthesized from modified biobased compounds such as fatty oils, sugar derivatives, terpenes, etc. 3.1. From Chitosan

Chitosan is obtained from the deacetylation of chitin. Namely, chitin is the most abundant natural polymer in the world after cellulose. Indeed, it is present in the shells of crustaceans, insects, and fungi walls. The global availability of chitin is estimated at 109 tons.234 The resulting polymer is called chitosan when partial deacetylation of chitin is greater than 60% (Scheme 34).235

Scheme 30. Synthesis of a Benzoxazine Monomer and Its Polymerization226,227

14192

DOI: 10.1021/acs.chemrev.6b00486 Chem. Rev. 2016, 116, 14181−14224

Chemical Reviews

Review

Scheme 32. Reaction between an Aceto−Acetic Ester Function and an Amine

Scheme 33. Reaction between Amines and Silanes to Give Silazanes

Scheme 34. Chitosan’s Structure

Deacetylation is carried out in basic medium, in the presence of a reducing agent, in an inert atmosphere and at a temperature of about 110 °C. The deacetylation content can be varied, so has to influence the properties of the chitosan (solubility, biodegradability, biological properties, ...). This biopolymer is cheap and has interesting properties: it is biocompatible,236 biodegradable, bioresorbable,237 nontoxic,238 and antibacterial.239 Chitosan is insoluble in common organic solvents and in aqueous solutions at neutral and basic pH. However, it is soluble in acidic solutions (such as acetic acid solution 1% mol); the amine groups are protonated and the presence of positive charges lead to ionic repelling and promote the solubilization.235 Chitosan is a weak acid; its pKa is 6.5. It has several functions on its backbone: primary alcohol, secondary alcohol, amine, and amide if the deacetylation of chitin is not complete. Despite great interest of natural amine functions, the high molar mass of chitosan makes it difficult to use directly as

a reactive amine. Therefore, it is necessary to carry out chemical, physical, or enzymatic modifications to increase its solubility/miscibility and thus increase its interest. Several methods have already been reported. For instance, chitosan can be depolymerized to reduce its molar mass and improve its solubility.240 Acid-degradation methods (with strong acids such as nitric or hydrochloric acid) are very common and fast ones but are not specific. Indeed, the hydrolysis proceeds randomly, generating a large amount of monomers and oligomers with various length. Under mild conditions, Illy et al.241 performed this depolymerization by using H2O2 in an acetic acid medium under microwave irradiation. Soluble oligomers in aqueous solution were obtained, with a degree of polymerization lower than 10. Another method consists of mixing the chitosan with

Figure 3. Summary of amine polymerization reactions. 14193

DOI: 10.1021/acs.chemrev.6b00486 Chem. Rev. 2016, 116, 14181−14224

Chemical Reviews

Review

3.2. From Amino Acids

hydroxybenzotriazole (HOBt) to solubilize it in water. HOBt complexes with the amine groups of chitosan and thereby forms an organic salt soluble in water (Figure 4).242

With chitosan, peptides and of course amino acids are among the most important molecules having amines groups on their backbones. They are used either in organic or macromolecular chemistry.5 Two cases will be distinguished here: the first one where amino acids are used without any chemical modification (i.e., conserving the amino acid’s structure as lysine) and the second one where amino acids are modified either by decarboxylation, to obtain mono- or diamine, or by deamination, to obtain other amino acids; this reaction has been done on lysine for instance or diacids, for example glutamic acids. 3.2.1. Why Amino Acids? The few available examples of renewable amines are limited to amino acids. They are a huge source of natural amines, after their decarboxylation. The biological chemistry (enzymes and bacteria) allows modifying biomass in several basic amino acids in a specific way, which enables planning strategies to obtain precise molecules. Conceptually this chemistry could be resumed as seen in Scheme 36.

Figure 4. Chitosan complexed with HOBt. Reprinted with permission from ref 242. Copyright 2006 John Wiley and Sons.

Tishchenko et al.243 made a comparison between hydrolysis, oxidative reaction, and microwave irradiation and concluded that the choice of the preferable chitosan fragmentation procedure depends on the further aims: enzymatic hydrolysis gives low polymerization degree (2−10), chemical degradation seems more preferable at large scale with molar masses around 5−15 kg mol−1, and microwave irradiation does not modify the deacetylation degree but is very time-consuming. To improve its solubility in organic solvents, chitosan can be reacted in dimethylformamide (DMF) with phthalic anhydrides which react with the amine groups with a degree of substitution of 98%.244−246 A well-known coupling using a carboxylic acid on the amine and alcohol functions of chitosan, in the presence of carbodiimide could be performed as well. Fangkangwanwong et al.242 conjugated carboxylic functions on amine and alcohol functions of soluble chitosan-HOBt system in water and thus obtained good coupling yields. These reactions took place in water, at room temperature. Kim et al.247 conjugated a catechol derivative, 3,4-dihydroxyhydrocinnamique acid, on the chitosan thanks to this method. This greatly increases the solubility of chitosan at pH 7 and provides adhesion property to the system. Mohy Eldin et al.248 chose to modify the chitin prior to deacetylation. They grafted benzoquinone onto the alcohol functions of chitosan, and then they turned it into amine functions with ethylenediamine prior deacetylation of the polymer (Scheme 35).

Scheme 36. Possible Deamination or Decarboxylation of Amino Acids

From Scheme 36, there are two ways to modify for amino acids-monomers: one where deamination is performed, which is out of the interest of this review, and another way where decarboxylation is performed to give a mono- or a diamine depending on the starting material. The variability of groups G makes the variety of the difunctional products usable in step growth polymerization. Medici et al.249−251 worked a lot on decarboxylation reaction via biochemical routes. Indeed, a great variety of amines could be obtained. However, the amine of interest for material applications should have at least two amine functions or one amine function and another reactive function to participate in a polymerization process. Some examples are given in Table 1. Cysteamine (1) is a very interesting molecule since its SH function allows grafting amine on versatile molecules,252 or polymers253,254 via a thiol−ene coupling reaction on double

Scheme 35. Amino-Chitosan’s Synthesis Starting from Chitina

a

Reprinted with permission from ref 248. Copyright 2012 John Wiley and Sons. 14194

DOI: 10.1021/acs.chemrev.6b00486 Chem. Rev. 2016, 116, 14181−14224

Chemical Reviews

Review

Table 1. Amino Acidsa and Their Corresponding Amines after Modification

a

All structures are idealized. Normally they are under zwitterion form, which means as dipolar ions (neutrals) bearing at the same time a positive and a negative charge.

C6 diamines that have been used for years in PA6-6 or PA6-10 syntheses. 3.2.2. Lysine as a Platform for Chemicals. The two most studied amino acids are glutamic acid (used almost exclusively as animal feed) and lysine. Indeed the annual production of glutamic acid and lysine are respectively of 3109 t/year and 1210 t/year. The most interesting is actually lysine because it bears already two amine functions on its backbone, which is, in most of the cases, a compulsory condition to create polymers. Therefore, lysine has been intensively studied in the past few years. Lysine production increases by about 4%/year. The low cost of lysine and its interesting functionality with two amines and one acid function, gives to lysine a great potential to become a platform chemical for polymer synthesis (Figure 5). Indeed, it could become a key building block for both commodity and high-value chemicals for the production of a wide range of value-added materials. Its conversion to 1,5-pentanediamine (also known as cadaverine260,261), ε-caprolactam,262,263 5aminovaleric acid,264 and lysinol6 has received recent attention in the literature. The main industrial manufacturers are Ajinomoto in Japan, ADM in the U.S., Evonik in Germany, and DSM in The

bonds or a Michael addition on activated double bonds. Thiol− ene coupling of cysteamine is one of the most famous routes to synthesize polyfunctional amines from renewable resources. Thus, a high versatility of amines could be reached.255−258 This coupling could also be performed with an amine salt (hydrochloride cysteamine) to avoid some side-reaction, such as the addition of the amine on activated double bonds; in this case the amine function is obtained after further neutralization with a base (Scheme 37). Scheme 37. Thiol-Ene Coupling between Cysteamine and a Double Bond

Serine (2) has been decarboxylated with Pseudomonas putida 512 to obtain ethanolamine.250 Aspartic acid (3) could also be decarboxylated with L-aspartate α decarboxylase to give βalanine. 259 Tyrosine (4) gives γ amino phenol after decarboxylation which could be used in polyepoxide chemistry. Ornithin (5) gives C4 diamines which compete with traditional 14195

DOI: 10.1021/acs.chemrev.6b00486 Chem. Rev. 2016, 116, 14181−14224

Chemical Reviews

Review

Figure 5. Different functionalizations of lysine for the synthesis of useful building blocks.

Scheme 38. Synthesis Way of L-Lysine from Cellulose Thanks to the Corynebacterium Glutamicum and from Lysine to Pentanediamine

leads to the α-aminocaprolactam which gives by deamination the corresponding classical ε-caprolactam.262,263 In 2005, Frost263 reported the conversion of lysine into α-amino-εcaprolactam (Scheme 40). The preferred solvent for the

Netherlands. Recently, Ajinomoto, the world’s largest lysine manufacturer, has used nonedible cellulose feedstock such as corn straw for lysine synthesis. The decarboxylation of lysine leads to 1,5-pentanediamine which is considered as the concurrent of the hexamethylene diamine (Scheme 38). The company Ajimoto is also the biggest manufacturer of 1,5-pentanediamine. The Cathay Industrial Biotech company proposes as well some polymers from this diamine. Historically, it has to be reminded that the lysine has been used to create lysine diisocyanate (LDI). Actually, SNPE Group did the esterification of the carboxylic acid to obtain the 2alkylester pentanediamine265 followed by a transformation in diisocyanate thanks to a classical method (Scheme 39).

Scheme 40. From L-Lysine to ε-Caprolactam through αAmino-ε-caprolactam

cyclization is 1,2-propanediol with very good yield (96%) in 2 h. Successive deamination was done at −5 °C, in the presence of KOH and hydroxylamine-O-sulfonic acid (HOSA) with the formation of N2 and K2SO4. Finally, ε-caprolactam was purified by sublimation in 75% yield. In a more recent patent262 the lysine and/or salts do not need to be purified from the fermentation broth prior to being heated to form the intermediate product, α-amino-ε-caprolactam. The Genomatica company develops by fermentation, besides caprolactam, hexamethylene diamine and adipic acid and is integrated to the producer of biobased basic products for the creation of polyamides.266 Then, the reduction of lysine by catalytic hydrogenation by Du Pont de Nemours could lead to lysinol. Lysine hydrogenation proceeds under relatively smooth conditions with Ru/ C catalyst in water (100−150 °C, 48−70 bar, pH 1.5−2) to give lysinol in good yield.6 Lysinol is an aliphatic diamino alcohol which possesses clear structural and functional similarity to ethylenamines and ethanolamines (Scheme 41). Finally, polyamines dendrimers could be obtained from Lysine. During the first step, Lysine-N-carboxyanhydride spontaneously polymerizes in water by step growth polymerization. The linear polylysine, called first generation (G1) is

Scheme 39. LDI Synthesis from Lysinea

a

Reprinted with permission from ref 265. Copyright 1993 John Wiley and Sons.

This very interesting product was proposed too early to meet success, since at that time the interest for biobased chemicals was not so important; therefore, the synthesis steps number, and therefore the high cost, led to the stop of its industrial production. The deamination of the lysine into 5-amino valeric acid (precursor of the nylon 5) was proposed by Pukin et al.264 The direct cyclization of the zwitterion of the amino acid, actually the intramolecular amidification in diluted conditions, 14196

DOI: 10.1021/acs.chemrev.6b00486 Chem. Rev. 2016, 116, 14181−14224

Chemical Reviews

Review

This work is preceded by a study of the synthesis of aminobutyric acid (AA) by decarboxylation of the glutamic acid by an enzyme supported on a calcium alginate.270 The glutamic acid represent a good building block for the creation of different amines or useful products such as N-vinyl2-pyrrolidone (NVP) or acrylonitrile, which is an attractive molecule for the synthesis of amines from alcohols.

Scheme 41. Synthesis Pathway for the Synthesis of Lysinol from L-Lysine

3.3. From Vegetable Oils and Derivatives

thus obtained. It is used as a macroinitiator for the step growth polymerization of N-ε-trifluoroacetyl-L-lysine-N-carboxyanhydride (Lys(Tfa)-NCA) to form generation 2 (G2) polymer which can be used as macroinitiator and so on until the formation of generation 5 dendrimers (Figure 6).267

Additionally to natural amines, this part and the following ones are dedicated to the syntheses of biobased amines by modifications of renewable raw materials or building blocks. This first part is focused on vegetable oils as cheap and renewable starting materials. Some efforts have been made to produce biobased polymers derived from vegetable oils. Derivatization reactions with nitrogen-containing reagents are good pathways to create a range of new products. These molecule types are generally composed from ester/acid and double bonds. The addition of an amine function on such structures could be done on these two functions. Some attempts to convert lipids or lipid derivatives into amines or polyamines have been reported. Biermann et al.271 reviewed different methods to functionalize unsaturated fatty acids. 3.3.1. Amino-Acids/Esters Synthesis. For more than 50 years, scientists have been working on the reaction involving unsaturations of vegetable oil in order to transform them into amines.272 McDonald and Gruger were the first to propose some methods to synthesize ω-amino acids from vegetable oils. They described two different methods: (1) purified oleic acid was converted into polyhalo-acids and then into polyamino acids by reaction with liquid ammonia and (2) the double bonds of fatty acids were reacted with tetranitromethane followed by a subsequent reduction reaction. Moreover, amino acids can be obtained by reaction of alkyl halide with ammonia. As an example, 11-amino undecylenic acid used by Arkema for the Rilsan 11 synthesis was made from ricinoleic acid extracted from castor oil. It was manufactured in five steps: (I) methanolysis of triglycerides in basic media, (II) pyrolysis to obtain the heptanal and the methyl undecylenate ester, (III) hydrolysis of the ester, (IV) the terminal alkene can react with acid bromide for the synthesis of the 11-bromo undecylenic acid (Scheme 43). The addition of Br at the chain end is very important for the next step and is only possible in the presence of peroxide (antiMarkovnikov addition273). This 11-bromo undecylenic acid is known to react with an aqueous, alcoholic, or aqueous-alcoholic ammonia solution (V) to obtain the corresponding aminocarboxylic acid, thanks to the Hoffman method for preparing amines.274 The reaction takes place at relatively low temperatures and speeds up with temperature. In contrast, the percentage of amino acid decreases when the temperature increases. Recently, Arkema proposed another method, carried out under mild conditions and greatly reducing the reaction time, in order to limit the side reactions which create impurities such as secondary amines.275 This new solution45 proposes the synthesis of the ω-amino acid thanks to a conjugated method for the production of amino alkanoic acids and polyol carbonate from a natural monounsaturated fatty acid. The ammonia released during the reaction between urea and polyol to give the carbonate was used in the last step of ω-amino acid synthesis.

Figure 6. Polylysine dendrimers of generations 1, 2, 3, and 4 (G1, G2, G3, and G4). Reprinted with permission from ref 267. Copyright 2010 John Wiley and Sons.

Scholl et al.268,269 did not used classical protonic acids as catalyst (sulfuric acid (H2SO4) and p-toluenesulfonic acid (pTSA)) but transition metal derivatives. Thus, they could work at lower temperature (150 °C). The most well-known organometallic compounds for this reaction are Ti(OR)4 and Zr(OR)4, because of their good resistance to high temperature, even if they are sensitive to hydrolysis. 3.2.3. Glutamic Acid. Glutamic acid is one of the two most studied amino acids with a higher production than lysine. It is also a significant constituent of waste streams from biofuel production (bioethanol) from corn or wheat.5 Many modification ways of the glutamic acid have been performed to reach amine derivatives; Scheme 42 summarizes them. Scheme 42. Platform of Molecule from Glutamic Acid

14197

DOI: 10.1021/acs.chemrev.6b00486 Chem. Rev. 2016, 116, 14181−14224

Chemical Reviews

Review

Scheme 43. Transformation of Castor Oil into ω-Amino Acid

3.3.2. Diamines. Because fatty nitriles are becoming of interest in the framework of biofuels and for the valorization of the oil part of biomass to form fine chemicals, the production of long-chain fatty nitriles by direct reaction of esters with ammonia has been the subject of numerous patents and academic publications. For example, several catalysts were tested for the ammonia dehydration of lauric acid methyl ester into nitrile in a gas-phase continuous downstream process, at high temperature, 300 °C. Catalysts tested were used with various acid−base features.45,285,286 And, as we are concerned by amines, the catalytic reduction of nitrile compounds could lead to amine function. An interesting case concerns sebacic acid or 10-undecenoic acid obtained from the castor oil methyl ester, which could be transformed into the corresponding 1,10decamethylenediamine (C10) after additional chemical reactions, as those described in section 2.2.1.4. Another example concerns dimers or trimers of fatty acid. Figure 7 shows typical diacid and triacid that can be obtained by a Diels−Alder reaction of 9,11-linoleic acid.

The manufacture of amines from vegetable oils is really an industrial stake. Amino acids can also be obtained from biotechnology. Using a completely new, two-phase fermentation technology, Evonik is now able to manufacture ωaminolauric acid (ALA).276,277 The production of amino lauric acid (amino acid in C11) has been produced for two years by Evonik from renewable raw material as palm kernel oil.278 A simple hydrolysis could lead to the corresponding amino acid. Also, from vernolic acid, Ayorinde et al.279 obtained the 12aminododecanoic acid and 11-aminoundecylenic acid via a reaction sequence (four steps) that includes the formation of 12-oxododecanoic acid oxime (Scheme 44). Scheme 44. Creation of the 12-Aminododecanoic Acid from Fatty Acids

Spiccia et al.280 reported also a multicatalytic sequence to obtain a valuable PA-11 precursor, the methyl-11-aminoundecanoate, which is prepared by a ruthenium-catalyzed cross-metathesis and a highly regioselective palladium-catalyzed amination-hydrogenation reaction from canola oil (purified rapeseed oil). Another way to develop amine with metal catalysts was reported by Kamiyama et al.281 They found that an oxidative dehydrogenation took place between fatty acids and ammonia in the presence of some metal salts such as [Cu(NH3)4]SO4 (higher activity) in aqueous media and gave α- or β-amino acids. Lastly, Schaffer et al.282 created an enzymatic method for the oxidation or amination of fatty acids and their esters using whole cell biocatalysts. Schrewe et al.283 achieved the direct and regioselective amino-functionalization of nonactivated carbon of fatty acids. By one recombinant microbial catalyst, coupling oxygenase and transaminase catalysis in vivo, both substrates were converted with absolute regioselectivity to the terminal amine via two sequential oxidation reactions followed by an amination step (three chemical reactions with a single catalyst). Recently, Song et al.284 proposed a biotransformation of longchain fatty acids (e.g., oleic acid, ricinoleic, and lesquerolic acid) into ω-aminocarboxylic acids (C11, C12, and C13) using recombinant Escherichia coli based biocatalysts.

Figure 7. Formulas of dimer and trimer of 9,11-linoleic fatty acid.

Even if these products are called dimers, there is a small percentage of trimers (and also residual amounts of the initial monomer) and consequently the functionality can vary between 2 and 2.1. Some grades are also hydrogenated to remove the double bonds. Only one producer, Croda287 has developed a new type of vegetable oil derived amine, based on its own biobased fatty acid dimer technology. Priamine dimer diamines (Figure 8A) are hydrophobic and apolar diamines, giving excellent barrier properties to coatings and yet provide excellent flexibility and impact resistance to materials. Another important advantage of the dimer diamine is its low viscosity allowing use in high solid contents and even solvent-free formulations. This new technology will give to formulators a great flexibility as these dimer amines can be used solely or as a cohardener and will bring true sustainable coatings one big step closer. Dimer diamine is proposed in four versions (1071, 1073, 1074, and 1075), the purity of the dimer increases with the 14198

DOI: 10.1021/acs.chemrev.6b00486 Chem. Rev. 2016, 116, 14181−14224

Chemical Reviews

Review

Several companies and industrial consortia have begun to develop industrial productions at large scale.297 An important application of biobased succinic acid is the production of 1,4butanediol (1,4-BDO), which is a starting material for tetrahydrofuran and γ-butyrolactone and also, of particular interest here, for 1,4-BDA. 3.4.1.2. Adipic Acid. Adipic acid is a C6 platform chemical that can be converted by chemical processes into 1,6hexanediol and adipodinitrile through amidation and dehydration. Adipodinitrile is the starting material for the production of 1,6-hexamethylene diamine (HMDA) as described in a patent of Voit et al. (Scheme 46).298 A partial hydrogenation of aliphatic α,ω-dinitriles leading to the corresponding ω-aminonitriles, more precisely the 6aminocapronitrile, could be performed in the presence of selective and efficient catalyst.299 6-Aminocapronitrile can be converted into ε-caprolactam after cyclization (Scheme 47).300 To synthesize adipic acid, one process propose the classical petrochemical route adapted to biobased benzene (Scheme 48) but still produces the powerful greenhouse gas N2O as a byproduct. There are considerable activities to produce adipic acid through fermentation from carbohydrates. Fermentative production of adipic acid from glucose was first published by Frost and Draths in the 90s,301 but its industrial production at a large scale was proposed later than the one of succinic acid. Several processes are in development.302 Bioengineered microorganisms that are capable of producing adipic acid from glucose were recently announced and proposed at pilot scale by two companies.266 Another route proposed by another company is the catalytic conversion of glucose to adipic acid via glucaric acid. In this case, the key parameter is the choice of the catalyst (Scheme 49).303 3.4.1.3. Azelaic Acid. Via azelaic acid, extracted from rapeseed oil, 1,9-nonandiamine, could be synthesized (Scheme 50).304 Usually this diamine is produced from fossil resources via butadiene. 3.4.2. From Bio-Based Polyols. Pera-Titus and Shi305 reported that only a little research has been done for the catalytic amination of biobased alcohols, and that ruthenium or iridium catalysts could be performant to do such reactions. Recently, Vogt et al.306 proposed a catalytic system able to selectively convert various bioalcohols and diols into primary amines using ammonia at high temperatures. Parameter optimization showed the importance of ammonia concentration and Ru/P ratio and the catalyst can be reused at least 6 times without significant loss of activity or selectivity.306 However, some biobased alcohols have been transformed into amine derivatives via different methods. Glucose, mannose,

Figure 8. Scheme of diamines derived from the dimer fatty acid one by direct amination (A) and the other one by amidification with a diamine (B).

reference number. New diamines could also be obtained more easily and at a lower price by amidification of the acidic function (Figure 8B). However, it is necessary to remove the huge excess of diamine used to achieve this reaction. 3.3.3. Miscellaneous Amines. Some polyamines, called here miscellaneous amines have been created using fatty backbones. For instance, the use of manganese III acetate mediated addition of sodium azide followed by the catalytic reduction of the azide group to give the corresponding amine was explored but not extended to triglycerides. Aziridines have been synthesized from unsaturated fatty acids in a two-step process, but this function does not behave as an amine group in cross-linking of polymer.288,289 Biswas et al.290 discovered another method to produce an oleate-aniline adduct using catalytic amount of ionic liquids. The use of nitrile compounds as precursors of amines after catalytic reduction has also been tested with fatty acids.45 Recently, Zhao et al.291 prepared secondary amines in a fivestep process starting from epoxidized triglycerides. The oxirane groups were transformed into diols which were brominated using triphenyl phosphine. At last, the displacement of the bromine groups with sodium azide and the subsequent catalytic reduction gave the expected polyamine.292,293 The addition of an excess of diamine onto an epoxidized vegetable oil under controlled experimental conditions can lead to an aminated fatty acid.292,293 Hence these applied methodologies imply multistep processes with low yields and formation of many byproducts. Furthermore, a simpler new solution was proposed by Stemmelen et al.,294 which involves the thiol−ene reaction to graft amine on unsatured triglycerides. They realized the addition of a cysteamine hydrochloride (CAHC) on grapeseed oil (GSO) under UV (Scheme 45). Thanks to thiol−ene reaction, Turunc et al.295 proposed also the synthesis of new fatty acid-derived amine functional monomers. 3.4. From Sugar Derivatives

3.4.1. From Bio-Based Diacid. 3.4.1.1. Succinic Acid. Diamine such as 1,4-butanediamine (1,4-BDA) could be obtained from succinic acid.296 Biobased succinic acid is obtained owing to different types of fermentation processes of carbohydrates, mainly glucose.

