Next Article in Journal
Molecular Mechanism Study on the Effect of Microstructural Differences of Octylphenol Polyoxyethylene Ether (OPEO) Surfactants on the Wettability of Anthracite
Next Article in Special Issue
Cascading One-Pot Synthesis of Biodegradable Uronic Acid-Based Surfactants from Oligoalginates, Semi-Refined Alginates, and Crude Brown Seaweeds
Previous Article in Journal
Analytical Quality by Design-Compliant Development of a Cyclodextrin-Modified Micellar ElectroKinetic Chromatography Method for the Determination of Trimecaine and Its Impurities
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Photocatalytic Transformation of Biomass and Biomass Derived Compounds—Application to Organic Synthesis

by
Mario Andrés Gómez Fernández
and
Norbert Hoffmann
*
CNRS, Université de Reims Champagne-Ardenne, ICMR, Equipe de Photochimie, UFR Sciences, B.P. 1039, 51687 Reims, France
*
Author to whom correspondence should be addressed.
Molecules 2023, 28(12), 4746; https://doi.org/10.3390/molecules28124746
Submission received: 27 April 2023 / Revised: 6 June 2023 / Accepted: 9 June 2023 / Published: 13 June 2023
(This article belongs to the Special Issue Sustainable Chemistry in France 2.0)

Abstract

:
Biomass and biomass-derived compounds have become an important alternative feedstock for chemical industry. They may replace fossil feedstocks such as mineral oil and related platform chemicals. These compounds may also be transformed conveniently into new innovative products for the medicinal or the agrochemical domain. The production of cosmetics or surfactants as well as materials for different applications are examples for other domains where new platform chemicals obtained from biomass can be