Scheme 45. Amination of GSO by UV Initiated Thiol−Ene Couplinga

a

Reprinted with permission from ref 294. Copyright 2011 John Wiley and Sons. 14199

DOI: 10.1021/acs.chemrev.6b00486 Chem. Rev. 2016, 116, 14181−14224

Chemical Reviews

Review

Scheme 46. From Adipic Acid to HMDA

(D, L, and D,L compounds).45 However, the yield after the hydrogenation step remains very low. Dehydrogenases and transaminases are engineered to convert the isosorbide into diamine in a two-step biocatalytic process, which includes cofactor recycling.314,315 3.4.2.5. From Hydroxymethylfurfural (HMF). Aminomethylhydroxymethylfuran derivatives (Scheme 54) are well-known for their widely recognized pharmaceutical activities. These structures are generally produced from either furfural alcohol or furfural.316 However, the reported procedures usually require harsh reaction conditions to convert pentosanes or pentoses into furfurylamine, with lower selectivity, in a two-step process, involving acid-induced dehydration to furfural followed by reductive amination.317 Cuklovic and Stevens reported a twostep reductive amination of HMF in the absence of catalyst (Scheme 54).318 This two-step procedure starts with imine formation, which limits the scope of amine substrates, followed by the use of excess NaBH4, which generates copious amounts of waste besides the use of a costly hydrogenation reagent. Xu et al.319 described the direct reductive amination of 5-HMF with various primary and secondary amines by dichloro-bis(2,9dimethyl-1,10-phenanthroline) and ruthenium(II) (Ru(DMP)2Cl2) as catalyst for hydrogenation. Similar reactions can be also performed between furfurylamine and formaldehyde, aldehyde or ketone,320,321 to obtain amine dimers. The reaction plan (Scheme 55) is proposed by Holfinger et al.320 This 2,5-bis(aminomethyl) furan (BAF) is a very interesting pseudoaromatic diamine which can be obtained through different ways.322,323 In one of these ways, formyl group of biobased 5-hydroxymethylfurfural (HMF) is reduced to yield 2,5-bis(hydroxymethyl)furan. Then the hydroxyl groups could be converted into primary amines (Scheme 56) owing to various chemical routes (not necessarily green chemistry).324 Another way consists of the catalytic oxidation of HMF into 2,5-diformylfuran (DFF) followed by reductive amination to yield 2,5-bis(aminomethyl)furan. This last step requires delicate conditions to prevent the formation of byproducts such as secondary, tertiary, and polymer amines due to the condensation and/or hydrogenation of the reactive dialdehyde groups and high nucleophilic amine groups. Recently, the direct reductive amination of DFF with ammonia into BAF was demonstrated, for the first time, over the commercial type Nickel-Raney and acid treated Nickel-Raney catalysts (Scheme 57).325 2,5-Bis(aminomethyl)tetrahydrofuran was also prepared from 2,5-furandicarboxylic acid (FDCA).296

Scheme 47. Synthesis of ε-Caprolactam from ω-Aminonitrile

2,3-butanediol (2,3-BDO), isosorbide or hydromethylfurfural have been for instance modified into amines. 3.4.2.1. Glucose. Pseudomonas putida S12 was engineered for the production of monoethanolamine (MEA) from glucose via the decarboxylation of the central metabolite L-serine, which is catalyzed by the enzyme L-serine decarboxylase (SDC).250 3.4.2.2. From D-Mannose. Matsuoka et al.307 synthesized a trisaccharidic glycomonomer from D-mannose via this method. The alcoholic function of D-mannose was converted into triflate, unstable intermediate which is immediately treated with potassium phthalimide to obtain the desired phthaloylic derivative. A final hydrolysis gives the amine. 3.4.2.3. From 2,3-Butanediol (2,3-BDO). Alternatively, HMDA and ε-caprolactam can also be produced from biobased butadiene (Scheme 51). One pathway to biobased butadiene is from catalytic efficient dehydration of biobased 2,3-BDO. The 2,3-BDO would be produced by fermenting carbon monoxide with LanzaTech’s proprietary microorganism to make ethanol and 2,3-BDO as coproducts, up to 50% of the volume (and the rest in ethanol).308 3.4.2.4. From Dianhydrohexitols. Dianhydrohexitols are heterocyclic compounds that are obtained by double dehydration of the corresponding hexitol. Among these diols, the 1,4:3,6-dianhydrohexitols are well-known under the names isosorbide, isoidide, and isomanide. Isosorbide is the only one produced at a large commercial scale. Evonik has patented a simple amination with ammonia to create aminated isosorbide.191,309 The reducing amination (Scheme 52, way A) was also studied by the team of Wiggins,310,311 but the yields were not given. Two other methods using the diazoture (cf. section 2.2.3.2) and Gabriel’s reaction (cf. section 2.2.3.3) allowed the best yields to be obtained. The reduction reaction of azotures was used on isoidide (Scheme 52, way B) in allowed good yield (88%) but the use of potentially very explosive diazoture prevents it from being developed at an industrial scale.116 The Gabriel reaction allowed to obtain the diaminoisoidide (cf. section 2.2.3.3). This method (Scheme 52, way C) has been proposed for the first time by Cope and Shen,312 and then the yields were improved to 47% by Thiyagarajan et al.313 via a four-step pathway. The catalytic hydrogenation of nitriles into amines was patented by Arkema (Scheme 53) on isosorbide and its isomers

3.5. From Terpenes

The early works of functionalization of terpenes with amine groups were reported by Keim et al.326,327 They synthesized mixtures of primary, secondary, and tertiary amines via terpene

Scheme 48. Dominant Petrochemical Route to Adipic Acid

14200

DOI: 10.1021/acs.chemrev.6b00486 Chem. Rev. 2016, 116, 14181−14224

Chemical Reviews

Review

Scheme 49. Chemical Catalytic Route for Adipic Acid from Glucose

asymmetric transfer hydrogenations of aromatic alkyl ketones332 or for enantioselective alkynyl zinc additions to aromatic and aliphatic aldehydes.333 Furthermore, interesting amines were synthesized such as 1,8-diamino-p-menthane or menthane diamine (MNDA). It is a primary alicyclic diamine in which both amino groups are attached to tertiary carbon atoms (Figure 9). MNDA is obtained from terpene through a complex process in the presence of water, hydrogen cyanide, and sulfuric acid at low temperature, typically 40−70 °C. The first synthesis of MNDA was claimed in a patent.334 More recently, primary amines were synthesized from citronellal by biphasic reductive amination.335 Finally, Alves et al.336 were the only ones to report the synthesis of amino terpene for further polymer synthesis. Amino groups were grafted by thiol−ene coupling of cysteamine on dihydromyrcenol.336,337 Synthesized amino terpenes were also used for the synthesis of surfactants by reaction with dextran.

Scheme 50. Synthesis Pathway from Azaleic Acid to 1,9Nonanediamine

Scheme 51. Synthesis of Adipodinitrile from Butadiene

amination, catalyzed with Pd acetylacetonate or tributyl phosphite. Most of the synthesized amines from terpenes relate to medical applications. Thus, Lochynski et al.328 and Gajcy et al.329 proposed the synthesis of amino terpenes in several steps for neurological applications. They presented the possibility of using terpene such as (+)-3-carene and (−)-menthol, as synthons for the preparation of amino acids with predetermined stereogenic centers. These compounds might be considered as structural analogues of γ-aminobutyric acid (GABA) and thus are expected to be of interest as potential inhibitors of GABA neuro-receptors. Amino derivatives from limonene were also investigated for their effect on egg hatchability and mortality.330 Moreover, two regioisomers of limonene β-amino alcohol derivatives were found to be significantly effective against in vitro cultures of the Leishmania (Viannia) braziliensis.331 Amino derivatives of terpene were also used as ligand in catalytic enantioselective reaction. Thus, limonene derived amino alcohol was used for the catalytic

3.6. From Cardanol

The cardanol, extracted from cashew nut shell liquid, represents a good alternative to create aromatic amine. Due to its aromatic ring, cardanol confers partial rigidity and good mechanical and thermal properties to materials. For instance, in order to reduce the toxicity problems due to the use of phenol, as well as high energy consumption due to the use of solution in water,338 and in order to use renewable resources, the Mannich reaction could be performed on cardanol with paraformaldehyde (a formaldehyde oligomer). Thiyagarajan et al.339 conducted this reaction with hexamethylenediamine (HMDA) or diethylene-

Scheme 52. Synthesis of Unsubstituted Diaminoisoidide by Three Different Ways

14201

DOI: 10.1021/acs.chemrev.6b00486 Chem. Rev. 2016, 116, 14181−14224

Chemical Reviews

Review

Scheme 53. Double Amination of Isosorbide

Scheme 54. Synthesis Routes for Preparing Aminomethyl-Hydroxymethylfuran

Scheme 55. Mechanism of the Reaction between Furfurylamine and Aldehydea

a Reproduced from Holfinger, M. S.; Conner, A. H.; Hill, C. G., Jr. Ind. Eng. Chem. Res. 1997, 36, 605−613. Copyright 1997 American Chemical Society.

obtained according to the mechanism in Scheme 58. The final product is a mixture of two phenalkamines (A) and (B) (and sometimes the starting diamine in excess) since both amino groups of the HMDA can participate in the reaction. Generally only the product (B) is obtained or mentioned in the literature. Zhang et al.340 used melamine as reactant for the creation of the corresponding phenalkamine. However, they functionalized the corresponding amine into alcohol for the synthesis of polyurethanes. Some other diamines have been used, such as DETA, TETA, or xylylene diamine derivatives. As presented in section 2.3.5, the reaction of two equivalents of paraformaldehyde with one phenolic function and a primary amine yields to a benzoxazine. Already some various benzoxazines have been synthesized from phenols (Figure 10).340−349 Particularly, Lochab et al.343 synthesized benzoxazines with aniline as amine precursor in bulk and with quantitative yield. Sini Nalakathu Kolanadiyil et al.344 synthesized a particular series of itaconimide/nadimide-functionalized benzoxazine and used cardanol to vary the corresponding characteristic of the curable monomer.

Scheme 56. Synthesis and Amination of 2,5Bis(hydroxymethyl)furan

Scheme 57. Synthesis and Reductive Amination of 2,5Diformylfuran (DFF)

Figure 9. Formula of menthane diamine.

triamine to confer good properties of flexibility and water resistance to materials. The targeted phenalkamine was 14202

DOI: 10.1021/acs.chemrev.6b00486 Chem. Rev. 2016, 116, 14181−14224

Chemical Reviews

Review

Scheme 58. Synthesis of a Phenalkamine from Cardanol and Paraformaldehyde

Additionally, another way for the synthesis of cardanol-based amines consists of nitration followed by reduction. Nitration of hydrogenated cardanol with nitric acid leads to a mixture of mono-, di-, or trisubstituted nitro isomers, with respectively one, two, or three equivalents of acid.352 A simple reduction of these nitro groups gives the corresponding mono, di or triamine. Sadavarte et al.353 first removed the phenol function of the saturated cardanol then performed the nitration and the reduction to obtain an aromatic diamine (Scheme 60). It is important to carry out this nitration on saturated cardanol since double bonds reacts with strong acids, forming cardanol oligomers. Jadhav et al.354 created an aromatic diamine with a long spacer by reacting cardanol monoaminated with 4,4′-dichlorodiphenyl sulfone (Scheme 61).

Figure 10. General benzoxazines reported in the literature.

One of the main issues of this synthesis pathway is the use of formaldehyde, which is highly hazardous (CMR), and the nonfunctionalization of the aliphatic chain conducting to a limited amine. Moreover, the thiol−ene reaction has been performed by Darroman et al.350 with cysteamine hydrochloride on the double bonds of cardanol. In order to increase the amine functionality, allylation of the phenol was performed with allyl bromide, prior thiol−ene coupling. Thus, they obtained, after reaction, a polyamine with a functionality of 3.7 active hydrogens. Mhaske et al.351 proposed two routes for the synthesis of monoamines from cardanol for end-capping polyimides in order to study effect on solubility of the polymers. After saturation of the double bonds of cardanol, a first route consists of the amination reaction on the aromatic group with diazotized sulfanilic acid and sodium dithionite (I). In the second route, the amination is performed with 1-chloro-4-nitrobenzene or 1,2-dichloro-4-nitrobenzene in the presence of potassium carbonate under reflux, followed by a pallado-reduction of the nitro group (II) (Scheme 59).

3.7. From Lignin Derivatives

Lignin is the main biobased source of aromatic molecules, because its chemical structure is composed of phenylpropane units, originating from three different aromatic alcohols: pcoumaryl, coniferyl, and sinapyl alcohol.355 If it is interesting to study the synthesis of amines from lignin, it has to be stated that there is no direct work on the synthesis of amines from lignin but only from lignin oligomers or from lignin-derived monomers. Few teams worked on the synthesis of amines from direct Mannich reaction on the lignin (Scheme 62). In fact, the reaction of lignin with amine consists generally to add another function as alcohol (reaction with ethanol amine or diethanol amine) in order to use for the synthesis of polyurethanes-based lignin or an acid function (reaction with amino acids)356 for the creation of proton exchange resins or to create ammonium resins for chelation with other cationic compounds.357 Indeed, it is a hard work to depolymerize lignins into oligomers or small molecules. Some works have already been reported in the literature, and there is several different techniques for transforming lignin into small organic molecule, and more particularly phenols, but only two routes are usually performed.355 The first one is a thermal treatment which leads to a complex mixture of phenolic compounds. The second one is a thermo-oxidative treatment. In this last case the mixture is composed principally of vanillin and syringaldehyde. Hence, among the great variety of aromatic compounds obtained from different routes of lignin depolymerization, only vanillin and, more recently, syringaldehyde were extracted with good yield from wood lignin. Thus, the valorization of these two molecules become recently of great interest.358 Production of these molecules is already done at industrial scale. Therefore, functionalization routes of these compounds were developed to valorize them in the synthesis of biobased aromatic polymers. The rest of this chapter deals only with the vanillin which

Scheme 59. Synthesis of Monoamine from Cardanol with Diazotized Sulfanilic Acid/Sodium Dithionite (I) or by Reaction with Chloro-4-nitrobenzene Derivatives Followed by Pallado-Reduction (II)

14203

DOI: 10.1021/acs.chemrev.6b00486 Chem. Rev. 2016, 116, 14181−14224

Chemical Reviews

Review

Scheme 60. Synthesis of a Diamine Based Cardanol from Hydrogenated Cardanol

Scheme 61. Synthesis of a Diamine from Mono-Aminated Monomer by Reaction of 4,4′-Dichlorophenyl Sulfone

Scheme 62. Schiff Base Preparation from Lignin

Scheme 63. Synthesis of Vanillylamine from Vanillin

Du et al.362 also managed to isolate vanillylamine from the bioconversion of lignin degradation product (vanillin) with CV2025 ω-transaminase. At low concentration of vanillin (100 residues) due to unavoidable deletions

This method involves simple reagents and allows preparing high molar mass polymers in both good yield and large quantity with no detectable racemization at the chiral centers. The considerable variety of NCAs that have been synthesized (>200) allows exceptional diversity in the types of polypeptides that can be prepared for various application, such as drug delivery systems.402−406 4.2.2. Polyamide 6,6, Polyamide 6, and Substitutes. Just a reminder concerning the PA nomenclature: the different PAs, diamine plus a diacid, A2B2 type are identified by numbers denoting the number of carbon atoms in the monomers with diamine first. Both polyamide PA 6,6 and PA 6 are widely used in many different markets and applications due to their excellent performance/cost ratios. They are, by far, the most used PAs 14206

DOI: 10.1021/acs.chemrev.6b00486 Chem. Rev. 2016, 116, 14181−14224

Chemical Reviews

Review

globally. PA 6 has a slightly lower temperature resistance versus PA 6,6 and is also slightly less expensive. As said previously, biobased adipic acid is a C6 platform chemical that can be converted by chemical processes into 1,6-hexamethylene diamine (HMDA) and caprolactam. The final aim is to prepare 100% biobased PA 6,6 and PA 6 for some special applications. Another platform chemical is lysine which is also able to be converted into caprolactam, but also into 1,5-pentanediamine (1,5-PDA), 5-aminovaleric acid, and lysinol. Lysinol/adipic acid copolymer is insoluble and cannot be melt processed.6 On the contrary 1,5-PDA has the potential to substitute HMDA for the synthesis of several PAs. 1,5-PDA is a clear liquid at room temperature, melting point = 9 °C compared to 42−45 °C for HMDA. Several companies (Ajinomoto and Toray, Cathay) propose different commercial PAs: PA 5,4; PA 5,6; PA 5,10; and PA 5,12. They possess such desirable properties as high melting points and low water absorption.260,407 It is said that manufactured PA using 1,5-PDA is pleasant to the touch, yet has the same strength and heat resistance as conventional PA fibers made from the hexamethylenediamine. It also absorbs and desorbs moisture nearly as well as cotton. 5-aminovaleric acid has been proposed for the synthesis of the homopolymer PA 5. Although PA 5 has not achieved commercial realization, it possesses properties similar to those of PA 4,6, and serves as a suitable substitute.408 The amino acid aspartic acid could lead to alanine which can be step growth polymerized and gives PA 3 with groups IV metal oxides as catalysts.259 DSM Engineering Plastics currently uses 1,4-butane diamine (also named putrescine) derived from castor oil to produce PA 4,6 and PA 4,10 polyamides. Microbial putrescine production from simple sugars, however, would add economic and sustainability benefits.261 4.2.3. Polyamides and Derivatives from Long Aliphatic Chains. Most of the biobased commercially available PAs today are synthesized from oilseed and more precisely castor oil. The starting material is the castor bean (Ricinus communis) and its oil derivatives, which are synthesized into monomers (section 3.3.1). However, as the castor oil market has been historically volatile in terms of pricing and supply, the use of sugar cane feedstock in producing biobased PAs could be a welcome change for the polyamides market. Thanks to their excellent chemical resistance, low water absorption, and good dimensional stability these PAs are used for several specific applications. PA homopolymers are produced from amino acid monomers. PA 11 marketed under the trade name Rilsan (from Arkema) is certainly the first commercialized biobased PA. The biobased ω-amino-lauric acid (ALA) from Evonik is an alternative to petroleum-based laurolactam (LL). The amino acid ALA replaces the lactam monomer LL in the manufacture of an identical polymer, PA 12. But the key ingredient in castor oil for PA copolymers synthesis is sebacic acid also known as 1,10 decanedioic acid. It has been available for a long time, but only recently it has attracted interest from industry as a biobased source for polyamides. PA 10,10 which is the step growth polymerization product of 1,10-decamethylene diamine and sebacic acid is 100% biobased and can be an alternative to long-chain highperformance polyamides PA 12 and PA 12,12 (Figure 13). It provides lower cost options versus PA 11 and 12.The PA 10,12 made from biobased C10 diamine and fossil sebacic acid polymer has a biobased carbon content of 45%.

Figure 13. Formulas of PA 10,10 and 10,12.

Considering other long-chain fatty acids, many new diacids and diamines could be obtained which could lead to new specialty PAs with properties not achievable by the fossil product PAs. One example could be PA 9,9 from 1,9nonandiamine and azelaic acid.409,410 Du Prez et al.411 described the synthesis of substituted polyamides from AB′ monomer obtained with a thiolactone derivative of 10-undecenoic acid. Nucleophilic aminolysis of thiolactone releases a thiol which reacts in a stepwise thiol−ene photopolymerization reaction with primary amines. The addition of diamines such as dimer diamine allowed increasing mechanical properties. New supramolecular ionic polyamides networks are synthesized by proton transfer reaction between Croda dimer diamines and a series of naturally occurring carboxylic acids such as malonic acid, citric acid or tartaric acid.412 The resulting solid soft materials exhibit a thermoreversible transition: it is a viscoelastic liquid at high temperatures and it exhibits self-healing properties at room temperature. Finally, Meier et al.295 proposed, as previously mentioned, the synthesis of new fatty acid-derived amine functional monomers thanks to a thiol−ene reaction. Methyl 10undecanoate, methyl oleate and methyl erucate were aminated thermally or under UV activation and were then homopolymerized to yield new biobased polyamides. 4.2.4. Semiaromatic and Rigid Polyamides and Substitutes. 4.2.4.1. Polyphthalimides, PPAs. The incorporation of aromatic units like terephthalic acid (TA) into aliphatic PAs raises the glass transition temperature Tg, as well as the melting temperature, Tm. Commercial products are either copolymers of HMDA and terephthalic acid with small amounts of isophthalic acid or homopolymers from long-chain diamines with terephthalic acid. Basically, for the realization of biobased polyphtalamides (PPAs), the same limits occurs as for the aliphatic polyamides: the availability of the monomers at an acceptable cost. The situation might be different for the PAs with longer aliphatic chains. Introduction of aromatic units can confer specific properties to the polymers that can be of high interest for different applications. One example is the high temperature Rilsan proposed by Arkema. 4.2.4.2. From Cardanol. One another biobased fatty long chain could come from amino-cardanol. Indeed, some difunctional amine monomers from cardanol were used for the synthesis of a variety of high performance polymers such as aromatic polyimides, poly(amideimide)s, polyamide, ... (Schemes 68 and 69)353,354,413,414 The effects of incorporation this pentadecyl chains on solubility and thermal properties of the polymers were evaluated (Table 2). The results demonstrated that the presence of the C15 alkyl chain hanging off the polymer backbone disrupted the packing of polymer chains as well as enhanced the solubility of these polymers when compared with reference polymers, i.e., polymers without the C15 alkyl chain. The polymers containing this pentadecyl chains exhibited a depression in Tg which could be attributed to these chains that decrease the intermolecular interactions and increase the segmental motion in the polymer backbone. A 14207

DOI: 10.1021/acs.chemrev.6b00486 Chem. Rev. 2016, 116, 14181−14224

Chemical Reviews

Review

Scheme 68. Structures of Poly(amideimide)s Obtained from Amino-Cardanol

biobased polyamides can be played by 1,4;3,6-dianhydrohexitols, or more precisely by their diamine derivatives. Hence, the synthesis of diamines from these dianhydrohexides have been the subject of a lot of papers and patents415 (cf. section 3.4.2.4). Isosorbide-(bis(propan-1-amine)) (DAPIS) was synthesized and used to prepare fully biobased PAs. To prevent degradation of sugar-derived diamines during polymerization and to ensure sufficiently high molar mass of the final products, a low temperature melt step growth polymerization of “nylon salts” followed by solid-state polymerization have been applied. The majority of these reactions have been carried out under solventfree conditions. Step growth polymerization with various contributions of DAPIS afforded a family of homo- and copolyamides, which were characterized using complementary spectroscopic techniques (FTIR, NMR). DSC and TGA, demonstrating that the increasing amounts of isohexide diamines efficiently decrease PAs melting point and slightly decrease their thermal stability.416 Similarly, 2,5-diamino-2,5-dideoxy-1,4−3,6-dianhydroiditol (isoidide diamine, IIDA) was synthesized by Wageningen UR Food and Biobased Research.313,339 A series of PAs was synthesized from heptanedioic acid (pimelic acid) and different ratios of isoidide diamine (IIDA) and butane-1,4-diamine (BDA) (Figure 14).

Scheme 69. Structures of Polyamides Obtained from AminoCardanol

Table 2. Summary of the Glass Transition Temperature of Polyimide, Polyamideimide, and Polyamide Obtained from Cardanol Diamines or Equivalent with Aliphatic Chain diamine

coreactant

Tga (°C)

Tgb (°C)

A1

Ar2 Ar3 Ar4 Ar5 Ar1 Ar2 Ar4 Ar5

206 176 159 158 286 244 206 223

330 320 305 297 NDc 316 306 330

A2

a

Glass transition of correspondent polymer with an amine from cardanol. bGlass transition of correspondent polymer with an amine without pendant chain of the cardanol. cND: not detected.

Figure 14. Polymer from the condensation between heptanedioic acid (pimelic acid) and different ratios of isoidide diamine (IIDA) and butane-1,4-diamine (BDA).

large window between glass transition temperatures and initial decomposition temperatures of the polymers containing pentadecyl chains was observed which gives an opportunity for these polymers to be melt-processed or compression molded. 4.2.4.3. From 1,4:3,6-Dianhydrohexides Derivatives. It is well-known that 1,4:3,6-dianhydrohexitols (isosorbide, isomannide and isoidide) units introduced in a polymer chain increase the rigidity and the Tg of the material. Moreover, their aliphatic structure confers to polymers a higher stability to UV than aromatic ones. Therefore, a special role in the synthesis of

The introduction of the isoidide-based diamine reduces the degree of crystallinity to form a completely amorphous polymer for high content of IIDA. However, at the same time it enhanced the rigidity of the polymer backbone, which enabled the synthesis of amorphous PA with Tg values up to 100 °C. The molar mass values measured by SEC (performed in HFIP) for the IIDA-based were in the range of 3000−10 700 g mol−1. From the authors, the lower molar masses and lower dispersities when increasing the IIDA fraction in the recipe 14208

DOI: 10.1021/acs.chemrev.6b00486 Chem. Rev. 2016, 116, 14181−14224

Chemical Reviews

Review

are caused by two phenomena: The first reason is the lower reactivity of the amine groups of IIDA being attached to a secondary carbon atom compared to the amine groups of BDA, which are attached to a primary carbon atom. The second reason is the limited thermal stability of the isoidide diamine monomer 4.2.4.4. From Furan Derivatives. Another way to prepare substituents of polyphtalamides is to introduce furanic units in the polymer backbone. Usually people are using 2,5furandicarboxylic acid (FDCA). However, a patent from Solvay propose to use indifferently the diacid or the diamine, 2,5bis(aminomethyl) furan (BAF). Semicrystalline PAs have thus been obtained.296 4.2.5. Thermoplastic Elastomers with Polyamide Blocks. Poly(ether-block-amide), PEBA, are block copolymers obtained by step growth polymerization of a carboxylic acid telechelic polyamide (PA6, PA11, and PA12) with an alcohol telechelic polyether (polytetramethylene oxide, PTMO, or poly(ethylene oxide), PEO). PEBA are high performance thermoplastic elastomers (TPEs). They are used to replace common thermoplastic polyurethanes or polyester elastomers. Some new commercial products, like Pebax Rnew 80R53 based on the castor oil chemistry (from Arkema), can present more than 90% biobased carbon.

temperatures for lysine comprising networks with a cycloaliphatic-type epoxide of 91 and 195 °C, respectively. Lysine has also directly been used as a hardener of cycloaliphatic diepoxide by Li et al.425 The adduct obtained from the cycloaliphatic epoxide and the amine could react with the carboxylate of the zwitterion of the lysine leading to a cross-linked polymer. The Tg is equal to 175 °C and the thermal stability, even though lower than materials obtained with conventional hardeners like anhydride begin decreasing near 200 °C. Concerned applications are almost in lead-free electronics Shibata et al.426,427 focused their works on commercial glycerol polyglycidyl ether (GPE; epoxide functionality =2) and polyglycerol polyglycidyl ether (PGPE; epoxide functionality = 4.1) reacted with ε-poly(L-lysine) (PL) at 110 °C in the presence of nonmodified montmorillonite (MMT). Interesting mechanical properties (modulus, tensile properties, etc.) have been published but always with a Tg lower than 60 °C. Interestingly is the aerobic biodegradability of the PGPE−PL networks in aqueous medium which was about 4% after 90 days and decreased with the MMT content. As explained previously (Scheme 41), lysinol (2,6-diamino-1hexanol) is prepared from the hydrogenation of lysine. An example of the potential utility of lysinol is demonstrated by its use as a diamine curing agent with standard epoxide monomer DGEBA. The thermosets were prepared using stoichiometrically precise amounts of amine and DGEBA and thus 1 mol lysinol (equal 4 mol N−H) per 2 mol of DGEBA and 1 mol diethylenetriamine (DETA) (equal 5 mol N−H) per 2.5 mol DGEBA. DETA is a typical petrochemical amine commonly used in adhesives and coatings application. The samples were allowed to cure at ambient temperature during 24 h with a postcure at 60−100 °C. The resulting thermosets were subjected to a number of tests to evaluate their mechanical and adhesive properties, and also, chemical resistance. The conclusion from the authors is that the thermosets prepared from lysinol and DETA exhibit similar properties.6 4.3.4. Hardeners from Vegetables Oils. Fatty renewable resources were also studied to synthesize biobased amines for epoxide curing, by direct amidification of oil or fatty acids428,429 or by thiol−ene coupling on double bonds of unsaturated fatty resources.294 In all cases, obtained materials exhibit interesting properties for coating applications with rather low Tg. For instance, Stemmelen et al.294 realized the addition of mercaptan bearing an amine on grapeseed oil under UV (Scheme 45, cf. section 3.3.3). Their amino-containing oils were used as novel curing agents for biobased epoxide monomer. They have thus proposed fully vegetable oil derived polyepoxide networks (Tg = −38 °C), from epoxidized vegetable oils and vegetable oilderived amines.294 The use of fatty acids in amine hardeners was also tested by Cornille et al.430 in order to confer a hydrophobic character before curing epoxidized biophenols. With the same thiol−ene reaction, Turunc et al.295 proposed the synthesis of new fatty acid-derived amine functional monomers. Amine-terminated polyamides may also be used as curing agents. There are some commercial products like UniRez from Arizona Chem. Priamine are also commercial products from Croda obtained from fatty acid dimers (Figure 8, cf. section 3.3.2.). Dworakowska et al.431 used these dimer diamines as a curing agent of epoxide monomers such as epoxidized cardanol, for the achievement of polyepoxide foams. These foams exhibited tailored mechanical and thermal properties and Tg

4.3. Epoxy Networks

4.3.1. Diamines as Hardeners for Epoxies. More recently, great attention was paid to renewable resourcesderived amine hardeners,3 particularly for polyepoxide thermosets,417 since they are cross-linked polymers and thus cannot be recycled easily. Therefore, biobased reactants are an interesting approach to reduce environmental impacts of these thermosets. To date, biobased acids or anhydrides are the most studied polyepoxide biobased curing agents;418,419 therefore, it is crucial to propose biobased amine hardeners. Indeed, current amines are good curing agents since they are very reactive with epoxide monomers, but they often exhibit high toxicity and high volatility, which increases their dangerousness; therefore, petro-based amines could well have to be replaced in the years to come All biobased diamines described in the previous section 3 and especially the ones used for PAs synthesis can be used as polyepoxide hardeners. We are just going to report some examples from the literature. 4.3.2. Hardeners from Chitosan. Illy et al.420 reported the use of chitosan as cross-linking agent for polyepoxides. The cross-linking reaction was performed with a diglycidyl ether of bisphenol A (DGEBA) prepolymer in suspension in water. The glass transition temperature Tg, of the obtained network was 38 °C and its degradation temperature was 420 °C. The final product was also used for coating applications. As amino hardener, chitosan was also reacted with epoxide functions of the poly(ethylene glycol) diglycidyl ether421 and was used as nanofibers in a natural rubber epoxidized biocomposite,422 to cross-link an epoxide cement,423 or even to obtain a hydrogel by graft copolymerization of epoxide terminated polydimethylsiloxane on chitosan.424 4.3.3. Hardeners from Amino-Acids. Lysine425 and tryptophan425 (Table 1, section 3.2.1) are two amino acids reported as ecofriendly cross-linking agents for epoxide monomers in the electronics industry. Both the amino and carboxyl functional groups are able to ring open an epoxide group. Li et al.425 reported the Tg and thermal degradation 14209

DOI: 10.1021/acs.chemrev.6b00486 Chem. Rev. 2016, 116, 14181−14224

Chemical Reviews

Review

close to room temperature. Recently, Zupancic et al.432 used dimer diamines with epoxide compound bearing 2 or more epoxide groups to achieve adhesive coatings with good strengths. Vijayalakshmi et al.433 reported amine functional polyamides prepared from C21 cycloaliphatic dicarboxylic acid or C36 dimer acids with various amines, namely, diethylenetriamine (DETA), triethylenetetramine (TETA), and tetraethylenepentamine (TEPA), which subsequently were reacted with epoxide monomers at different weight ratios. The cured product derived from an equivalent weight ratio of polyamide to epoxy showed better properties. 4.3.5. From Terpenes. Menthane diamine (MNDA) is somewhat less reactive than similar linear chain diamines. Its low reactivity has been used for a better control of B-stage for processing of composites. Full curing has to be done at high temperature, but too high a temperature could degrade the final network (due to the tertiary carbon atoms).434 Recently, some polyepoxide networks were also obtained by reaction between poly(propylene glycol) diglycidyl ether and menthane diamine. Thus, Wei et al.435,436 studied the effects of cross-link density on thermo-mechanical properties and shape-memory of the synthesized polyepoxide networks. Recently, Mi et al.437 reported the synthesis of a rosin based polyamide amine as a potential curing agent for epoxide systems. The cured polyepoxide networks exhibited comparable properties with petrochemical-cured polyepoxide networks: shear strength of 21.6 MPa, thermal decomposition temperature of 340 °C, and Tg of 147 °C. Wang et al.438 compared the curing behavior of rosin-based imidoamine (Figure 15) with a

compared their synthesized biobased amine with commercial cardanol amine (NX-5454 from Cardolite) for curing epoxidized cardanol (NC-514 from Cardolite). They obtained similar thermal resistance for both materials, around 320−330 °C. The glass transition temperatures were also closer, 30 and 19 °C respectively for polyepoxide with NX-5454 and cardanolcysteamine. 4.3.7. Aromatic Amine Hardeners. Vanillin (from lignin) and furfural (from sugar), both widely available renewable aromatic resources, were chosen as base building-blocks for preparing polypeoxide networks. Fache et al.440 synthesized two amine hardeners, vanillylamine and bis(furfurylamine) and used them to cure biobased epoxides, diglycidylether of methoxyhydroquinone and also triglycidylether of vanillylamine (Figure 16)

Figure 16. Formulas of vanillyl and bis(furfuryl) amines.

The polymers synthesized were bisphenol A-free, potentially fully biobased polyepoxide thermosets that displayed relatively high Tg (close to 100 °C) and degradation properties comparable to the DGEBA-Isophorone diamine, used as an industrial reference. Tg and cross-linking density were higher with the trifunctional epoxide. The vanillylamine hardener leads to lower Tg than the bis(furfuryl amine), 67 and 111 °C, respectively. Mendis et al.441 modified lignin using hydration and Mannich functionalization with triethylene tetraamine (TETA) and created polyepoxide composites with DGEBA. Aliphatic amines were added to formulations to synthesize lignin-epoxide material. The materials exhibited two-phase microstructures containing lignin-rich agglomerates. It has to be mentioned that BASF442 and Rhodia443 patented the synthesis of polyepoxide thermosets with aminated HMF called furan diaminomethyl (FDA) or aminated HMF derivatives called THFDA (tetrahydrofurandiaminoethyl for Rhodia). They both have been reacted with DGEBA and compared to MXDA networks (xylylene diamine/DGEBA. Glass transitions were measured around to 100 °C (101 and 114 °C, respectively) close to the Tg of the reference system MXDA/DGEBA (120 °C). However, these Tg are lower than the one of industrial standard isophoronediamine/DGEBA material (around 160 °C).

Figure 15. Imidoamines used for epoxide curing.

commercial aromatic amine, methylene dianiline (MDA), in order to investigate structure−property relationships. They found that epoxide cured with rosin-based imidoamine led to the highest Tg but slightly lower moduli and thermal stability than networks with MDA, which revealed the importance of the structure of curing agents. 4.3.6. From Cardanol. Cardanol is often used for the synthesis of epoxide monomers as reported before. Its use as hardener is rarer. Some team have studied its behavior as hardener for diglycidyl ether of bisphenol A (DGEBA), but obtained glass transitions lower than 100 °C which is very low.439 However, amines created from cardanol are often used to cure polyepoxide networks. Phenalkamine of DGEBA as polyepoxide network hardener leads to good properties, and greater hardness compares to the direct use of the amine. For example, the Tg is close to 80 °C with DGEBA when the amine is difunctional, and could be higher (130 °C) if the amine used for the creation of the phenalkamine is multifunctional (DETA).439 Concerning thermal stability, all materials have a degradation temperature close to 300 °C. Some teams have tried to create fully biobased polyepoxide networks. Daroman et al.350 synthesized a biobased cardanol amine by thiol−ene coupling with cysteamine. Then they

4.4. Miscellaneous Polymers

4.4.1. Other Thermoplastic Polymers. 4.4.1.1. Polyureas and Polyhydroxyurethanes. Fully renewable polyureas have been synthesized from a biobased polyester diol and combinations of biobased diisocyanates and chain extenders.444 The diol was the poly(1,2-dimethylethylene adipate) and the diisocyanates included L-lysine diisocyanate and isoidide diisocyanate. The chain extenders, which can greatly affect flow temperature and modulus, included 1,4-diaminobutane (produced from glucose by a metabolically engineered strain of E. coli), diaminoisoidide (from polysaccharide-based isomannide), and di(aminobutyl) urea (from putrescine and urea). Properties have been optimized and these fully biobased TPUs 14210

DOI: 10.1021/acs.chemrev.6b00486 Chem. Rev. 2016, 116, 14181−14224

Chemical Reviews

Review

are proposed for “a plethora of applications”, like furniture production, automotive industry, clothing, and sports appliances, as well as biomedical applications Concerning polyhydroxyurethanes (PHUs), bis-cyclic carbonates can be reacted with biobased aliphatic diamines. For example, a bis(cyclic carbonate) based on a dimer fatty acid was synthesized and reacted with dimer diamine to give a fully biobased nonisocyanate polyurethane (NIPU), without solvent and catalyst. The influence of the amine ratio and the average amine functionality were investigated by chemical, thermal, and rheological analyses. These parameters strongly impact the final NIPU properties.445 Dimer diamine was also used as comonomer by Duval et al.,446 who synthesized renewable NIPU with dicyclic carbonate prepared by reacting sebacic acid with glycerol carbonate in the presence of diciclocarbodiimide (DCC) as coupling agent and 4-dimethyaminopyridine (DMAP) as catalyst. This difunctional monomer was then reacted with several renewable linear and branched diamines such as dimer diamine to afford new non isocyanate polyhydroxyurethane materials. These PHU exhibited amorphous behavior with a Tg equal to −20 °C and were thermally stable until 200 °C. The same team has worked on the reaction between dimer diamine and dimethyl carbamate447 to synthesize new dicarbamate monomers in the presence of TBD as catalyst. Then the dimer diamine -dicarbamate was reacted with 2 diols to afford polymer in the presence of potassium carbonate (K2CO3). Material constituted with dimer diamine-dicarbamate and diol C10 (linear diol) gave a semicrystalline PHU with a Tm of 64 °C and a Tg of −19 °C. The material made with dimer diamine-dicarbamate and diolC12 (branched and unsaturated form) was amorphous (Tg = −38 °C). These 2 polymers had molar masses around 6000g/mol. Van Velthoven et al.448 presented PHU synthesized from diglycerol dicarbonate and dimer diamine in bulk conditions without any catalyst. This fully biobased PHU was yellow amorphous rubbery material (Tg = −7 °C), with a molar mass of 9000 g/mol and a higher thermal stability than the corresponding non biobased PHU. Maisonneuve et al.449 exhibited new tailor-made biobased thermoplastic poly(hydroxyurethane amide)s from dimer diamine and dicarbonates containing various amide linkages and methyl 10-undecenoate. This route allowed obtaining molar masses up to 30 000g/mol. The presence of ester or amide functions, and consequently of hydrogen bonds through amide linkages was found to drastically modify the PHUs properties. Another work reports the synthesis of new biobased isosorbide dicyclic carbonates from isosorbide. Then PHUs were synthesized by a cyclic carbonate-amine step growth polymerization with four commercial diamines including a biobased one, 1,10 diaminodecane. With this diamine, PHU exhibited a low Tg (15 °C) and a degradation temperature, Td5% = 243 °C.450 4.4.1.2. Poly(carbonate-amide). Noel et al.451 condensed ferulic acids with various hydroxylated amino acids as shown in Scheme 70. Some poly(amido carbonate) have been prepared with regioselectivity and specific utilization for sensing and/or imaging applications. Those polymers could only be obtained after esterification of the carboxylate of the amino acids. 4.4.2. Miscellaneous Thermosets Polymers. 4.4.2.1. Biobased Polyimides. An ester of lysine was recently used to

Scheme 70. Peptidic Coupling between Ferulic Acid and Protected Hydroxyl-Amino Acids Such as Serine Ester, Threonine Ester, or Tyrosine Ester

synthesized polyaspartamide with aromatic bismaleimides (BMI; Figure 17).452

Figure 17. Expected structure of bismaleimide monomer used for polyaspartamide synthesis. Reprinted with permission from ref 452. Copyright 2014 John Wiley and Sons.

The Tg of this polymer is 343 °C, Td5% is 389 °C and its properties are close to those of BMI cured with an aromatic methylene dianiline (MDA) which is known to be highly toxic. 4.4.2.2. Benzoxazines. Benzoxazine synthesis methodology is well developed on various biobased monomers synthesized from natural resources such as cardanol,341 vanillin,368 levulinic acid,453 and eugenol454 as shown in the precedent part. Cardanol-based polybenzoxazines showed improved toughness due to internal plasticization by long chain alkyl substituents. Mono- or bis-benzoxazines synthesized from cardanol have been also used in blends with petroleum-based bisphenol A benzoxazine monomers.343 It has to be noticed that these monomers are often used as viscosity (or Tg) modifier. The curing reaction of such products is usually studied by DSC analysis and could occur between 200 to 240 °C. A. The thermal stability of benzoxazine monomers from cardanol is pretty high (higher than 400 °C), generally higher than those from classical aromatic benzoxazines and depends on the amine used for the synthesis.343 Cardanol benzoxazine thus exhibits higher curing temperature but lead to soft material (low Tg). It has been proved that the thermic stability increases with the cardanol benzoxazine content.343 Agag et al.346 synthesized polybenzoxazine networks and added some wood flour to obtain composites for both reinforcement of polymer matrix. Biocomposites have also been prepared from jute fibers impregnated with cardanol based benzoxazines.341 Rao et al.345 created a low temperature curable liquid epoxide/benzoxazine monomer system from cardanol. 14211

DOI: 10.1021/acs.chemrev.6b00486 Chem. Rev. 2016, 116, 14181−14224

Chemical Reviews

Review

5. CONCLUSIONS AND FUTURE TRENDS The largest source for renewable amines remains chitin polymer due to large commercial availability and interesting properties.234 In the last decades unmodified chitosan has been widely used in a variety of applications:455 for example as wound dressing, in tissue engineering, cosmetics, food or textile industry, and in wastewater treatment. Production of amino acids is also an interesting source of reactive amines. The amino acid business is a multibillion dollar enterprise. All 20 amino acids are sold, albeit each in greatly different quantities. Amino acids are used as animal feed additives (lysine, methionine, and threonine), flavor enhancers (monosodium glutamic, serine, and aspartic acid), and as specialty nutrients in the medical field. Glutamic acid, lysine, and methionine account for the majority, by weight, of amino acids sold. Glutamic acid and lysine are made by fermentation; methionine is made by chemical synthesis. The major producers of amino acids are based in Japan, the U.S., South Korea, China, and Europe. The global market for amino acids is projected to reach US$ 20.4 × 109 by 2020, driven by robust demand from the animal feed end-use sector. The global production of glutamic acid, in monosodium glutamate form, was 3000kt in 2014 and is expected to reach 4000 kt in 2020. Llysine is commonly commercially produced as L-lysine monohydrochloride (L-lysine HCl), with purity higher than 98.5 wt %, which corresponds to 78.8 wt % of free lysine. Production of L-lysine exceeded 1500 kt in 2014 and is expected to reach 2500 kt in 2020.456 Decarboxylation of amino acids is of real interest for the synthesis of reactive building blocks in order to obtain other reactive amines. As a first conclusion, there are a lot of opportunities introduced by regulations to develop reactive amines from biomass for the preparation of renewable polymer formulations and materials. Developing highly efficient, safe, low waste, low toxicity, and atom economy processes are the keywords for green chemistry. Amination reactions have to be simple, practical, and operational, and catalysts are expected to play an important role. Ammonia plays a key-role in the synthesis of amines, and most of the commercial amines are synthesized directly or indirectly from ammonia. Renewable resources contain high quantities of hydroxyl groups; therefore, reductive oxidation could be a preferred way to prepare reactive amine from biobased molecules. This route is all the more interesting since it limits the formation of side-products. Moreover, use of cysteine-derived cysteamine could be a fully biobased way to access highly reactive aliphatic amines. Anyway, biobased reactive amines are key-molecules, needed for the developing production of numerous biobased polymers, such as polyamides and polyhydroxyurethanes or polyepoxides (and also bismaleimides, benzoxazines, and new isocyanates from Bayer). Additionally, the versatility of amine reactivity (more versatile than alcohol or acid) should be used to propose new biobased building blocks for the synthesis of hybrid polymers and particularly HNIPUs.457 Production of biobased polyepoxides and polyamides increased respectively by 25% and 13% between 2013 and 2014.458 Moreover, simple and clean processes are desired for the synthesis of biobased aromatic amines to yield polyaramides. Even if we find lots of industrial propositions concerning biobased PA, most of them concern niche markets. Concerning polyhydroxyurethanes, they are the most likely

substituents of polyurethanes, avoiding the use of toxic isocyanates, and PUs, with a global production of 18 Mt/y, rank sixth among all polymers in the world. Therefore, the biobased amines are really key molecules for the emerging industry of biobased polymers. Since very few biobased amines already exist, such as the Priamine of Croda and the diaminodecane of Arkema, both derived from vegetable oils, or the NC5454 phenalkamine of Cardolite from cashew nutshell liquid, the challenges consist really in the synthesis of new biobased and reactive amines via clean routes. Even if the literature contains lots of academic publications and industrial patents depicting the synthesis of new biobased amines, all biobased chemical platforms suffer from industrial developments and few industrial products are emerging. Some decisions are to be taken to choose access routes to biobased amines: for example, furfural-derived amines seem easier to synthesize compared to isosorbide-derived ones. From lysine, Bayer produces a C5 diamine for PA and also both lysinederived isocyanate and isocyanurate. Such platform developments leading to versatile biobased reactants that could be used in PA, PU, but also polyepoxide and PHU are really to be favored. Eventually, petroleum refining started at the end of the 19th century and industrial development of synthetic polymers began in the 1930s. These processes have been continuously optimized since the beginning, in terms of economy and environment. Biorefinery of biomass is nowadays just at the beginning of its history, and several processes are still to be developed or optimized, according to the type of building blocks (i.e., both hardeners and epoxy monomers) or chemical routes. Finally, unlike for vegetable oil and sugar derivatives, biorefining is particularly a challenge for aromatic molecules since they do not benefit from previous industrial agro- and food industry.

AUTHOR INFORMATION Corresponding Author

*E-mail: [email protected]. Notes

The authors declare no competing financial interest. Biographies Vincent Froidevaux was born in France in 1988. He first graduated from the engineering school of Chemistry of Montpellier in 2011. He obtained his Ph.D. degree in 2014 at the University of Montpellier on sulfonated polymers cross-linking with controlled thiol-Michael reaction with Hutchinson Company. Then he continued with a postdoctoral position with Hutchinson Company. In 2016, he joined Chryso as a research engineer. He is the coauthor of 4 scientific publications and 1 patent. Claire Negrell was born in France in 1977. She first graduated from the engineering school of Chemistry of Montpellier in 2000. In 2001, she joined CNRS in the laboratory of Pr. Guerin in the University of Créteil (FRANCE). In 2005, she joined the team “Macromolecular Engineering and Architectures” (IAM) of the Institute Charles Gerhardt (Montpellier, FRANCE). She obtained her Ph.D. degree in 2010 at the University Montpellier. Since 2010, her main research projects are in the field of phosphorus-containing polymers as well as in the phosphorus functionalization of biobased monomers and polymers. She is the coauthor of 29 scientific publications. 14212

DOI: 10.1021/acs.chemrev.6b00486 Chem. Rev. 2016, 116, 14181−14224

Chemical Reviews

Review

(9) Andreev, V. P.; Nizhnik, Y. P.; Lebedeva, N. S. New basicity/ nucleophilicity scale on the basis of parameters of formation of axial n,ν-complexes derived from tetraphenylporphyrinatozinc(II) and base/nucleophile as ligand. Russ. J. Org. Chem. 2008, 44, 906−915. (10) Hall, H. K.; Bates, R. B. Correlation of alkylamine nucleophilicities with their basicities. Tetrahedron Lett. 2012, 53, 1830−1832. (11) Bunting, J. W.; Mason, J. M.; Heo, C. K. M. Nucleophilicity towards a saturated carbon atom: rate constants for the aminolysis of methyl 4-nitrobenzenesulfonate in aqueous solution. A comparison of the n and N+ parameters for amine nucleophilicity. J. Chem. Soc., Perkin Trans. 2 1994, 2291−2300. (12) Heo, C. K. M.; Bunting, J. W. Nucleophilicity towards a vinylic carbon atom: rate constants for the addition of amines to the 1methyl-4-vinylpyridinium cation in aqueous solution. J. Chem. Soc., Perkin Trans. 2 1994, 2279−2290. (13) Edwards, J. O.; Pearson, R. G. The Factors Determining Nucleophilic Reactivities. J. Am. Chem. Soc. 1962, 84, 16−24. (14) Armstrong, V. C.; Farlow, D. W.; Moodie, R. B. The hydrolysis of amides, esters, and related compounds in acid solution. Part I. Amides, carbamates, and ureas. J. Chem. Soc. B 1968, 1099−1103. (15) Cohen, T.; Lipowitz, J. Acid-Catalyzed Amide Hydrolysis Assisted by a Neighboring Amide Group. J. Am. Chem. Soc. 1964, 86, 5611−5616. (16) Yates, K.; Riordan, J. C. The HA acidity function and the mechanism of amide hydrolysis in hydrochloric acid. Can. J. Chem. 1965, 43, 2328−2335. (17) Nagase, K.; Sakaguchi, K. Alkaline hydrolysis of polyacrylamide. J. Polym. Sci., Part A: Gen. Pap. 1965, 3, 2475−2482. (18) Micovic, V.; Mihailovic, M. The Reduction of Acid Amides with Lithium Aluminum Hydride. J. Org. Chem. 1953, 18, 1190−1200. (19) Wallis, E. S.; Lane, J. F. The Hofmann Reaction. Organic Reactions; John Wiley & Sons: New York, 2004. (20) Mader, P. M. Hydrolysis kinetics for p-dimethylaminophenyl isocyanate in aqueous solutions. J. Org. Chem. 1968, 33, 2253−2260. (21) Bion, N.; Saussey, J.; Haneda, M.; Daturi, M. Study by in situ FTIR spectroscopy of the SCR of NOx by ethanol on Ag/Al2O3 Evidence of the role of isocyanate species. J. Catal. 2003, 217, 47−58. (22) Fauss, R.; Meiners, H. J.; Rasshofer, W.; Meyborg, H., Bayer, A.G. Fed. Rep. Ger. Polyamines and their use in preparing polyurethanes, DE3223400A1, 1983. (23) Fauss, R.; Meiners, H. J.; Rasshofer, W.; Meyborg, H., Bayer, A.G. Fed. Rep. Ger. Polyamines and their use for producing polyurethanes, DE3223398A1, 1983. (24) Rasshofer, W.; Dieterich, D., Bayer, A.-G. Fed. Rep. Ger. Onestep production of polyamines from isocyanate-containing prepolymers and their use in preparing polyurethanes, DE3223397A1, 1983. (25) Knauf, T.; Merkel, M. Bayer Intellectual Property GmbH, Germany. Process for continuous production of nitrobenzene by nitration of benzene with improved product separation/purification, WO2013030223A1, 2013. (26) Kozlov, A. I.; Grunskii, V. N.; Zbarskii, V. L.; Starodubtsev, V. S.; Efremov, A. I.; Khitrov, N. V.; Zhubrikov, A. V. OAO “Khimprom”, Russia. Method for catalytic nitration of aromatic hydrocarbons on porous catalysts, RU2309142C1, 2007. (27) Boyer, C.; Otazaghine, B.; Boutevin, B.; Joly-Duhamel, C.; Robin, J.-J. Synthesis of maleimide-terminated n-butyl acrylate oligomers by atom transfer radical polymerization: Study of their copolymerization with vinyl ethers. J. Polym. Sci., Part A: Polym. Chem. 2005, 43, 4303−4322. (28) Lauwiner, M.; Rys, P.; Wissmann, J. Reduction of aromatic nitro compounds with hydrazine hydrate in the presence of an iron oxide hydroxide catalyst. I. The reduction of monosubstituted nitrobenzenes with hydrazine hydrate in the presence of ferrihydrite. Appl. Catal., A 1998, 172, 141−148. (29) Rollas, S. Reduction of aromatic and heteroaromatic azo compounds with hydrazine hydrate. Marmara Pharm. J. 2010, 14, 41− 46.

Sylvain Caillol was born in 1974. He received his M.Sc. Degree in Chemistry from the Engineering School of Chemistry of Montpellier in 1998. Then he received his Ph.D. degree in Polymer Science in 2001 from the University of Bordeaux. Subsequently he joined Rhodia group and headed the Polymer Department in the Research Center of Aubervilliers. In 2007 he joined CNRS in the University of Montpellier where he heads a research topic dedicated to Green Chemistry for polymer science. He is cofounder and Director of ChemSuD Chair. Coauthor of several articles and patents, he won the Innovative Techniques for Environment award. Jean-Pierre Pascault is Emeritus from 2005 at National Institute of Applied Science (INSA) Lyon France. He was Professor in the same Institute from 1983 to October 2005; Director of the Laboratory of Macromolecular Materials (Associated to CNRS) from 1982 to 1998; Director of a CNRS Polymer Network Group (FR CNRS) from 2000 to 2006; and President of the French Polymer Group, GFP, and of the Polymer Division of the French Chemical Society, SFC from 2001 to 2004. Key words for his research activities are polymer chemistry; polymer network formation including polyepoxies, UP resins, polyurethanes and polyacrylates; thermoplastic/thermoset blends; Nanostructured Thermosets; and for composites, adhesives and coatings applications. He has authored over 330 scientific publications including several book chapters, two books (Thermosetting Polymers and Epoxy Polymers: New Materials and Innovations), and 35 patents. Bernard Boutevin was born in 1948. He received his Ph.D. degree in 1974 from the University of Montpellier and joined CNRS. Subsequently he created a Polymer laboratory in Montpellier in 1985. Then he was director of several research units with more than 100 researchers. He was appointed full Professor in 1999. His research interests included the use of fluorine, phosphorous, and silicon in monomers and polymers. Recently, he became interested in the synthesis of new biobased building blocks and polymers. Now Professor Emeritus, he was codirector of more than 200 Ph.D. theses and coauthor of more than 250 patents and 1000 articles.

ACKNOWLEDGMENTS Authors would like to thank Ms. Coline Voirin for her help at the beginning of this work. REFERENCES (1) Blain, M.; Jean-Gérard, L.; Benazet, D.; Boutevin, B.; Andrioletti, B.; Caillol, S. Synthesis of Aminotelechelic Prepolymers to Circumvent the Carbonation of Amines in Epoxy Coatings. Macromol. Mater. Eng. 2016, 301, 682−693. (2) Shima, S.; Sakai, H. Polylysine produced by Streptomyces. Agric. Biol. Chem. 1977, 41, 1807−1809. (3) Ding, C.; Matharu, A. S. Recent Developments on Biobased Curing Agents: A Review of Their Preparation and Use. ACS Sustainable Chem. Eng. 2014, 2, 2217−2236. (4) Anastas, P. T.; Warner, J. C. Green Chemistry: Theory and Practice; Oxford University Press: New York, 1998. (5) Scott, E.; Peter, F.; Sanders, J. Biomass in the manufacture of industrial products-the use of proteins and amino acids. Appl. Microbiol. Biotechnol. 2007, 75, 751−762. (6) Metkar, P. S.; Scialdone, M. A.; Moloy, K. G. Lysinol: a renewably resourced alternative to petrochemical polyamines and aminoalcohols. Green Chem. 2014, 16, 4575−4586. (7) Kanzian, T.; Nigst, T. A.; Maier, A.; Pichl, S.; Mayr, H. Nucleophilic Reactivities of Primary and Secondary Amines in Acetonitrile. Eur. J. Org. Chem. 2009, 2009, 6379−6385. (8) Simonyan, G. S.; Beileryan, N. M.; Pirumyan, E. G.; Roque, J. P.; Boyer, B. Effects of Solvents and the Structure of Amines on the Rates of Reactions of α,β-Unsaturated Nitriles and Amides with Secondary Amines. Kinet. Catal. 2001, 42, 474−478. 14213

DOI: 10.1021/acs.chemrev.6b00486 Chem. Rev. 2016, 116, 14181−14224

Chemical Reviews

Review

(51) Bianchini, C.; Dal Santo, V.; Meli, A.; Oberhauser, W.; Psaro, R.; Vizza, F. Preparation, Characterization, and Performance of the Supported Hydrogen-Bonded Ruthenium Catalyst [(sulphos)Ru(NCMe)3](OSO2CF3)/SiO2. Comparisons with Analogous Homogeneous and Aqueous-Biphase Catalytic Systems in the Hydrogenation of Benzylideneacetone and Benzonitrile. Organometallics 2000, 19, 2433−2444. (52) Kukula, P.; Gabova, V.; Koprivova, K.; Trtik, P. Selective hydrogenation of unsaturated nitriles to unsaturated amines over amorphous CoB and NiB alloys doped with chromium. Catal. Today 2007, 121, 27−38. (53) Amundsen, L. H.; Nelson, L. S. Reduction of Nitriles to Primary Amines with Lithium Aluminum Hydride. J. Am. Chem. Soc. 1951, 73, 242−244. (54) Saavedra, J. Z.; Resendez, A.; Rovira, A.; Eagon, S.; Haddenham, D.; Singaram, B. Reaction of InCl3 with Various Reducing Agents: InCl3−NaBH4-Mediated Reduction of Aromatic and Aliphatic Nitriles to Primary Amines. J. Org. Chem. 2012, 77, 221−228. (55) Krimen, L. I.; Cota, D. J. The Ritter Reaction. Organic Reactions; John Wiley & Sons: New York, 2004. (56) Ritter, J. J.; Minieri, P. P. A New Reaction of Nitriles. I. Amides from Alkenes and Mononitriles. J. Am. Chem. Soc. 1948, 70, 4045− 4048. (57) Lebedev, M. Y.; Erman, M. B. Lower primary alkanols and their esters in a Ritter-type reaction with nitriles. An efficient method for obtaining N-primary-alkyl amides. Tetrahedron Lett. 2002, 43, 1397− 1399. (58) Fernholz, H.; Schmidt, H. J. tert-Butyl Acetate as Alkylating Agent. Angew. Chem., Int. Ed. Engl. 1969, 8, 521−521. (59) Guérinot, A.; Reymond, S.; Cossy, J. Ritter Reaction: Recent Catalytic Developments. Eur. J. Org. Chem. 2012, 2012, 19−28. (60) Werkmeister, S.; Bornschein, C.; Junge, K.; Beller, M. Selective Ruthenium-Catalyzed Transfer Hydrogenations of Nitriles to Amines with 2-Butanol. Chem. - Eur. J. 2013, 19, 4437−4440. (61) Hamid, M. H. S. A.; Slatford, P. A.; Williams, J. M. J. Borrowing Hydrogen in the Activation of Alcohols. Adv. Synth. Catal. 2007, 349, 1555−1575. (62) Guillena, G.; J. Ramón, D.; Yus, M. Hydrogen Autotransfer in the N-Alkylation of Amines and Related Compounds using Alcohols and Amines as Electrophiles. Chem. Rev. 2010, 110, 1611−1641. (63) Guérin, C.; Bellosta, V. r.; Guillamot, G. r.; Cossy, J. Mild Nonepimerizing N-Alkylation of Amines by Alcohols without Transition Metals. Org. Lett. 2011, 13, 3534−3537. (64) Lawrence, S. A. Amines: Synthesis, Properties and Applications; Cambridge University Press: New York, 2004. (65) Klein, H. P.; Burton, B. L.; Forkner, M. W.; Alexander, D. C.; Renken, T. L.; Godinich, C. E. Huntsman Petrochemical Corporation, U.S.A. Huntsman Spec. Chemical Corp. Polyether-polyamine curing agents for epoxy resins and polyureas, WO2004020506A2, 2004. (66) Imm, S.; Baehn, S.; Neubert, L.; Neumann, H.; Beller, M. An Efficient and General Synthesis of Primary Amines by RutheniumCatalyzed Amination of Secondary Alcohols with Ammonia. Angew. Chem., Int. Ed. 2010, 49, 8126−8129. (67) Fischer, A.; Mallat, T.; Baiker, A. Amination of diols and polyols to acyclic amines. Catal. Today 1997, 37, 167−189. (68) Cano, R.; Ramón, D. J.; Yus, M. Impregnated Ruthenium on Magnetite as a Recyclable Catalyst for the N-Alkylation of Amines, Sulfonamides, Sulfinamides, and Nitroarenes Using Alcohols as Electrophiles by a Hydrogen Autotransfer Process. J. Org. Chem. 2011, 76, 5547−5557. (69) Shi, F.; Tse, M. K.; Zhou, S.; Pohl, M.-M.; Radnik, J.; Hübner, S.; Jähnisch, K.; Brückner, A.; Beller, M. Green and Efficient Synthesis of Sulfonamides Catalyzed by Nano-Ru/Fe3O4. J. Am. Chem. Soc. 2009, 131, 1775−1779. (70) Kim, J. W.; Yamaguchi, K.; Mizuno, N. Heterogeneously catalyzed selective N-alkylation of aromatic and heteroaromatic amines with alcohols by a supported ruthenium hydroxide. J. Catal. 2009, 263, 205−208.

(30) Gowda, S.; Gowda, D. C. Application of hydrazinium monoformate as new hydrogen donor with Raney nickel: a facile reduction of nitro and nitrile moieties. Tetrahedron 2002, 58, 2211− 2213. (31) Raju, B.; Ragul, R.; Sivasankar, B. N. A new reagent for selective reduction of nitro group. Indian J. Chem., Sect. B: Org. Chem. Incl. Med. Chem. 2009, 48B, 1315−1318. (32) Göksu, H.; Ho, S. F.; Metin, Ö .; Korkmaz, K.; Mendoza Garcia, A.; Gültekin, M. S.; Sun, S. Tandem Dehydrogenation of Ammonia Borane and Hydrogenation of Nitro/Nitrile Compounds Catalyzed by Graphene-Supported NiPd Alloy Nanoparticles. ACS Catal. 2014, 4, 1777−1782. (33) Cantillo, D.; Moghaddam, M. M.; Kappe, C. O. Hydrazinemediated Reduction of Nitro and Azide Functionalities Catalyzed by Highly Active and Reusable Magnetic Iron Oxide Nanocrystals. J. Org. Chem. 2013, 78, 4530−4542. (34) Allais, C.; Grassot, J.-M.; Rodriguez, J.; Constantieux, T. MetalFree Multicomponent Syntheses of Pyridines. Chem. Rev. 2014, 114, 10829−10868. (35) Frank, R. L.; Seven, R. P. Pyridines. IV. A Study of the Chichibabin Synthesis. J. Am. Chem. Soc. 1949, 71, 2629−2635. (36) Alexander, F. P.; Simonov, A. M.; Doron’kin, V. N. Advances in the Study of the Chichibabin Reaction. Russ. Chem. Rev. 1978, 47, 1042. (37) Lawin, P. B.; Sherman, A. R.; Grendze, M. P. Reilly Industries, Inc., U.S.A. Improved Chichibabin aminations of pyridine bases, WO9600216A1, 1996. (38) McGill, C. K.; Bailey, T. D. Reilly Tar and Chemical Corp., U.S.A. 2-Alkylamino- and 2-amino-5-alkylpyridines, US4405790A, 1983. (39) Ikawa, T.; Fujita, Y.; Mizusaki, T.; Betsuin, S.; Takamatsu, H.; Maegawa, T.; Monguchi, Y.; Sajiki, H. Selective N-alkylation of amines using nitriles under hydrogenation conditions: facile synthesis of secondary and tertiary amines. Org. Biomol. Chem. 2012, 10, 293−304. (40) Gomez, S.; Peters, J. A.; Maschmeyer, T. The reductive amination of aldehydes and ketones and the hydrogenation of nitriles: mechanistic aspects and selectivity control. Adv. Synth. Catal. 2002, 344, 1037−1057. (41) Werkmeister, S.; Bornschein, C.; Junge, K.; Beller, M. Ruthenium-catalyzed transfer hydrogenation of nitriles: reduction and subsequent N-monoalkylation to secondary amines. Eur. J. Org. Chem. 2013, 2013, 3671−3674. (42) Enthaler, S.; Junge, K.; Addis, D.; Erre, G.; Beller, M. A practical and benign synthesis of primary amines through ruthenium-catalyzed reduction of nitriles. ChemSusChem 2008, 1, 1006−1010. (43) Suarez, T.; Fontal, B. Hydrogenation reactions with ruthenium complex RuCl2(TRIPHOS). J. Mol. Catal. 1988, 45, 335−344. (44) Chin, C. S.; Lee, B. Hydrogenation of nitriles with iridiumtriphenylphosphine complexes. Catal. Lett. 1992, 14, 135−140. (45) Dubois, J.-L.; Gillet, J.-P. Arkema, France. Coproduction of cyclic carbonates and fatty nitriles and/or fatty amines from natural oils, WO2008145941A2, 2008. (46) Gillet, J.-P. Arkema, France. New functional compounds with an isosorbide core or isosorbide isomer, fabrication process, and applications of these compounds, WO2008145921A2, 2008. (47) Xie, X.; Liotta, C. L.; Eckert, C. A. CO2-Protected Amine Formation from Nitrile and Imine Hydrogenation in Gas-Expanded Liquids. Ind. Eng. Chem. Res. 2004, 43, 7907−7911. (48) De Bellefon, C.; Fouilloux, P. Homogeneous and heterogeneous hydrogenation of nitriles in a liquid phase: chemical, mechanistic, and catalytic aspects. Catal. Rev.: Sci. Eng. 1994, 36, 459−506. (49) Hegedus, L.; Mathe, T. Selective heterogeneous catalytic hydrogenation of nitriles to primary amines in liquid phase. Appl. Catal., A 2005, 296, 209−215. (50) Yoshida, T.; Okano, T.; Otsuka, S. Catalytic hydrogenation of nitriles and dehydrogenation of amines with the rhodium(I) hydrido compounds [RhH[P(CHMe 2 ) 3 ] 3 ] and [Rh 2 H 2 (μ-N 2 )[P(cyclohexyl)3]4]. J. Chem. Soc., Chem. Commun. 1979, 870−871. 14214

DOI: 10.1021/acs.chemrev.6b00486 Chem. Rev. 2016, 116, 14181−14224

Chemical Reviews

Review

(71) Yamaguchi, K.; Koike, T.; Kim, J. W.; Ogasawara, Y.; Mizuno, N. Highly dispersed ruthenium hydroxide supported on titanium oxide effective for liquid-phase hydrogen-transfer reactions. Chem. - Eur. J. 2008, 14, 11480−11487. (72) Zhang, Y.-Q.; Wei, X.-W.; Yu, R. Fe3O4 NanoparticlesSupported Palladium-Bipyridine Complex: Effective Catalyst for Suzuki Coupling Reaction. Catal. Lett. 2010, 135, 256−262. (73) De Luca, L.; Porcheddu, A. Microwave-Assisted Synthesis of Polysubstituted Benzimidazoles by Heterogeneous Pd-Catalyzed Oxidative C-H Activation of Tertiary Amines. Eur. J. Org. Chem. 2011, 2011, 5791−5795. (74) Zhang, Y.; Qi, X.; Cui, X.; Shi, F.; Deng, Y. Palladium catalyzed N-alkylation of amines with alcohols. Tetrahedron Lett. 2011, 52, 1334−1338. (75) He, W.; Wang, L.; Sun, C.; Wu, K.; He, S.; Chen, J.; Wu, P.; Yu, Z. Pt-Sn/γ-Al2O3-Catalyzed Highly Efficient Direct Synthesis of Secondary and Tertiary Amines and Imines. Chem. - Eur. J. 2011, 17, 13308−13317. (76) Wang, L.; He, W.; Wu, K.; He, S.; Sun, C.; Yu, Z. Heterogeneous bimetallic Pt-Sn/γ-Al2O3 catalyzed direct synthesis of diamines from N-alkylation of amines with diols through a borrowing hydrogen strategy. Tetrahedron Lett. 2011, 52, 7103−7107. (77) He, W.; He, S.; Sun, C.; Wu, K.; Wang, L.; Yu, Z. Heterogeneous bimetallic Pt-Sn/γ-Al2O3 catalyzed N-alkylation of amines: efficient synthesis of secondary and tertiary amines. Cuihua Xuebao 2012, 33, 717−722. (78) Zotova, N.; Roberts, F. J.; Kelsall, G. H.; Jessiman, A. S.; Hellgardt, K.; Hii, K. K. Catalysis in flow: Au-catalyzed alkylation of amines by alcohols. Green Chem. 2012, 14, 226−232. (79) He, L.; Lou, X.-B.; Ni, J.; Liu, Y.-M.; Cao, Y.; He, H.-Y.; Fan, K.N. Efficient and clean Gold-catalyzed one-pot selective N-alkylation of amines with alcohols. Chem. - Eur. J. 2010, 16, 13965−13969. (80) Ishida, T.; Takamura, R.; Takei, T.; Akita, T.; Haruta, M. Support effects of metal oxides on gold-catalyzed one-pot N-alkylation of amine with alcohol. Appl. Catal., A 2012, 413−414, 261−266. (81) Peng, Q.; Zhang, Y.; Shi, F.; Deng, Y. Fe2O3-supported nanogold catalyzed one-pot synthesis of N-alkylated anilines from nitroarenes and alcohols. Chem. Commun. 2011, 47, 6476−6478. (82) Shimizu, K.; Nishimura, M.; Satsuma, A. γ-Alumina-Supported Silver Cluster for N-Benzylation of Anilines with Alcohols. ChemCatChem 2009, 1, 497−503. (83) Cui, X.; Zhang, Y.; Shi, F.; Deng, Y. Organic Ligand-Free Alkylation of Amines, Carboxamides, Sulfonamides, and Ketones by Using Alcohols Catalyzed by Heterogeneous Ag/Mo Oxides. Chem. Eur. J. 2011, 17, 1021−1028. (84) Winans, C. F.; Adkins, H. Alkylation of amines as catalyzed by nickel. J. Am. Chem. Soc. 1932, 54, 306−312. (85) Imabeppu, M.; Kiyoga, K.; Okamura, S.; Shoho, H.; Kimura, H. One-step amination of α,ω-alkylenediols over Cu/Ni-based catalysts. Catal. Commun. 2009, 10, 753−757. (86) Yu, X.; Liu, C.; Jiang, L.; Xu, Q. Manganese Dioxide Catalyzed N-Alkylation of Sulfonamides and Amines with Alcohols under Air. Org. Lett. 2011, 13, 6184−6187. (87) Likhar, P. R.; Arundhathi, R.; Kantam, M. L.; Prathima, P. S. Amination of Alcohols Catalyzed by Copper-Aluminum Hydrotalcite: A Green Synthesis of Amines. Eur. J. Org. Chem. 2009, 2009, 5383− 5389. (88) Shimizu, K.-i.; Shimura, K.; Nishimura, M.; Satsuma, A. Silver cluster-promoted heterogeneous copper catalyst for N-alkylation of amines with alcohols. RSC Adv. 2011, 1, 1310−1317. (89) He, J.; Yamaguchi, K.; Mizuno, N. Selective synthesis of secondary amines via N-alkylation of primary amines and ammonia with alcohols by supported copper hydroxide catalysts. Chem. Lett. 2010, 39, 1182−1183. (90) Martinez, R.; Ramon, D. J.; Yus, M. Selective N-monoalkylation of aromatic amines with benzylic alcohols by a hydrogen autotransfer process catalyzed by unmodified magnetite. Org. Biomol. Chem. 2009, 7, 2176−2181.

(91) Reddy, B. V. S.; Krishna, A. S.; Ganesh, A. V.; Kumar, G. G. K. S. N. Nano Fe3O4 as magnetically recyclable catalyst for the synthesis of α-aminophosphonates in solvent-free conditions. Tetrahedron Lett. 2011, 52, 1359−1362. (92) Cui, X.; Dai, X.; Deng, Y.; Shi, F. Development of a general nonnoble metal catalyst for the benign amination of alcohols with amines and ammonia. Chem. - Eur. J. 2013, 19, 3665−3675. (93) Grigg, R.; Mitchell, T. R. B.; Sutthivaiyakit, S.; Tongpenyai, N. Transition metal-catalyzed N-alkylation of amines by alcohols. J. Chem. Soc., Chem. Commun. 1981, 611−612. (94) Watanabe, Y.; Tsuji, Y.; Ohsugi, Y. The ruthenium catalyzed Nalkylation and N-heterocyclization of aniline using alcohols and aldehydes. Tetrahedron Lett. 1981, 22, 2667−2670. (95) Fujita, K.-i.; Li, Z.; Ozeki, N.; Yamaguchi, R. N-Alkylation of amines with alcohols catalyzed by a Cp*Ir complex. Tetrahedron Lett. 2003, 44, 2687−2690. (96) Abbenhuis, R. A. T. M.; Boersma, J.; van Koten, G. RutheniumComplex-Catalyzed N-(Cyclo)alkylation of Aromatic Amines with Diols. Selective Synthesis of N-(ω-Hydroxyalkyl)anilines of Type PhNH(CH2)nOH and of Some Bioactive Arylpiperazines. J. Org. Chem. 1998, 63, 4282−4290. (97) Huh, K. T.; Shim, S. C.; Doh, C. H. Ruthenium complex catalyzed synthesis of diamino compounds from α,ω-diols and secondary amines. Bull. Korean Chem. Soc. 1990, 11, 45−49. (98) Saidi, O.; Blacker, A. J.; Farah, M. M.; Marsden, S. P.; Williams, J. M. J. Iridium-catalysed amine alkylation with alcohols in water. Chem. Commun. 2010, 46, 1541−1543. (99) Blank, B.; Michlik, S.; Kempe, R. Selective iridium-catalyzed alkylation of (hetero)aromatic amines and diamines with alcohols under mild reaction conditions. Chem. - Eur. J. 2009, 15, 3790−3799. (100) Kroutil, W.; Fischereder, E.-M.; Fuchs, C. S.; Lechner, H.; Mutti, F. G.; Pressnitz, D.; Rajagopalan, A.; Sattler, J. H.; Simon, R. C.; Siirola, E. Asymmetric Preparation of prim-, sec-, and tert-Amines Employing Selected Biocatalysts. Org. Process Res. Dev. 2013, 17, 751− 759. (101) Gunanathan, C.; Milstein, D. Selective synthesis of primary amines directly from alcohols and ammonia. Angew. Chem., Int. Ed. 2008, 47, 8661−8664. (102) Shimizu, K.-i.; Kon, K.; Onodera, W.; Yamazaki, H.; Kondo, J. N. Heterogeneous Ni Catalyst for Direct Synthesis of Primary Amines from Alcohols and Ammonia. ACS Catal. 2013, 3, 112−117. (103) Pingen, D.; Mueller, C.; Vogt, D. Direct Amination of Secondary Alcohols Using Ammonia. Angew. Chem., Int. Ed. 2010, 49, 8130−8133. (104) Frauenkron, M.; Melder, J.-P.; Ruider, G.; Rossbacher, R.; Höke, H. Ethanolamines and Propanolamines. Ullmann’s Encyclopedia of Industrial Chemistry; John Wiley & Sons: New York, 2000. (105) Buter, R.; Steenbergen, A.; Geurink, P. J. A.; Scherer, T., Akzo Nobel, N. V. Aqueous crosslinkable binder composition, coating, its preparation, coated substrate and lacquer or sealing composition, WO2001090265A1, 2001. (106) Seeboth, H. The Bucherer Reaction and the Preparative Use of its Intermediate Products. Angew. Chem., Int. Ed. Engl. 1967, 6, 307− 317. (107) Mizuno, M.; Yamano, M. A new practical one-pot conversion of phenols to anilines. Org. Lett. 2005, 7, 3629−3631. (108) Weidner, J. J.; Peet, N. P. Direct conversion of hydroxy aromatic compounds to heteroarylamines via a one-pot Smiles rearrangement procedure. J. Heterocycl. Chem. 1997, 34, 1857−1860. (109) Fox, M. A.; Whitesell, J. K. Organic Chemistry, 3rd ed.; Wiley: New York, 2003. (110) Speranza, G. P.; Lichtenwalter, M. Jefferson Chemical Co., Inc. Aminoethylpiperazine, US3055901, 1962. (111) Malik, A. A.; Tzeng, D.; Cheng, P.; Baum, K. Synthesis of fluorinated diisocyanates. J. Org. Chem. 1991, 56, 3043−3044. (112) Walkowiak-Kulikowska, J.; Szwajca, A.; Boschet, F.; Gouverneur, V.; Ameduri, B. Iodine Transfer Copolymerization of Fluorinated α-Methylstyrenes with Styrene Using 1-Iodoperfluorohex14215

DOI: 10.1021/acs.chemrev.6b00486 Chem. Rev. 2016, 116, 14181−14224

Chemical Reviews

Review

ane as the Chain Transfer Agent. Macromolecules 2014, 47, 8634− 8644. (113) Szonyi, F.; Guennouni, F.; Cambon, A. New methods of preparation of 2-(perfluoroalkyl)ethylamines. J. Fluorine Chem. 1991, 55, 85−92. (114) Soules, A.; Vazquez, C. P.; Ameduri, B.; Joly-Duhamel, C.; Essahli, M.; Boutevin, B. Use of fluorinated maleimide and telechelic bismaleimide for original hydrophobic and oleophobic polymerized networks. J. Polym. Sci., Part A: Polym. Chem. 2008, 46, 3214−3228. (115) Bräse, S.; Gil, C.; Knepper, K.; Zimmermann, V. Organic Azides: An Exploding Diversity of a Unique Class of Compounds. Angew. Chem., Int. Ed. 2005, 44, 5188−5240. (116) Kolb, H. C.; Finn, M. G.; Sharpless, K. B. Click Chemistry: Diverse Chemical Function from a Few Good Reactions. Angew. Chem., Int. Ed. 2001, 40, 2004−2021. (117) Osby, J. O.; Martin, M. G.; Ganem, B. An exceptionally mild deprotection of phthalimides. Tetrahedron Lett. 1984, 25, 2093−2096. (118) Ragnarsson, U.; Grehn, L. Novel Gabriel reagents. Acc. Chem. Res. 1991, 24, 285−289. (119) Sheehan, J. C.; Bolhofer, W. A. An improved procedure for the condensation of potassium phthalimide with organic halides. J. Am. Chem. Soc. 1950, 72, 2786−2788. (120) Sen, S. E.; Roach, S. L. A convenient two-step procedure for the synthesis of substituted allylic amines from allylic alcohols. Synthesis 1995, 1995, 756−758. (121) Mongondry, P.; Bonnans-Plaisance, C.; Jean, M.; Tassin, J. F. Mild synthesis of amino-poly(ethylene glycol)s. Application to steric stabilization of clays. Macromol. Rapid Commun. 2003, 24, 681−685. (122) Liu, H.; Zhou, Y.; Yan, X.; Chen, C.; Liu, Q.; Xi, C. CopperMediated Amidation of Alkenylzirconocenes with Acyl Azides: Formation of Enamides. Org. Lett. 2013, 15, 5174−5177. (123) Shioiri, T.; Ninomiya, K.; Yamada, S. Diphenylphosphoryl azide. New convenient reagent for a modified Curtius reaction and for peptide synthesis. J. Am. Chem. Soc. 1972, 94, 6203−6205. (124) Linke, S.; Tisue, G. T.; Lwowski, W. Curtius and Lossen rearrangements. II. Pivaloyl azide. J. Am. Chem. Soc. 1967, 89, 6308− 6310. (125) Zakharkin, L. I.; Kalinin, V. N.; Gedymin, V. V. The Wolff, Beckmann, Hofmann, Curtius and Schmidt rearrangements in the series of 3-o-carborane derivatives: 1,2-Dicarba-closo-dodecaboranes. Tetrahedron 1971, 27, 1317−1322. (126) Léger, S.; Bayly, C. I.; Black, W. C.; Desmarais, S.; Falgueyret, J.-P.; Massé, F.; Percival, M. D.; Truchon, J.-F. Primary amides as selective inhibitors of cathepsin K. Bioorg. Med. Chem. Lett. 2007, 17, 4328−4332. (127) Moriwake, T.; Hamano, S.; Saito, S.; Torii, S.; Kashino, S. Synthesis of the chiral (8S)-7-aza-1,3(E),9-decatriene system from natural α-amino acids and its intramolecular Diels-Alder reaction directed toward chiral trans-hydroisoquinolones. J. Org. Chem. 1989, 54, 4114−4120. (128) Gomez, S.; Peters, J. A.; van der Waal, J. C.; van den Brink, P. J.; Maschmeyer, T. The rationalization of catalyst behaviour in the reductive amination of benzaldehyde with ammonia using a simple computer model. Appl. Catal., A 2004, 261, 119−125. (129) Bodis, J.; Lefferts, L.; Mueller, T. E.; Pestman, R.; Lercher, J. A. Activity and Selectivity Control in Reductive Amination of Butyraldehyde over Noble Metal Catalysts. Catal. Lett. 2005, 104, 23−28. (130) Kagan, F.; Rebenstorf, M. A.; Heinzelman, R. V. Preparation of glycamines. J. Am. Chem. Soc. 1957, 79, 3541−3544. (131) Lauritsen, A.; Madsen, R. Synthesis of naturally occurring iminosugars from D-fructose by the use of a zinc-mediated fragmentation reaction. Org. Biomol. Chem. 2006, 4, 2898−2905. (132) Sharma, S. K.; Songster, M. F.; Colpitts, T. L.; Hegyes, P.; Barany, G.; Castellino, F. J. Reductive amination with tritylamine as an ammonia equivalent: efficient preparation of the 5-(4-(9fluorenylmethyloxycarbonyl)aminomethyl-3,5-dimethoxyphenoxy)valeric acid (PAL) handle for peptide synthesis. J. Org. Chem. 1993, 58, 4993−4996.

(133) Borch, R. F.; Bernstein, M. D.; Durst, H. D. Cyanohydridoborate anion as a selective reducing agent. J. Am. Chem. Soc. 1971, 93, 2897−2904. (134) Miriyala, B.; Bhattacharyya, S.; Williamson, J. S. Chemoselective reductive alkylation of ammonia with carbonyl compounds: synthesis of primary and symmetrical secondary amines. Tetrahedron 2004, 60, 1463−1471. (135) Dangerfield, E. M.; Timmer, M. S. M.; Stocker, B. L. Total Synthesis Without Protecting Groups: Pyrrolidines and Cyclic Carbamates. Org. Lett. 2009, 11, 535−538. (136) Dangerfield, E. M.; Plunkett, C. H.; Win-Mason, A. L.; Stocker, B. L.; Timmer, M. S. M. Protecting-group-free synthesis of amines: synthesis of primary amines from aldehydes via reductive amination. J. Org. Chem. 2010, 75, 5470−5477. (137) Gawley, R. E. The Beckmann Reactions: Rearrangements, Elimination−Additions, Fragmentations, and Rearrangement−Cyclizations. Organic Reactions; John Wiley & Sons: New York, 2004. (138) Ayedi, M. A.; Le Bigot, Y.; Ammar, H.; Abid, S.; Gharbi, R. E.; Delmas, M. Synthesis of Primary Amines by One-Pot Reductive Amination of Aldehydes. Synth. Commun. 2013, 43, 2127−2133. (139) Jones, B. Kinetics and Mechanism of the Beckmann Rearrangement. Chem. Rev. 1944, 35, 335−350. (140) Yamabe, S.; Tsuchida, N.; Yamazaki, S. Is the Beckmann Rearrangement a Concerted or Stepwise Reaction? A Computational Study. J. Org. Chem. 2005, 70, 10638−10644. (141) Błachut, D.; Wojtasiewicz, K.; Krawczyk, K.; Maurin, J.; Szawkało, J.; Czarnocki, Z. Identification and synthesis of by-products found in 4-methylthioamphetamine (4-MTA) produced by the Leuckart method. Forensic Sci. Int. 2012, 216, 108−120. (142) Sommer, A.; Bruecker, R., Chemische Werke Huels, A.-G., Fed. Rep. Ger. 3-Aminomethyl-3,5,5-trimethylcyclohexylamine, DE3011656A1, 1981. (143) Disteldorf, J.; Huebel, W.; Broschinski, L. Chemische Werke Huels A.-G. Primary mono- and diamines from oxo compounds, DE3021955A1, 1981. (144) Merger, F.; Priester, C. U.; Witzel, T.; Koppenhoefer, G.; Harder, W. BASF A.-G., Germany. Process for preparing 3(aminomethyl)-3,5,5-trimethylcyclohexylamine, EP449089A1, 1991. (145) Gillet, J. P.; Kervennal, J.; Pralus, M. New process for isophoronediamine synthesis. Stud. Surf. Sci. Catal. 1993, 78, 321−328. (146) Carothers, W. H. du Pont, E. I. de Nemours & Co. Linear condensation “superpolymers” suitable for production of pliable, strong, elastic fibers, US2071250, 1937. (147) Chaudhari, P. S.; Salim, S. D.; Sawant, R. V.; Akamanchi, K. G. Sulfated tungstate: a new solid heterogeneous catalyst for amide synthesis. Green Chem. 2010, 12, 1707−1710. (148) Jouffret, F.; Madec, P. J. Polyamide synthesis by ester aminolysis. I. A new route for nylon-6,6 synthesis. J. Polym. Sci., Part A: Polym. Chem. 1996, 34, 2363−2370. (149) Han, C.; Porco. Synthesis of Carbamates and Ureas Using Zr(IV)-Catalyzed Exchange Processes. Org. Lett. 2007, 9, 1517−1520. (150) Culbertson, B. M. Cyclic imino ethers in step-growth polymerizations. Prog. Polym. Sci. 2002, 27, 579−626. (151) Verbicky, J. W., Jr.; Williams, L. Thermolysis of N-alkylsubstituted phthalamic acids. Steric inhibition of imide formation. J. Org. Chem. 1981, 46, 175−177. (152) Takahashi, Y.; Miyashi, T.; Yoon, U. C.; Oh, S. W.; Mancheno, M.; Su, Z.; Falvey, D. F.; Mariano, P. S. Mechanistic Studies of the Azomethine Ylide-Forming Photoreactions of N-(Silylmethyl)phthalimides and N-Phthaloylglycine. J. Am. Chem. Soc. 1999, 121, 3926−3932. (153) Adeppa, K.; Rupainwar, D. C.; Misra, K. A modified industrial procedure for large scale preparation of N-alkylimides and Narylimides in improved yields. Proc. Natl. Acad. Sci., India, Sect. B 2010, 80, 30−36. (154) Jin, S.; Yee, A. F. Preparation and characterization of maleimide-terminated poly(arylene ether sulfone) oligomers of various molecular weights. J. Appl. Polym. Sci. 1991, 43, 1849−1858. 14216

DOI: 10.1021/acs.chemrev.6b00486 Chem. Rev. 2016, 116, 14181−14224

Chemical Reviews

Review

maleimide/vinyl ether systems. Macromol. Chem. Phys. 2000, 201, 1493−1503. (175) Pozos-Vazquez, C. Ph.D. Thesis; University of Montpellier II: Montpellier, France, 2007. (176) Vazquez-Pozos, C.; Joly-Duhamel, C.; Boutevin, B.; Robin, J. J. Photopolymerization without photoinitiator of bismaleimide-containing oligo(oxypropylene)s: effect of oligoethers chain length. Macromol. Chem. Phys. 2009, 210, 269−278. (177) Gao, Y.; Zhang, Z.; Chen, L.; Gu, W.; Li, Y. Synthesis of 6N,N,N-Trimethyltriazole Chitosan via “Click Chemistry” and Evaluation for Gene Delivery. Biomacromolecules 2009, 10, 2175− 2182. (178) Wolfgang, H. B. Self-healing Polymers: From Principles to Applications; Wiley: New York, 2013. (179) Froidevaux, V.; Negrell, C.; Laborbe, E.; Auvergne, R.; Boutevin, B. Thermosetting material by a thermo responsive crosslinking using retroDiels-Alder and, in situ, Thia-Michael reactions. Eur. Polym. J. 2015, 69, 510−522. (180) Laborbe, E.; Le Rosignol, B.; Froidevaux, V.; Boutevin, B.; Auvergne, R. Cure-On-Demand Liquid Sealant Composition, Process For The Preparation Thereof And Uses Thereof, US20130137817A1, 2013. (181) Vazquez, C. P.; Joly-Duhamel, C.; Boutevin, B. PhotoinitiatorFree, Open-Air Acceptor/Donor Copolymerization of Bismaleimides: Simple Polymerization Conditions for New Thermoplastic Elastomer Production. Macromol. Chem. Phys. 2013, 214, 1621−1628. (182) Pozos-Vazquez, C.; Tayouo, R.; Joly-Duhamel, C.; Boutevin, B. UV-curable bismaleimides containing poly(dimethylsiloxane): Use as hydrophobic agent. J. Polym. Sci., Part A: Polym. Chem. 2010, 48, 2123−2134. (183) van Sonntag, J.; Knolle, W. Maleimides as electron-transfer photoinitiators: quantum yields of triplet states and radical-ion formation J. J. Photochem. Photobiol., A 2000, 136, 133−139. (184) McElhanon, J. R.; Zifer, T.; Kline, S. R.; Wheeler, D. R.; Loy, D. A.; Jamison, G. M.; Long, T. M.; Rahimian, K.; Simmons, B. A. Thermally Cleavable Surfactants Based on Furan Maleimide Diels Alder Adducts. Langmuir 2005, 21, 3259−3266. (185) Boros, M.; Koekoesi, J.; Vamos, J.; Koevesdi, I.; Noszal, B. Methods for syntheses of N-methyl-DL-aspartic acid derivatives. Amino Acids 2007, 33, 709−717. (186) Mather, B. D.; Viswanathan, K.; Miller, K. M.; Long, T. E. Michael addition reactions in macromolecular design for emerging technologies. Prog. Polym. Sci. 2006, 31, 487−531. (187) Boyer, C.; Loubat, C.; Robin, J. J.; Boutevin, B. Synthesis of functionalized amine oligomers by free-radical Telomerization of methyl methacrylate with a peculiar telogen: 2-aminoethanethiol hydrochloride. J. Polym. Sci., Part A: Polym. Chem. 2004, 42, 5146− 5160. (188) Boyer, C.; Boutevin, G.; Robin, J. J.; Boutevin, B. Study of the telomerization of dimethylaminoethyl methacrylate (DMAEMA) with mercaptoethanol. Application to the synthesis of a new macromonomer. Polymer 2004, 45, 7863−7876. (189) Long, T. E.; Broske, A. D.; Bradley, D. J.; McGrath, J. E. Synthesis and characterization of poly(t-butyl methacrylate-b-isopreneb-t-butyl methacrylate) block copolymers by anionic techniques. J. Polym. Sci., Part A: Polym. Chem. 1989, 27, 4001−4012. (190) Sun, G.; Wu, D.; Liu, Y.; He, C.; Chung, T. S.; Goh, S. H. pHControllable cyclic threading/dethreading of polypseudorotaxane obtained from cyclodextrins and poly(amino ester). Polymer 2005, 46, 3355−3362. (191) Elbert, D. L.; Hubbell, J. A. Conjugate Addition Reactions Combined with Free-Radical Cross-Linking for the Design of Materials for Tissue Engineering. Biomacromolecules 2001, 2, 430−441. (192) Wu, C. S.; Liu, Y. L.; Chiu, Y.-S. Synthesis and characterization of new organosoluble polyaspartimides containing phosphorus. Polymer 2002, 43, 1773−1779. (193) Dix, L. R.; Ebdon, J. R.; Hodge, P. Chain extension and crosslinking of telechelic oligomersII. Michael additions of bisthiols

(155) Endrey, A. L. du Pont, E. I. de Nemours & Co. Aromatic polyimide particles from polycyclic diamines, US3179631, 1965. (156) Altıntaş, Z.; Karataş, S.; Kayaman-Apohan, N.; Güngör, A. The maleimide modified epoxy resins for the preparation of UV-curable hybrid coatings. Polym. Adv. Technol. 2011, 22, 270−278. (157) Auvergne, R.; Morel, M.-H.; Menut, P.; Guilbert, S.; Robin, J.-J. The impact of bifunctional molecules on the gluten network during mixing. React. Funct. Polym. 2011, 71, 70−79. (158) Harwood, H. J.; Pyriadi, T. M. Use of acetyl chloridetriethylamine and acetic anhydride-triethylamine mixtures in the synthesis of isomaleimides from maleamic acids. J. Org. Chem. 1971, 36, 821−823. (159) Alagic, A.; Koprianiuk, A.; Kluger, R. Hemoglobin−Superoxide DismutaseChemical Linkages That Create a Dual-Function Protein. J. Am. Chem. Soc. 2005, 127, 8036−8043. (160) Syrett, J. A.; Mantovani, G.; Barton, W. R. S.; Price, D.; Haddleton, D. M. Self-healing polymers prepared via living radical polymerisation. Polym. Chem. 2010, 1, 102−106. (161) Heath, W. H.; Palmieri, F.; Adams, J. R.; Long, B. K.; Chute, J.; Holcombe, T. W.; Zieren, S.; Truitt, M. J.; White, J. L.; Willson, C. G. Degradable Cross-Linkers and Strippable Imaging Materials for Stepand-Flash Imprint Lithography. Macromolecules 2008, 41, 719−726. (162) Mizawa, T.; Takenaka, K.; Shiomi, T. Synthesis of αmaleimide-ω-dienyl heterotelechelic poly(methyl methacrylate) and its cyclization by the intramolecular Diels−Alder reaction. J. Polym. Sci., Part A: Polym. Chem. 2000, 38, 237−246. (163) Goussé, C.; Gandini, A. Diels−Alder polymerization of difurans with bismaleimides. Polym. Int. 1999, 48, 723−731. (164) White, J. E.; Scaia, M. D.; Snider, D. A. Reactions of diaminoalkanes with bismaleimides: Synthesis of some unusual polyimides. J. Appl. Polym. Sci. 1984, 29, 891−899. (165) Dolci, E.; Froidevaux, V.; Joly-Duhamel, C.; Auvergne, R.; Boutevin, B.; Caillol, S. Maleimides as a bulding block for the synthesis of high performance polymers. Polym. Rev. 2016, 56, 512−556. (166) Hagiwara, T.; Suzuki, I.; Takeuchi, K.; Hamana, H.; Narita, T. Synthesis and polymerization of N-(4-vinylphenyl)maleimide. Macromolecules 1991, 24, 6856−6858. (167) Hagiwara, T.; Shimizu, T.; Someno, T.; Yamagishi, T.; Hamana, H.; Narita, T. Anionic polymerization of N-substituted maleimide. 4. ″Living″ characteristics of anionic polymerization of Nphenylmaleimide. Macromolecules 1988, 21, 3324−3327. (168) Hagiwara, T.; Mizota, J.; Hamana, H.; Narita, T. Anionic polymerization of N-substituted maleimide, 1. Polymerization of Nphenylmaleimide. Makromol. Chem., Rapid Commun. 1985, 6, 169− 174. (169) Hagiwara, T.; Shimizu, T.; Hamana, H.; Narita, T. Anionic polymerization of N-substituted maleimide. VI. Effects of counter cation on the anionic polymerization of N-phenylmaleimide with alkali metal tert-butoxides. J. Polym. Sci., Part A: Polym. Chem. 1990, 28, 2437−2444. (170) Hagiwara, T.; Shimizu, T.; Uda, T.; Hamana, H.; Narita, T. Anionic polymerization of N-substituted maleimide. V.A study on the kinetic features of anionic polymerization of N-phenylmaleimide. J. Polym. Sci., Part A: Polym. Chem. 1990, 28, 185−192. (171) Hagiwara, T.; Someno, T.; Hamana, H.; Narita, T. Anionic polymerization of N-substitutes maleimide. II. Polymerization of Nethylmaleimide. J. Polym. Sci., Part A: Polym. Chem. 1988, 26, 1011− 1020. (172) Zhang, X.; Chen, G.-C.; Collins, A.; Jacobson, S.; Morganelli, P.; Dar, Y. L.; Musa, O. M. Thermally degradable maleimides for reworkable adhesives. J. Polym. Sci., Part A: Polym. Chem. 2009, 47, 1073−1084. (173) Morel, F.; Decker, C.; Jonsson, S.; Clark, S. C.; Hoyle, C. E. Kinetic study of the photo-induced copolymerization of N-substituted maleimides with electron donor monomers. Polymer 1999, 40, 2447− 2454. (174) Decker, C.; Bianchi, C.; Morel, F.; Jonsson, S.; Hoyle, C. Mechanistic study of the light-induced copolymerization of 14217

DOI: 10.1021/acs.chemrev.6b00486 Chem. Rev. 2016, 116, 14181−14224

Chemical Reviews

Review

(215) Diakoumakos, C. D.; Kotzev, D. L. Non-isocyanate-based polyurethanes derived upon the reaction of amines with cyclocarbonate resins. Macromol. Symp. 2004, 216, 37−46. (216) Helou, M.; Guillaume, S.; Carpentier, J.-F.; Miserque, O. Total Petrochemicals Research Feluy, Belgium, Centre National de la Recherche Scientifique. Catalytic process for polymerizing cyclic carbonates issued from renewable resources, WO2010012562A1, 2010. (217) Ochiai, B.; Inoue, S.; Endo, T. Salt effect on polyaddition of bifunctional cyclic carbonate and diamine. J. Polym. Sci., Part A: Polym. Chem. 2005, 43, 6282−6286. (218) Lambeth, R. H.; Henderson, T. J. Organocatalytic synthesis of (poly)hydroxyurethanes from cyclic carbonates and amines. Polymer 2013, 54, 5568−5573. (219) Clements, J. H. Reactive Applications of Cyclic Alkylene Carbonates. Ind. Eng. Chem. Res. 2003, 42, 663−674. (220) Deepa, P.; Jayakannan, M. Solvent-free and nonisocyanate melt transurethane reaction for aliphatic polyurethanes and mechanistic aspects. J. Polym. Sci., Part A: Polym. Chem. 2008, 46, 2445−2458. (221) Deepa, P.; Jayakannan, M. Solvent-induced self-organization approach for polymeric architectures of micropores, hexagons and spheres based on polyurethanes prepared via novel melt transurethane methodology. J. Polym. Sci., Part A: Polym. Chem. 2007, 45, 2351− 2366. (222) Li, S.; Zhao, J.; Zhang, Z.; Zhang, J.; Yang, W. Synthesis and characterization of aliphatic thermoplastic poly(ether urethane) elastomers through a non-isocyanate route. Polymer 2015, 57, 164− 172. (223) Mukherjee, S.; Yang, J. W.; Hoffmann, S.; List, B. Asymmetric Enamine Catalysis. Chem. Rev. (Washington, DC, U. S.) 2007, 107, 5471−5569. (224) Cordova, A. The Direct Catalytic Asymmetric Mannich Reaction. Acc. Chem. Res. 2004, 37, 102−112. (225) Shukla, V.; Bajpai, M.; Habib, F. Mannich base: a cost effective, low temperature curing agent for epoxides. Pigm. Resin Technol. 2007, 36, 169−171. (226) Baqar, M.; Agag, T.; Ishida, H.; Qutubuddin, S. Methylolfunctional benzoxazines as precursors for high-performance thermoset polymers: Unique simultaneous addition and condensation polymerization behavior. J. Polym. Sci., Part A: Polym. Chem. 2012, 50, 2275− 2285. (227) Ghosh, N. N.; Kiskan, B.; Yagci, Y. PolybenzoxazinesNew high performance thermosetting resins: Synthesis and properties. Prog. Polym. Sci. 2007, 32, 1344−1391. (228) Lehn, J.-M.; Skene, W. G. Fr. Polyhydrazones and polyimines exhibiting reversible formation and component exchange, WO2004003044A2, 2004. (229) Pickard, P. L.; Tolbert, T. L. Improved method of ketimine synthesis. J. Org. Chem. 1961, 26, 4886−4887. (230) Yamashkin, S. A.; Kucherenko, N. Y.; Yurovskaya, M. A. Reactions of acetoacetic ester with aryl- and heteroarylamines. Chem. Heterocycl. Compd. (N. Y., NY, U. S.) 1997, 33, 499−514. (231) Leonard, D. P.; Truesdale, J. H.; Scherrer, J. H.; Wright, H. J. Cook Paint and Varnish Co., U.S.A. Crosslinked acetoacetic ester polymers, EP227454A2, 1987. (232) Leitao, E. M.; Jurca, T.; Manners, I. Catalysis in service of main group chemistry offers a versatile approach to p-block molecules and materials. Nat. Chem. 2013, 5, 817−829. (233) Bellini, C.; Carpentier, J.-F.; Tobisch, S.; Sarazin, Y. BariumMediated Cross-Dehydrocoupling of Hydrosilanes with Amines: A Theoretical and Experimental Approach. Angew. Chem., Int. Ed. 2015, 54, 7679−7683. (234) Jolles, P., Muzzarelli, R. A. A., Eds. Chitin and Chitinases; Birkhaeuser Verlag: Basel, 1999. (235) Riva, R.; Ragelle, H.; des Rieux, A.; Duhem, N.; Jerome, C.; Preat, V. Chitosan and chitosan derivatives in drug delivery and tissue engineering. Adv. Polym. Sci. 2011, 244, 19−44. (236) Khor, E.; Lim, L. Y. Implantable applications of chitin and chitosan. Biomaterials 2003, 24, 2339−2349.

to bismaleimides, bismaleates and bis(acetylene ketone)s to give linear and crosslinked polymers. Eur. Polym. J. 1995, 31, 653−658. (194) Billiet, S.; Van Camp, W.; Hillewaere, X. K. D.; Rahier, H.; Du Prez, F. E. Development of optimized autonomous self-healing systems for epoxy materials based on maleimide chemistry. Polymer 2012, 53, 2320−2326. (195) Gokmen, M. T.; Brassinne, J.; Prasath, R. A.; Du Prez, F. E. Revealing the nature of thio-click reactions on the solid phase. Chem. Commun. 2011, 47, 4652−4654. (196) Ali, M. A.; Tateyama, S.; Oka, Y.; Kaneko, D.; Okajima, M. K.; Kaneko, T. Syntheses of High-Performance Biopolyamides Derived from Itaconic Acid and Their Environmental Corrosion. Macromolecules (Washington, DC, U. S.) 2013, 46, 3719−3725. (197) White, J. E. Synthesis and properties of high-molecular-weight step-growth polymers from bismaleimides. Ind. Eng. Chem. Prod. Res. Dev. 1986, 25, 395−400. (198) Crivello, J. V. Polyaspartimides: Condensation of aromatic diamines and bismaleimide compounds. J. Polym. Sci., Polym. Chem. Ed. 1973, 11, 1185−1200. (199) Fullerton, R.; Roylance, D.; Allred, R.; Acton, A. Cure analysis of printed wiring boards containing reactive adhesive layers. Polym. Eng. Sci. 1988, 28, 372−376. (200) Pascault, J.-P.; Sautereau, H.; Verdu, J.; Williams, R. J. J. Thermosetting Polymers; Marcel Dekker, Inc.: New York, 2002. (201) Delebecq, E.; Pascault, J.-P.; Boutevin, B.; Ganachaud, F. On the Versatility of Urethane/Urea Bonds: Reversibility, Blocked Isocyanate, and Non-isocyanate Polyurethane. Chem. Rev. 2013, 113, 80−118. (202) Bayer, O. The diisocyanate polyaddition process (polyurethanes). Description of a new principle for building up highmolecular compounds (1937−1945). Angew. Chem. 1947, 59, 257− 272. (203) Maisonneuve, L.; Lamarzelle, O.; Rix, E.; Grau, E.; Cramail, H. Isocyanate-Free Routes to Polyurethanes and Poly(hydroxy Urethane)s. Chem. Rev. 2015, 115, 12407−12439. (204) Kihara, N.; Endo, T. Synthesis and properties of poly(hydroxyurethane)s. J. Polym. Sci., Part A: Polym. Chem. 1993, 31, 2765−2773. (205) Rokicki, G.; Lewandowski, M. Epoxy resins modified by carbon dioxide. Angew. Makromol. Chem. 1987, 148, 53−66. (206) Whelan, J. M., Jr.; Cotter, R. J. Union Carbide Corp. Cyclic carbonate polymers, US3072613, 1963. (207) Mikheev, V. V.; Svetlakov, N. V.; Sysoev, V. A.; Gumerova, R. K. Reaction of organic carbonates with amines. III. Reaction of amines with alkylene carbonates. Zh. Org. Khim. 1983, 19, 498−501. (208) Kihara, N.; Kushida, Y.; Endo, T. Optically active poly(hydroxyurethane)s derived from cyclic carbonate and L-lysine derivatives. J. Polym. Sci., Part A: Polym. Chem. 1996, 34, 2173−2179. (209) Tomita, H.; Sanda, F.; Endo, T. Structural analysis of polyhydroxyurethane obtained by polyaddition of bifunctional fivemembered cyclic carbonate and diamine based on the model reaction. J. Polym. Sci., Part A: Polym. Chem. 2001, 39, 851−859. (210) Tomita, H.; Sanda, F.; Endo, T. Polyaddition behavior of bis(five- and six-membered cyclic carbonate)s with diamine. J. Polym. Sci., Part A: Polym. Chem. 2001, 39, 860−867. (211) Tomita, H.; Sanda, F.; Endo, T. Model reaction for the synthesis of polyhydroxyurethanes from cyclic carbonates with amines: Substituent effect on the reactivity and selectivity of ring-opening direction in the reaction of five-membered cyclic carbonates with amine. J. Polym. Sci., Part A: Polym. Chem. 2001, 39, 3678−3685. (212) Leykin, A.; Beilin, D.; Birukova, O.; Figovsky, O.; Shapovalov, L. Nonisocyanate polyurethanes based on cyclic carbonates: part 1. Chemistry. Sci. Isr. Technol. Advantages 2009, 11, 160−176. (213) Figovsky, O. L.; Shapovalov, L. D. Features of reaction aminocyclocarbonate for production of new type nonisocyanate polyurethane coatings. Macromol. Symp. 2002, 187, 325−332. (214) Kreye, O.; Mutlu, H.; Meier, M. A. R. Sustainable routes to polyurethane precursors. Green Chem. 2013, 15, 1431−1455. 14218

DOI: 10.1021/acs.chemrev.6b00486 Chem. Rev. 2016, 116, 14181−14224

Chemical Reviews

Review

precursors, block copolymers and graft copolymers. RSC Polym. Chem. Ser. 2013, 6, 217−235. (259) Steunenberg, P.; Koenst, P. M.; Scott, E. L.; Franssen, M. C. R.; Zuilhof, H.; Sanders, J. P. M. Polymerisation of β-alanine through catalytic ester-amide exchange. Eur. Polym. J. 2013, 49, 1773−1781. (260) Kind, S.; Wittmann, C. Bio-based production of the platform chemical 1,5-diaminopentane. Appl. Microbiol. Biotechnol. 2011, 91, 1287−1296. (261) Adkins, J.; Pugh, S.; McKenna, R.; Nielsen, D. R. Engineering microbial chemical factories to produce renewable ″biomonomers″. Front. Microbiol. 2012, 3, 313. (262) Wicks, D. A. Draths Corporation, U.S.A. Preparation of cyclic amide monomers for polyamides, WO2010011967A1, 2010. (263) Frost, J. W. Board of Trustees of Michigan State University, U.S.A. Synthesis of caprolactam from lysine, WO2005123669A1, 2005. (264) Pukin, A. V.; Boeriu, C. G.; Scott, E. L.; Sanders, J. P. M.; Franssen, M. C. R. An efficient enzymatic synthesis of 5-aminovaleric acid. J. Mol. Catal. B: Enzym. 2010, 65, 58−62. (265) Storey, R. F.; Wiggins, J. S.; Mauritz, K. A.; Puckett, A. D. Bioabsorbable composites. II: Nontoxic, L-lysine-based poly(esterurethane) matrix composites. Polym. Compos. 1993, 14, 17−25. (266) Burgard, A. P.; Pharkya, P.; Osterhout, R. E. Genomatica, Inc. Metabolic pathway engineering of microorganisms for the production of adipic acid and other compounds, US20090305364A1, 2009. (267) Collet, H.; Souaid, E.; Cottet, H.; Deratani, A.; Boiteau, L.; Dessalces, G.; Rossi, J.-C.; Commeyras, A.; Pascal, R. An Expeditious Multigram-Scale Synthesis of Lysine Dendrigraft (DGL) Polymers by Aqueous N-Carboxyanhydride Polycondensation. Chem. - Eur. J. 2010, 16, 2309−2316. (268) Scholl, M.; Nguyen, T. Q.; Bruchmann, B.; Klok, H.-A. Controlling polymer architecture in the thermal hyperbranched polymerization of L-lysine. Macromolecules 2007, 40, 5726−5734. (269) Scholl, M.; Nguyen, T. Q.; Bruchmann, B.; Klok, H.-A. The thermal polymerization of amino acids revisited; synthesis and structural characterization of hyperbranched polymers from L-lysine. J. Polym. Sci., Part A: Polym. Chem. 2007, 45, 5494−5508. (270) Lammens, T. M.; De Biase, D.; Franssen, M. C. R.; Scott, E. L.; Sanders, J. P. M. The application of glutamic acid [small alpha]decarboxylase for the valorization of glutamic acid. Green Chem. 2009, 11, 1562−1567. (271) Biermann, U.; Furmeier, S.; Metzger, J. O. Oleochemical Manufacture and Applications. New Chem. Oils Fats 2001, 9, 266−299. (272) McDonald, R. N.; Gruger, E. H., Jr. New products from fish oils. II. Polyamino fatty acids derived from fish oils. Comm. Fisheries Rev. 1956, 18, 7−11. (273) Kharasch, M. S.; Mayo, F. R. Peroxide effect in the addition of reagents to unsaturated compounds. I. The addition of hydrogen bromide to allyl bromide. J. Am. Chem. Soc. 1933, 55, 2468−2490. (274) Genas, M. Organico. Amino acids, FR988699, 1951. (275) Pees, B.; Lebrun, S., Arkema France, Fr. Process for the ammonolysis of 11-bromoundecanoic acid in preparation of 11aminoundecanoic acid, FR2940282A1, 2010. (276) http://www.rsc.org/chemistryworld/2013/08/palm-oilpolyamide-evonik; accessed 01/15/2016. (277) http://corporate.evonik.com/en/media/search/pages/newsdetails.aspx?newsid=40329; accessed 01/15/2016. (278) Schmid, A.; Buehler, B.; Ladkau, N.; Hermann, I. Evonik Degussa GmbH, Germany. Enzymic ω-aminocarboxylic acid condensation, WO2012156151A1, 2012. (279) Ayorinde, F. O.; Nana, E. Y.; Nicely, P. D.; Woods, A. S.; Price, E. O.; Nwaonicha, C. P. Syntheses of 12-aminododecanoic and 11aminoundecanoic acids from vernolic acid. J. Am. Oil Chem. Soc. 1997, 74, 531−538. (280) Spiccia, N. D.; Border, E.; Illesinghe, J.; Jackson, W. R.; Robinson, A. J. Preparation of a nylon-11 precursor from renewable canola oil. Synthesis 2013, 45, 1683−1688. (281) Kamiyama, T.; Inoue, M.; Enomoto, S. Amination of fatty acids using copper(II)-ammine complex. Chem. Lett. 1975, 4, 165−166.

(237) Hu, Y.; Liu, Y.; Qi, X.; Liu, P.; Fan, Z.; Li, S. Novel bioresorbable hydrogels prepared from chitosan-graft-polylactide copolymers. Polym. Int. 2012, 61, 74−81. (238) Rinaudo, M. Chitin and chitosan: Properties and applications. Prog. Polym. Sci. 2006, 31, 603−632. (239) Kong, M.; Chen, X. G.; Xing, K.; Park, H. J. Antimicrobial properties of chitosan and mode of action: A state of the art review. Int. J. Food Microbiol. 2010, 144, 51−63. (240) Mourya, V. K.; Inamdar, N. N.; Choudhari, Y. M. Chitooligosaccharides: Synthesis, characterization and applications. Polym. Sci., Ser. A 2011, 53, 583−612. (241) Illy, N.; Robitzer, M.; Auvergne, R.; Caillol, S.; David, G.; Boutevin, B. Synthesis of water-soluble allyl-functionalized oligochitosan and its modification by thiol-ene addition in water. J. Polym. Sci., Part A: Polym. Chem. 2014, 52, 39−48. (242) Fangkangwanwong, J.; Akashi, M.; Kida, T.; Chirachanchai, S. Chitosan-hydroxybenzotriazole aqueous solution: a novel water-based system for chitosan functionalization. Macromol. Rapid Commun. 2006, 27, 1039−1046. (243) Tishchenko, G.; Simunek, J.; Brus, J.; Netopilik, M.; Pekarek, M.; Walterova, Z.; Koppova, I.; Lenfeld, J. Low-molecular-weight chitosans: Preparation and characterization. Carbohydr. Polym. 2011, 86, 1077−1081. (244) Liu, L.; Li, Y.; Fang, Y.-e.; Chen, L. Microwave-assisted graft copolymerization of ε-caprolactone onto chitosan via the phthaloyl protection method. Carbohydr. Polym. 2005, 60, 351−356. (245) Kurita, K.; Ikeda, H.; Yoshida, Y.; Shimojoh, M.; Harata, M. Chemoselective Protection of the Amino Groups of Chitosan by Controlled Phthaloylation: Facile Preparation of a Precursor Useful for Chemical Modifications. Biomacromolecules 2002, 3, 1−4. (246) Hua, D.; Tang, J.; Cheng, J.; Deng, W.; Zhu, X. A novel method of controlled grafting modification of chitosan via RAFT polymerization using chitosan-RAFT agent. Carbohydr. Polym. 2008, 73, 98−104. (247) Kim, K.; Ryu, J. H.; Lee, D. Y.; Lee, H. Bio-inspired catechol conjugation converts water-insoluble chitosan into a highly watersoluble, adhesive chitosan derivative for hydrogels and LbL assembly. Biomater. Sci. 2013, 1, 783−790. (248) Mohy Eldin, M. S.; Soliman, E. A.; Hashem, A. I.; Tamer, T. M. Antimicrobial activity of novel aminated chitosan derivatives for biomedical applications. Adv. Polym. Technol. 2012, 31, 414−428. (249) Medici, R.; Dominguez de Maria, P.; Otten, L. G.; Straathof, A. J. J. A high-throughput screening assay for amino acid decarboxylase activity. Adv. Synth. Catal. 2011, 353, 2369−2376. (250) Foti, M.; Medici, R.; Ruijssenaars, H. J. Biological production of monoethanolamine by engineered Pseudomonas putida S12. J. Biotechnol. 2013, 167, 344−349. (251) Straathof, A. J. J. Transformation of Biomass into Commodity Chemicals Using Enzymes or Cells. Chem. Rev. 2014, 114, 1871−1908. (252) Fache, M.; Darroman, E.; Besse, V.; Auvergne, R.; Caillol, S.; Boutevin, B. Vanillin, a promising biobased building-block for monomer synthesis. Green Chem. 2014, 16, 1987−1998. (253) Silverstein, J. S.; Casey, B. J.; Natoli, M. E.; Dair, B. J.; Kofinas, P. Rapid Modular Synthesis and Processing of Thiol-Ene Functionalized Styrene-Butadiene Block Copolymers. Macromolecules 2012, 45, 3161−3167. (254) Auvergne, R.; Desroches, M.; Clerc, S.; Carlotti, S.; Caillol, S.; Boutevin, B. New biobased epoxy hardeners: Thiol-ene addition on oligobutadiene. React. Funct. Polym. 2012, 72, 393−401. (255) Stanford, M. J.; Pflughaupt, R. L.; Dove, A. P. Synthesis of Stereoregular Cyclic Poly(lactide)s via ″Thiol-Ene″ Click Chemistry. Macromolecules 2010, 43, 6538−6541. (256) Lowe, A. B. Thiol-ene ″click″ reactions and recent applications in polymer and materials synthesis: a first update. Polym. Chem. 2014, 5, 4820−4870. (257) Hoyle, C. E.; Bowman, C. N. Thiol-Ene Click Chemistry. Angew. Chem., Int. Ed. 2010, 49, 1540−1573. (258) Boutevin, B.; Auvergne, R.; David, G. Thiol-ene radical coupling: a powerful technique for the synthesis of polymer 14219

DOI: 10.1021/acs.chemrev.6b00486 Chem. Rev. 2016, 116, 14181−14224

Chemical Reviews

Review

(282) Schaffer, S.; Hauberg, M.; Wessel, M.; Hennemann, H.-G.; Pfeffer, J.-C.; Haas, T.; Haeger, H. Evonik Industries AG, Germany. Enzymatic ω-oxidation and -amination of fatty acids and their esters using whole cell biocatalysts, EP2639308A1, 2013. (283) Schrewe, M.; Ladkau, N.; Buehler, B.; Schmid, A. Direct Terminal Alkylamino-Functionalization via Multistep Biocatalysis in One Recombinant Whole-Cell Catalyst. Adv. Synth. Catal. 2013, 355, 1693−1697. (284) Song, J.-W.; Lee, J.-H.; Bornscheuer, U. T.; Park, J.-B. Microbial Synthesis of Medium-Chain α,ω-Dicarboxylic Acids and ωAminocarboxylic Acids from Renewable Long-Chain Fatty Acids. Adv. Synth. Catal. 2014, 356, 1782−1788. (285) Mekki-Berrada, A.; Bennici, S.; Gillet, J.-P.; Couturier, J.-L.; Dubois, J.-L.; Auroux, A. Ammoniation-Dehydration of Fatty Acids into Nitriles: Heterogeneous or Homogeneous Catalysis? ChemSusChem 2013, 6, 1478−1489. (286) Mekki-Berrada, A.; Bennici, S.; Gillet, J. P.; Couturier, J. L.; Dubois, J. L.; Auroux, A. Fatty acid methyl esters into nitriles: Acidbase properties for enhanced catalysts. J. Catal. 2013, 306, 30−37. (287) http://www.crodacoatingsandpolymers.com/home.aspx?s= 139&r=669&p=3325; accessed 01/21/2016. (288) Biermann, U.; Friedt, W.; Lang, S.; Lühs, W.; Machmüller, G.; Metzger, J. O.; Rüsch gen Klaas, M.; Schäfer, H. J.; Schneider, M. P. New Syntheses with Oils and Fats as Renewable Raw Materials for the Chemical Industry. Angew. Chem., Int. Ed. 2000, 39, 2206−2224. (289) Baumann, H.; Bühler, M.; Fochem, H.; Hirsinger, F.; Zoebelein, H.; Falbe, J. Natural Fats and OilsRenewable Raw Materials for the Chemical Industry. Angew. Chem., Int. Ed. Engl. 1988, 27, 41−62. (290) Biswas, A.; Sharma, B. K.; Doll, K. M.; Erhan, S. Z.; Willett, J. L.; Cheng, H. N. Synthesis of an Amine−Oleate Derivative Using an Ionic Liquid Catalyst. J. Agric. Food Chem. 2009, 57, 8136−8141. (291) Zhao, H.-P.; Zhang, J.-F.; Susan Sun, X.; Hua, D. H. Syntheses and properties of cross-linked polymers from functionalized triglycerides. J. Appl. Polym. Sci. 2008, 110, 647−656. (292) Earls, J. D.; White, J. E.; López, L. C.; Lysenko, Z.; Dettloff, M. L.; Null, M. J. Amine-cured ω-epoxy fatty acid triglycerides: Fundamental structure−property relationships. Polymer 2007, 48, 712−719. (293) Hoyle, C. E.; Lee, T. Y.; Roper, T. Thiol−enes: Chemistry of the past with promise for the future. J. Polym. Sci., Part A: Polym. Chem. 2004, 42, 5301−5338. (294) Stemmelen, M.; Pessel, F.; Lapinte, V.; Caillol, S.; Habas, J. P.; Robin, J. J. A fully biobased epoxy resin from vegetable oils: From the synthesis of the precursors by thiol-ene reaction to the study of the final material. J. Polym. Sci., Part A: Polym. Chem. 2011, 49, 2434− 2444. (295) Turunc, O.; Firdaus, M.; Klein, G.; Meier, M. A. R. Fatty acid derived renewable polyamides via thiol-ene additions. Green Chem. 2012, 14, 2577−2583. (296) Jeol, S.; Decampo, F.; Li, P. Rhodia Operations, Solvay China Co., Ltd., France. Tetrahydrofuran-containing polyamide, preparation methods and applications, WO2014106485A1, 2014. (297) Cok, B.; Tsiropoulos, I.; Roes, A. L.; Patel, M. K. Succinic acid production derived from carbohydrates: An energy and greenhouse gas assessment of a platform chemical toward a bio-based economy. Biofuels, Bioprod. Biorefin. 2014, 8, 16−29. (298) Voit, G.; Fischer, R.; Bassler, P.; Ansmann, A.; Luyken, H.; Merger, M.; Ohlbach, F.; Rehfinger, A., BASF A.-G., Germany. Hydrogenation process and catalysts for producing aliphatic α,ωdiamines from the corresponding dinitriles, WO9944983A1, 1999. (299) Mares, F.; Galle, J. E.; Diamond, S. E.; Regina, F. J. Preparation and characterization of a novel catalyst for the hydrogenation of dinitriles to amino nitriles. J. Catal. 1988, 112, 145−156. (300) Ritz, J.; Fuchs, E.; Voit, G.; Achhammer, G.; Fischer, R. BASF A.-G., Germany. Manufacture of caprolactam by cyclization of 6aminocapronitrile, US5693793A, 1997.

(301) Frost, J. W.; Draths, K. M. Purdue Research Foundation, U.S.A. Synthesis of adipic acid from biomass-derived carbon sources, WO9507996A1, 1995. (302) Pavone, A. Bio-Based adipic acid Process Economics Program 2012, Report No. 284. (303) Boussie, T. R.; Dias, E. L.; Fresco, Z. M.; Murphy, V. J.; Shoemaker, J.; Archer, R.; Jiang, H. Rennovia, Inc. Production of adipic acid and derivatives from carbohydrate-containing materials, WO2010144862A2, 2010. (304) Elias, H. G. Synthesis, Materials, and Technology. Macromolecules; Pub, K. A.; 2013; Vol. 2, Proteins, pp 1015−1051. (305) Pera-Titus, M.; Shi, F. Catalytic Amination of Biomass-Based Alcohols. ChemSusChem 2014, 7, 720−722. (306) Pingen, D.; Diebolt, O.; Vogt, D. Direct Amination of BioAlcohols Using Ammonia. ChemCatChem 2013, 5, 2905−2912. (307) Matsuoka, K.; Kohzu, T.; Hakumura, T.; Koyama, T.; Hatano, K.; Terunuma, D. Synthetic construction of a Lex determinant via Gabriel amine synthesis and the glycopolymer involving highly clustered Lex residues. Tetrahedron Lett. 2009, 50, 2593−2596. (308) http://www.arpae-summit.com/paperclip/exhibitor_docs/ 14AE/LanzaTech_Inc._131.pdf; accessed 01/21/2016. (309) Sieber, V.; Grammann, K.; Ruehmann, B.; Haas, T.; Pfeffer, J.; Doderer, K.; Rollmann, C.; Skerra, A.; Rausch, C.; Lerchner, A. Evonik Degussa GmbH, Germany. Enzymic transamination of multicyclic dianhydro diuloses, WO2010089171A2, 2010. (310) Montgomery, R.; Wiggins, L. F. Anhydrides of polyhydric alcohols. V. 2,5-Diamino-1,4,3,6-dianhydromannitol and -sorbitol and their sulfanilamide derivatives. J. Chem. Soc. 1946, 0, 393−396. (311) Bashford, V. G.; Wiggins, L. F. Anhydrides of polyhydric alcohols. XIII. The amino derivatives of 1, 4:3, 6-dianhydromannitol, -sorbitol, and L-iditol and their behavior towards nitrous acid. J. Chem. Soc. 1950, 0, 371−374. (312) Cope, A. C.; Shen, T. Y. The Stereochemistry of 1,4:3,6Dianhydrohexitol Derivatives. J. Am. Chem. Soc. 1956, 78, 3177−3182. (313) Thiyagarajan, S.; Gootjes, L.; Vogelzang, W.; van Haveren, J.; Lutz, M.; van Es, D. S. Renewable Rigid Diamines: Efficient, Stereospecific Synthesis of High Purity Isohexide Diamines. ChemSusChem 2011, 4, 1823−1829. (314) Imm, S.; Bähn, S.; Zhang, M.; Neubert, L.; Neumann, H.; Klasovsky, F.; Pfeffer, J.; Haas, T.; Beller, M. Improved RutheniumCatalyzed Amination of Alcohols with Ammonia: Synthesis of Diamines and Amino Esters. Angew. Chem., Int. Ed. 2011, 50, 7599− 7603. (315) Lerchner, A.; Achatz, S.; Rausch, C.; Haas, T.; Skerra, A. Coupled Enzymatic Alcohol-to-Amine Conversion of Isosorbide using Engineered Transaminases and Dehydrogenases. ChemCatChem 2013, 5, 3374−3383. (316) Feriani, A.; Gaviraghi, G.; Toson, G.; Mor, M.; Barbieri, A.; Grana, E.; Boselli, C.; Guarneri, M.; Simoni, D.; Manfredini, S. Cholinergic Agents Structurally Related to Furtrethonium. 2. Synthesis and Antimuscarinic Activity of a Series of N-[5-[(1′-Substitutedacetoxy)methyl]-2-furfuryl]dialkylamines. J. Med. Chem. 1994, 37, 4278−4287. (317) Müller, C.; Diehl, V.; Lichtenthaler, F. W. Building blocks from sugars. Part 23. Hydrophilic 3-pyridinols from fructose and isomaltulose. Tetrahedron 1998, 54, 10703−10712. (318) Cukalovic, A.; Stevens, C. V. Production of biobased HMF derivatives by reductive amination. Green Chem. 2010, 12, 1201−1206. (319) Xu, Z.; Yan, P.; Xu, W.; Jia, S.; Xia, Z.; Chung, B.; Zhang, Z. C. Direct reductive amination of 5-hydroxymethylfurfural with primary/ secondary amines via Ru-complex catalyzed hydrogenation. RSC Adv. 2014, 4, 59083−59087. (320) Holfinger, M. S.; Conner, A. H.; Hill, C. G., Jr. Kinetic Model for the Acid-Catalyzed Formation of Difurfuryldiamines from Furfurylamine and Aldehydes. Ind. Eng. Chem. Res. 1997, 36, 605−613. (321) Conner, A. H.; Hill, C. G., Jr.; McKillip, W. J.; Reimann, R. H. United States Dept. of Agriculture. One-step method for the preparation of difurfuryldiamines, US5292903A, 1994. 14220

DOI: 10.1021/acs.chemrev.6b00486 Chem. Rev. 2016, 116, 14181−14224

Chemical Reviews

Review

(322) Gandini, A.; Belgacem, M. N. Furans in polymer chemistry. Prog. Polym. Sci. 1997, 22, 1203−1379. (323) Moreau, C.; Belgacem, M. N.; Gandini, A. Recent Catalytic Advances in the Chemistry of Substituted Furans from Carbohydrates and in the Ensuing Polymers. Top. Catal. 2004, 27, 11−30. (324) Jeol, S. Rhodia Operations, France. Novel polyamide, process for preparing same and uses thereof, WO2013007585A1, 2013. (325) Le Tien, C.; Lacroix, M.; Ispas-Szabo, P.; Mateescu, M.-A. Nacylated chitosan: hydrophobic matrices for controlled drug release. J. Controlled Release 2003, 93, 1−13. (326) Keim, W.; Roeper, M. Terpene amine synthesis via palladiumcatalyzed isoprene telomerization with ammonia. J. Org. Chem. 1981, 46, 3702−3707. (327) Keim, W.; Kurtz, K. R.; Roeper, M. Palladium catalyzed telomerization of isoprene with secondary amines and conversion of the resulting terpene amines to terpenols. J. Mol. Catal. 1983, 20, 129− 138. (328) Lochynski, S.; Kuldo, J.; Frackowiak, B.; Holband, J.; Wojcik, G. Stereochemistry of terpene derivatives. Part 2: synthesis of new chiral amino acids with potential neuroactivity. Tetrahedron: Asymmetry 2000, 11, 1295−1302. (329) Gajcy, K.; Pekala, J.; Frackowiak-Wojtasek, B.; Librowski, T.; Lochynski, S. Stereochemistry of terpene derivatives. Part 7: Novel rigidified amino acids from (+)-3-carene designed as chiral GABA analogues. Tetrahedron: Asymmetry 2010, 21, 2015−2020. (330) Ferrarini, S. R.; Duarte, M. O.; da Rosa, R. G.; Rolim, V.; Eifler-Lima, V. L.; von Poser, G.; Ribeiro, V. L. S. Acaricidal activity of limonene, limonene oxide and β-amino alcohol derivatives on Rhipicephalus (Boophilus) microplus. Vet. Parasitol. 2008, 157, 149−153. (331) Ferrarini, S. R.; Graebin, C. S.; Limberger, J.; Canto, R. F. S.; Dias, D. O.; da Rosa, R. G.; Madeira, M. d. F.; Eifler-Lima, V. L. Synthesis of limonene β-amino alcohol derivatives in support of new antileishmanial therapies. Mem. Inst. Oswaldo Cruz 2008, 103, 773− 777. (332) Watts, C. C.; Thoniyot, P.; Cappuccio, F.; Verhagen, J.; Gallagher, B.; Singaram, B. Catalytic asymmetric transfer hydrogenation of ketones using terpene-based chiral β-amino alcohols. Tetrahedron: Asymmetry 2006, 17, 1301−1307. (333) Watts, C. C.; Thoniyot, P.; Hirayama, L. C.; Romano, T.; Singaram, B. Enantioselective alkynylations of aromatic and aliphatic aldehydes catalyzed by terpene derived chiral amino alcohols. Tetrahedron: Asymmetry 2005, 16, 1829−1835. (334) McKeever, C. H.; Washburne, R. N. Rohm & Haas Co. 1,8Diamino-p-menthane, US2955138, 1960. (335) Behr, A.; Wintzer, A.; Luebke, C.; Mueller, M. Synthesis of primary amines from the renewable compound citronellal via biphasic reductive amination. J. Mol. Catal. A: Chem. 2015, 404−405, 74−82. (336) Alves, M.-H.; Sfeir, H.; Tranchant, J.-F.; Gombart, E.; Sagorin, G.; Caillol, S.; Billon, L.; Save, M. Terpene and Dextran Renewable Resources for the Synthesis of Amphiphilic Biopolymers. Biomacromolecules 2014, 15, 242−251. (337) Woolard, F. X.; McPhee, D. J. Amyris, Inc., U.S.A. Surfactants from terpene hydrocarbons via Diels-Alder reactions, WO2012103156A1, 2012. (338) Cummings, L. O. Water-insoluble phenol-formaldehyde polyamines, curing of polyepoxides and the resulting product, US4491654A, 1985. (339) Thiyagarajan, S.; Gootjes, L.; Vogelzang, W.; Wu, J.; van Haveren, J.; van Es, D. Chiral building blocks from biomass: 2,5Diamino-2,5-dideoxy-1,4−3,6-dianhydroiditol. Tetrahedron 2011, 67, 383−389. (340) Zhang, M.; Zhang, J.; Chen, S.; Zhou, Y. Synthesis and fire properties of rigid polyurethane foams made from a polyol derived from melamine and cardanol. Polym. Degrad. Stab. 2014, 110, 27−34. (341) Calo, E.; Maffezzoli, A.; Mele, G.; Martina, F.; Mazzetto, S. E.; Tarzia, A.; Stifani, C. Synthesis of a novel cardanol-based benzoxazine monomer and environmentally sustainable production of polymers and bio-composites. Green Chem. 2007, 9, 754−759.

(342) Raquez, J. M.; Deleglise, M.; Lacrampe, M. F.; Krawczak, P. Thermosetting (bio)materials derived from renewable resources: A critical review. Prog. Polym. Sci. 2010, 35, 487−509. (343) Lochab, B.; Varma, I. K.; Bijwe, J. Thermal behaviour of cardanol-based benzoxazines. J. Therm. Anal. Calorim. 2010, 102, 769− 774. (344) Nalakathu Kolanadiyil, S.; Bijwe, J.; Varma, I. K. Synthesis of itaconimide/nadimide-functionalized benzoxazine monomers: Structural and thermal characterization. React. Funct. Polym. 2013, 73, 1544−1552. (345) Rao, B. S.; Palanisamy, A. Synthesis of bio based low temperature curable liquid epoxy, benzoxazine monomer system from cardanol: Thermal and viscoelastic properties. Eur. Polym. J. 2013, 49, 2365−2376. (346) Agag, T.; An, S. Y.; Ishida, H. 1,3-bis(benzoxazine) from cashew nut shell oil and diaminodiphenyl methane and its composites with wood flour. J. Appl. Polym. Sci. 2013, 127, 2710−2714. (347) Rao, B. S.; Palanisamy, A. A new thermo set system based on cardanol benzoxazine and hydroxy benzoxazoline with lower cure temperature. Prog. Org. Coat. 2012, 74, 427−434. (348) Minigher, A.; Benedetti, E.; De Giacomo, O.; Campaner, P.; Aroulmoji, V. Synthesis and characterization of novel cardanol based benzoxazines. Nat. Prod. Commun. 2009, 4, 521−528. (349) Kasemsiri, P.; Hiziroglu, S.; Rimdusit, S. Effect of cashew nut shell liquid on gelation, cure kinetics, and thermomechanical properties of benzoxazine resin. Thermochim. Acta 2011, 520, 84−92. (350) Darroman, E.; Bonnot, L.; Auvergne, R.; Boutevin, B.; Caillol, S. New aromatic amine based on cardanol giving new biobased epoxy networks with cardanol. Eur. J. Lipid Sci. Technol. 2015, 117, 178−189. (351) Mhaske, S. B.; Bhingarkar, R. V.; Sabne, M. B.; Mercier, R.; Vernekar, S. P. Synthesis and characterization of end-capped polyimides and their gas permeability properties. J. Appl. Polym. Sci. 2000, 77, 627−635. (352) Attanasi, O. A.; Berretta, S.; Fiani, C.; Filippone, P.; Mele, G.; Saladino, R. Synthesis and reactions of nitro derivatives of hydrogenated cardanol. Tetrahedron 2006, 62, 6113−6120. (353) Sadavarte, N. V.; Halhalli, M. R.; Avadhani, C. V.; Wadgaonkar, P. P. Synthesis and characterization of new polyimides containing pendent pentadecyl chains. Eur. Polym. J. 2009, 45, 582−589. (354) Jadhav, A. S.; Vernekar, S. P.; Maldar, N. N. Synthesis and characterization of new aromatic sulfone ether polyamides containing pendant pentadecyl groups. Polym. Int. 1993, 32, 5−11. (355) Laurichesse, S.; Averous, L. Chemical modification of lignins: Towards biobased polymers. Prog. Polym. Sci. 2014, 39, 1266−1290. (356) Lü, Q.-F.; Huang, Z.-K.; Liu, B.; Cheng, X. Preparation and heavy metal ions biosorption of graft copolymers from enzymatic hydrolysis lignin and amino acids. Bioresour. Technol. 2012, 104, 111− 118. (357) Wang, X.; Zhang, Y.; Hao, C.; Dai, X.; Zhou, Z.; Si, N. Ultrasonic-assisted synthesis of aminated lignin by a Mannich reaction and its decolorizing properties for anionic azo-dyes. RSC Adv. 2014, 4, 28156−28164. (358) Fache, M.; Boutevin, B.; Caillol, S. Vanillin, a key-intermediate of biobased polymers. Eur. Polym. J. 2015, 68, 488−502. (359) Gannett, P. M.; Nagel, D. L.; Reilly, P. J.; Lawson, T.; Sharpe, J.; Toth, B. Capsaicinoids: their separation, synthesis, and mutagenicity. J. Org. Chem. 1988, 53, 1064−1071. (360) Kaga, H.; Miura, M.; Orito, K. A facile procedure for synthesis of capsaicin. J. Org. Chem. 1989, 54, 3477−3478. (361) Kim, S. W.; Park, J. H.; Jung, S. J.; Choi, T. B.; Hur, M. G.; Yang, S. D.; Yu, K. H. Synthesis and elucidation of deuterated vanillylamine hydrochloride and capsaicin. J. Labelled Compd. Radiopharm. 2009, 52, 563−565. (362) Du, C. J.; Rios-Solis, L.; Ward, J. M.; Dalby, P. A.; Lye, G. J. Evaluation of CV2025 ω-transaminase for the bioconversion of lignin breakdown products into value-added chemicals: synthesis of vanillylamine from vanillin. Biocatal. Biotransform. 2014, 32, 302−313. (363) Anderson, M.; Afewerki, S.; Berglund, P.; Cordova, A. Total synthesis of capsaicin analogues from lignin-derived compounds by 14221

DOI: 10.1021/acs.chemrev.6b00486 Chem. Rev. 2016, 116, 14181−14224

Chemical Reviews

Review

combined heterogeneous metal, organocatalytic and enzymatic cascades in one pot. Adv. Synth. Catal. 2014, 356, 2113−2118. (364) Jagadish, R. S.; Divyashree, K. N.; Viswanath, P.; Srinivas, P.; Raj, B. Preparation of N-vanillyl chitosan and 4-hydroxybenzyl chitosan and their physico-mechanical, optical, barrier, and antimicrobial properties. Carbohydr. Polym. 2012, 87, 110−116. (365) Petrovic, V. P.; Simijonovic, D.; Zivanovic, M. N.; Kosaric, J. V.; Petrovic, Z. D.; Markovic, S.; Markovic, S. D. Vanillic Mannich bases: synthesis and screening of biological activity. Mechanistic insight into the reaction with 4-chloroaniline. RSC Adv. 2014, 4, 24635−24644. (366) Van, A.; Chiou, K.; Ishida, H. Use of renewable resource vanillin for the preparation of benzoxazine resin and reactive monomeric surfactant containing oxazine ring. Polymer 2014, 55, 1443−1451. (367) Sini, N. K.; Bijwe, J.; Varma, I. K. Thermal behaviour of bisbenzoxazines derived from renewable feed stock ’vanillin’. Polym. Degrad. Stab. 2014, 109, 270−277. (368) Sini, N. K.; Bijwe, J.; Varma, I. K. Renewable benzoxazine monomer from Vanillin: Synthesis, characterization, and studies on curing behavior. J. Polym. Sci., Part A: Polym. Chem. 2014, 52, 7−11. (369) Suvannasara, P.; Tateyama, S.; Miyasato, A.; Matsumura, K.; Shimoda, T.; Ito, T.; Yamagata, Y.; Fujita, T.; Takaya, N.; Kaneko, T. Biobased Polyimides from 4-Aminocinnamic Acid Photodimer. Macromolecules 2014, 47, 1586−1593. (370) Goy, R. C.; de Britto, D.; Assis, O. B. G. A review of the antimicrobial activity of chitosan. Polim.: Cienc. Tecnol. 2009, 19, 241− 247. (371) Shahidi, F.; Arachchi, J. K. V.; Jeon, Y.-J. Food applications of chitin and chitosans. Trends Food Sci. Technol. 1999, 10, 37−51. (372) Raafat, D.; von Bargen, K.; Haas, A.; Sahl, H.-G. Insights into the mode of action of chitosan as an antibacterial compound. Appl. Environ. Microbiol. 2008, 74, 3764−3773. (373) Nasti, A.; Zaki, N. M.; de Leonardis, P.; Ungphaiboon, S.; Sansongsak, P.; Rimoli, M. G.; Tirelli, N. Chitosan/TPP and chitosan/ TPP-hyaluronic acid nanoparticles: systematic optimization of the preparative process and preliminary biological evaluation. Pharm. Res. 2009, 26, 1918−1930. (374) Kast, C. E.; Bernkop-Schnürch, A. Thiolated polymers  thiomers: development and in vitro evaluation of chitosan− thioglycolic acid conjugates. Biomaterials 2001, 22, 2345−2352. (375) Hornof, M. D.; Kast, C. E.; Bernkop-Schnürch, A. In vitro evaluation of the viscoelastic properties of chitosan−thioglycolic acid conjugates. Eur. J. Pharm. Biopharm. 2003, 55, 185−190. (376) Clausen, A. E.; Kast, C. E.; Bernkop-Schnuerch, A. The Role of Glutathione in the Permeation Enhancing Effect of Thiolated Polymers. Pharm. Res. 2002, 19, 602−608. (377) Francesko, A.; Soares da Costa, D.; Lisboa, P.; Reis, R. L.; Pashkuleva, I.; Tzanov, T. GAGs-thiolated chitosan assemblies for chronic wounds treatment: control of enzyme activity and cell attachment. J. Mater. Chem. 2012, 22, 19438−19446. (378) Xu, T.; Xin, M.-H.; Li, M.-C.; Huang, H.-L.; Zhou, S.-Q.; Liu, J.-Z. Synthesis, characterization, and antibacterial activity of N,Oquaternary ammonium chitosan. Carbohydr. Res. 2011, 346, 2445− 2450. (379) Kulbokaite, R.; Ciuta, G.; Netopilik, M.; Makuska, R. NPEG’ylation of chitosan via ″click chemistry″ reactions. React. Funct. Polym. 2009, 69, 771−778. (380) Oliveira, J. R.; Martins, M. C. L.; Mafra, L.; Gomes, P. Synthesis of an O-alkynyl-chitosan and its chemoselective conjugation with a PEG-like amino-azide through click chemistry. Carbohydr. Polym. 2012, 87, 240−249. (381) Morimoto, M.; Nakajima, T.; Ishikura, M.; Shigemasa, Y.; Ifuku, S.; Saimoto, H. Synthesis of organo-soluble chitosan derivatives with polyphenolic side chains. Carbohydr. Polym. 2012, 90, 1259− 1264. (382) Zampano, G.; Bertoldo, M.; Ciardelli, F. Defined Chitosanbased networks by C-6-Azide-alkyne ″click″ reaction. React. Funct. Polym. 2010, 70, 272−281.

(383) Fernandes, M. M.; Francesko, A.; Torrent-Burgues, J.; Tzanov, T. Effect of thiol-functionalisation on chitosan antibacterial activity: Interaction with a bacterial membrane model. React. Funct. Polym. 2013, 73, 1384−1390. (384) Hong, C.-Y.; You, Y.-Z.; Bai, R.-K.; Pan, C.-Y.; Borjihan, G. Controlled polymerization of acrylic acid under 60Co irradiation in the presence of dibenzyl trithiocarbonate. J. Polym. Sci., Part A: Polym. Chem. 2001, 39, 3934−3939. (385) Bough, W. A.; Salter, W. L.; Wu, A. C. M.; Perkins, B. E. Influence of manufacturing variables on the characteristics and effectiveness of chitosan products. I. Chemical composition, viscosity, and molecular-weight distribution of chitosan products. Biotechnol. Bioeng. 1978, 20, 1931−1943. (386) Muthukrishnan, S.; Pan, E. H.; Stenzel, M. H.; BarnerKowollik, C.; Davis, T. P.; Lewis, D.; Barner, L. Ambient Temperature RAFT Polymerization of Acrylic Acid Initiated with Ultraviolet Radiation in Aqueous Solution. Macromolecules 2007, 40, 2978−2980. (387) Huang, M.; Jin, X.; Li, Y.; Fang, Y. e., Syntheses and characterization of novel pH-sensitive graft copolymers of maleoylchitosan and poly(acrylic acid). React. Funct. Polym. 2006, 66, 1041− 1046. (388) No, H. K.; Cho, Y. I.; Kim, H. R.; Meyers, S. P. Effective Deacetylation of Chitin under Conditions of 15 psi/121 °C. J. Agric. Food Chem. 2000, 48, 2625−2627. (389) Khan, T. A.; Peh, K. K.; Ch’ng, H. S. Reporting degree of deacetylation values of chitosan: the influence of analytical methods. J. Pharm. Pharm. Sci. 2002, 5, 205−212. (390) Maghami, G. G.; Roberts, G. A. F. Evaluation of the viscometric constants for chitosan. Makromol. Chem. 1988, 189, 195−200. (391) Luckachan, G. E.; Pillai, C. K. S. Chitosan/oligo L-lactide graft copolymers: Effect of hydrophobic side chains on the physicochemical properties and biodegradability. Carbohydr. Polym. 2006, 64, 254−266. (392) Wu, Y.; Zheng, Y.; Yang, W.; Wang, C.; Hu, J.; Fu, S. Synthesis and characterization of a novel amphiphilic chitosan-polylactide graft copolymer. Carbohydr. Polym. 2005, 59, 165−171. (393) Sahoo, D.; Sahoo, S.; Mohanty, P.; Sasmal, S.; Nayak, P. L. Chitosan: a new versatile bio-polymer for various applications. Des. Monomers Polym. 2009, 12, 377−404. (394) Fasman, G. D., Ed.; Prediction of Protein Structure and the Principles of Protein Conformation; Plenum: New York, 1989. (395) Fasman, G. D. Poly-α-Amino Acids, Protein Models for Conformational Studies; Dekker: New York, 1967. (396) Branden, C.; Tooze, J. Introduction to Protein Structure; Garland Science: New York, 1992. (397) Deming, T. J. The Practice of Peptide Synthesis, 2nd by M. Bodansky and A. Bodansky. J. Polym. Sci., Part A: Polym. Chem. 1995, 33, 2289. (398) Cheng, J.; Deming, T. J. Synthesis of polypeptides by ringopening polymerization of α-amino acid N-carboxyanhydrides. Top. Curr. Chem. 2011, 310, 1−26. (399) Kwon, G. S.; Naito, M.; Kataoka, K.; Yokoyama, M.; Sakurai, Y.; Okano, T. Block copolymer micelles as vehicles for hydrophobic drugs. Colloids Surf., B 1994, 2, 429−434. (400) Kricheldorf, H. R. α-Amino acid N-carboxy anhydrides and related heterocycles: syntheses, properties, peptide synthesis, polymerization; Springer-Verlag: Berlin, 1987. (401) Szwarc, M. Carbanions, Living Polymers and Electron Transfer Processes; Interscience: New York, 1968. (402) Deming, T. J. Synthesis and Self-Assembly of Well-Defined Block Copolypeptides via Controlled NCA Polymerization. Adv. Polym. Sci. 2013, 262, 1−37. (403) Caillol, S.; Lecommandoux, S.; Mingotaud, A.-F.; Schappacher, M.; Soum, A.; Bryson, N.; Meyrueix, R. Synthesis and Self-Assembly Properties of Peptide−Polylactide Block Copolymers. Macromolecules 2003, 36, 1118−1124. (404) Cao, J.; Hu, P.; Lu, L.; Chan, B. A.; Luo, B.-H.; Zhang, D. Nonionic water-soluble ″clickable″ [small alpha]-helical polypeptides: 14222

DOI: 10.1021/acs.chemrev.6b00486 Chem. Rev. 2016, 116, 14181−14224

Chemical Reviews

Review

synthesis, characterization and side chain modification. Polym. Chem. 2015, 6, 1226−1229. (405) Wang, Q.-M.; Gao, Z.; Liu, S.; Fan, B.; Kang, L.; Huang, W.; Jin, M. Hybrid polymeric micelles based on bioactive polypeptides as pH-responsive delivery systems against melanoma. Biomaterials 2014, 35, 7008−7021. (406) Yakovlev, I.; Deming, T. J. Controlled Synthesis of Phosphorylcholine Derivatives of Poly(serine) and Poly(homoserine). J. Am. Chem. Soc. 2015, 137, 4078−4081. (407) Qian, Z.-G.; Xia, X.-X.; Lee, S. Y. Metabolic engineering of Escherichia coli for the production of cadaverine: a five carbon diamine. Biotechnol. Bioeng. 2011, 108, 93−103. (408) Bermudez, M.; Leon, S.; Aleman, C.; Munoz-Guerra, S. Comparison of lamellar crystal structure and morphology of nylon 46 and nylon 5. Polymer 2000, 41, 8961−8973. (409) Wang, M.-S.; Huang, J.-C. Nylon 1010 properties and applications. J. Polym. Eng. 1994, 13, 155−174. (410) Nieschlag, H. J.; Rothfus, J. A.; Sohns, V. E.; Perkins, R. B., Jr. Nylon-1313 from brassylic acid. Ind. Eng. Chem. Prod. Res. Dev. 1977, 16, 101−107. (411) Goethals, F.; Martens, S.; Espeel, P.; van den Berg, O.; Du Prez, F. E. Diversely Substituted Polyamide Structures through ThiolEne Polymerization of Renewable Thiolactone Building Blocks. Macromolecules 2014, 47, 61−69. (412) Aboudzadeh, A.; Fernandez, M.; Munoz, M. E.; Santamaria, A.; Mecerreyes, D. Ionic Supramolecular Networks Fully Based on Chemicals Coming from Renewable Sources. Macromol. Rapid Commun. 2014, 35, 460−465. (413) Ghatge, N. D.; Maldar, N. N. Polyimides from dianhydride and diamine: structure property relations by thermogravimetric analysis. Polymer 1984, 25, 1353−1356. (414) Sadavarte, N. V.; Avadhani, C. V.; Naik, P. V.; Wadgaonkar, P. P. Regularly alternating poly(amideimide)s containing pendent pentadecyl chains: Synthesis and characterization. Eur. Polym. J. 2010, 46, 1307−1315. (415) Fenouillot, F.; Rousseau, A.; Colomines, G.; Saint-Loup, R.; Pascault, J. P. Polymers from renewable 1,4:3,6-dianhydrohexitols (isosorbide, isomannide and isoidide): A review. Prog. Polym. Sci. 2010, 35, 578−622. (416) Wroblewska, A.; Zych, A.; Thiyagarajan, S.; Dudenko, D.; van Es, D.; Hansen, M. R.; Koning, C.; Duchateau, R.; Jasinska-Walc, L. Towards sugar-derived polyamides as environmentally friendly materials. Polym. Chem. 2015, 6, 4133−4143. (417) Auvergne, R.; Caillol, S.; David, G.; Boutevin, B.; Pascault, J.-P. Biobased Thermosetting Epoxy: Present and Future. Chem. Rev. 2014, 114, 1082−1115. (418) Ismail, T. N. M. T.; Hassan, H. A.; Hirose, S.; Taguchi, Y.; Hatakeyama, T.; Hatakeyama, H. Synthesis and thermal properties of ester-type crosslinked epoxy resins derived from lignosulfonate and glycerol. Polym. Int. 2009, 59, 181−186. (419) Witayakran, S.; Ragauskas, A. J. Modification of high-lignin softwood kraft pulp with laccase and amino acids. Enzyme Microb. Technol. 2009, 44, 176−181. (420) Illy, N.; Benyahya, S.; Durand, N.; Auvergne, R.; Caillol, S.; David, G.; Boutevin, B. The influence of formulation and processing parameters on the thermal properties of a chitosan-epoxy prepolymer system. Polym. Int. 2014, 63, 420−426. (421) Kiuchi, H.; Kai, W.; Inoue, Y. Preparation and characterization of poly(ethylene glycol) crosslinked chitosan films. J. Appl. Polym. Sci. 2008, 107, 3823−3830. (422) Lertwattanaseri, T.; Ichikawa, N.; Mizoguchi, T.; Tanaka, Y.; Chirachanchai, S. Epoxidized natural rubber bionanocomposite: a model case of bionanocomposite using nanofibrous chitosan and its consequent functional properties. Chem. Lett. 2009, 38, 798−799. (423) Cestari, A. R.; Vieira, E. F. S.; Alves, F. J.; Silva, E. C. S.; Andrade, M. A. S., Jr. A novel and efficient epoxy/chitosan cement slurry for use in severe acidic environments of oil wells-Structural characterization and kinetic modeling. J. Hazard. Mater. 2012, 213214, 109−116.

(424) Kim, I. Y.; Kim, S. J.; Shin, M.-S.; Lee, Y. M.; Shin, D.-I.; Kim, S. I. pH- and thermal characteristics of graft hydrogels based on chitosan and poly(dimethylsiloxane). J. Appl. Polym. Sci. 2002, 85, 2661−2666. (425) Li, Y.; Xiao, F.; Moon, K.-S.; Wong, C. P. Novel curing agent for lead-free electronics: Amino acid. J. Polym. Sci., Part A: Polym. Chem. 2006, 44, 1020−1027. (426) Shibata, M.; Nakai, K. Preparation and properties of biocomposites composed of bio-based epoxy resin, tannic acid, and microfibrillated cellulose. J. Polym. Sci., Part B: Polym. Phys. 2010, 48, 425−433. (427) Takada, Y.; Shinbo, K.; Someya, Y.; Shibata, M. Preparation and properties of bio-based epoxy montomorillonite nanocomposites derived from polyglycerol polyglycidyl ether and ϵ-polylysine. J. Appl. Polym. Sci. 2009, 113, 479−484. (428) Das, G.; Karak, N. Mesua ferrea L. seed oil based amido-amine modified nanoclay/epoxy nanocomposites. J. Appl. Polym. Sci. 2012, 124, 2403−2414. (429) Desroches, M.; Benyahya, S.; Besse, V.; Auvergne, R.; Boutevin, B.; Caillol, S. Synthesis of bio-based building blocks from vegetable oils: A platform chemicals approach. Lipid Technol. 2014, 26, 35−38. (430) Cornille, A.; Froidevaux, V.; Negrell, C.; Caillol, S.; Boutevin, B. Thiol-ene coupling: An efficient tool for the synthesis of new biobased aliphatic amines for epoxy curing. Polymer 2014, 55, 5561− 5570. (431) Dworakowska, S.; Cornille, A.; Bogdal, D.; Boutevin, B.; Caillol, S. Formulation of bio-based epoxy foams from epoxidized cardanol and vegetable oil amine. Eur. J. Lipid Sci. Technol. 2015, 117, 1893−1902. (432) Zupancic, J. J.; Marine, A. A.; Vietti, D. E.; Vinci, D. Rohm and Haas Company, Dow Global Technologies LLC, U.S.A. Adhesive twocomponent composition of epoxy-terminated polyester, WO2015073959A1, 2015. (433) Vijayalakshmi, P.; Rao, T. C.; Kale, V.; Balakrishna, R. S.; Subbarao, R. Reactive polyamides from cycloaliphatic C21 dicarboxylic acid and their evaluation as epoxy curing agents. Polymer 1992, 33, 3252−3256. (434) Druelle, S.; Galy, J.; Michel-Dansac, F.; Pascault, J. P.; Montarnal, S.; Sautereau, H. Dow Chemical Co. Manufacture of epoxy resin compositions, WO8702686A1, 1987. (435) Wei, K.; Zhu, G.; Tang, Y.; Liu, T.; Xie, J. The effects of crosslink density on thermo-mechanical properties of shape-memory hydro-epoxy resin. J. Mater. Res. 2013, 28, 2903−2910. (436) Wei, K.; Zhu, G.; Tang, Y.; Niu, L. Shape-memory effects of a hydro-epoxy resin system. J. Polym. Res. 2013, 20, 1−7. (437) Mi, Z.; Nie, X.-a.; Wang, Y.-g.; Xia, C.; Lin, G.-f. Study on synthesis and performance of Rosin-derived polyamide as epoxy curing agent. J. For. Prod. Ind. 2013, 2, 5−11 17 pp. (438) Wang, H.; Wang, H.; Zhou, G. Synthesis of rosin-based imidoamine-type curing agents and curing behavior with epoxy resin. Polym. Int. 2011, 60, 557−563. (439) Liu, Y.; Wang, J.; Xu, S. Synthesis and curing kinetics of cardanol-based curing agents for epoxy resin by in situ depolymerization of paraformaldehyde. J. Polym. Sci., Part A: Polym. Chem. 2014, 52, 472−480. (440) Fache, M.; Monteremal, C.; Boutevin, B.; Caillol, S. Amine hardeners and epoxy cross-linker from aromatic renewable resources. Eur. Polym. J. 2015, 73, 344−362. (441) Mendis, G. P.; Hua, I.; Youngblood, J. P.; Howarter, J. A. Enhanced dispersion of lignin in epoxy composites through hydration and mannich functionalization. J. Appl. Polym. Sci. 2015, 132, 41263− 41271. (442) Panchenko, A.; Charrak, M.; Dahmen, K.; Brinks, M.; Schelwies, M.; Blank, B.; Piepenbrink, M.; Backes, R. BASF SE, Germany. Use of 2,5-bis(aminomethyl)furan as hardener for epoxy resins, EP2837645A1, 2015. (443) Jeol, S. Rhodia Operations, Fr. Curable composition and curable prepolymer composition and manufacture of an epoxy 14223

DOI: 10.1021/acs.chemrev.6b00486 Chem. Rev. 2016, 116, 14181−14224

Chemical Reviews

Review

thermoset for adhering two or more substrates., WO2015001061A1, 2015. (444) Tang, D.; Thiyagarajan, S.; Noordover, B. A. J.; Koning, C. E.; van Es, D. S.; van Haveren, J. Fully renewable thermoplastic poly(ester urethane urea)s from bio-based diisocyanates. J. Renewable Mater. 2013, 1, 222−229. (445) Carre, C.; Bonnet, L.; Averous, L. Original biobased nonisocyanate polyurethanes: solvent- and catalyst-free synthesis, thermal properties and rheological behaviour. RSC Adv. 2014, 4, 54018−54025. (446) Duval, C.; Kebir, N.; Jauseau, R.; Burel, F. Organocatalytic synthesis of novel renewable non-isocyanate polyhydroxyurethanes. J. Polym. Sci., Part A: Polym. Chem. 2016, 54, 758−764. (447) Duval, C.; Kebir, N.; Charvet, A.; Martin, A.; Burel, F. Synthesis and properties of renewable nonisocyanate polyurethanes (NIPUs) from dimethylcarbonate. J. Polym. Sci., Part A: Polym. Chem. 2015, 53, 1351−1359. (448) van Velthoven, J. L. J.; Gootjes, L.; van Es, D. S.; Noordover, B. A. J.; Meuldijk, J. Poly(hydroxy urethane)s based on renewable diglycerol dicarbonate. Eur. Polym. J. 2015, 70, 125−135. (449) Maisonneuve, L.; More, A. S.; Foltran, S.; Alfos, C.; Robert, F.; Landais, Y.; Tassaing, T.; Grau, E.; Cramail, H. Novel green fatty acidbased bis-cyclic carbonates for the synthesis of isocyanate-free poly(hydroxyurethane amide)s. RSC Adv. 2014, 4, 25795−25803. (450) Besse, V.; Auvergne, R.; Carlotti, S.; Boutevin, G.; Otazaghine, B.; Caillol, S.; Pascault, J.-P.; Boutevin, B. Synthesis of isosorbide based polyurethanes: An isocyanate free method. React. Funct. Polym. 2013, 73, 588−594. (451) Noel, A.; Borguet, Y. P.; Raymond, J. E.; Wooley, K. L. Poly(carbonate-amide)s Derived from Bio-Based Resources: Poly(ferulic acid-co-tyrosine). Macromolecules (Washington, DC, U. S.) 2014, 47, 2974−2983. (452) Ozawa, Y.; Shibata, M. Biobased thermosetting resins composed of L-lysine methyl ester and bismaleimide. J. Appl. Polym. Sci. 2014, 131, 40379−40387. (453) Zúñiga, C.; Lligadas, G.; Ronda, J. C.; Galià, M.; Cádiz, V. Renewable polybenzoxazines based in diphenolic acid. Polymer 2012, 53, 1617−1623. (454) Dumas, L.; Bonnaud, L.; Olivier, M.; Poorteman, M.; Dubois, P. High performance benzoxazine/CNT nanohybrid network − An easy and scalable way to combine attractive properties. Eur. Polym. J. 2014, 58, 218−225. (455) Illy, N.; Couture, G.; Auvergne, R.; Caillol, S.; David, G.; Boutevin, B. New prospects for the synthesis of N-alkyl phosphonate/ phosphonic acid-bearing oligo-chitosan. RSC Adv. 2014, 4, 24042− 24052. (456) The Global Amino Acids Market − Trends, drivers and projections; Global Industry Analysis Inc.: San Jose, CA, 2016. (457) Leykin, A.; Shapovalov, L.; Figovsky, O. Non-isocyanate polyurethanes: yesterday, today and tomorrow. Al'tern. Energ. Ekol. 2016, 03−04, 95−108. (458) Carus, M. Bio-based Building Blocks and Polymers in the World; Nova-Institut GmbH: Hürth, Germany, 2015.

14224

DOI: 10.1021/acs.chemrev.6b00486 Chem. Rev. 2016, 116, 14181−14224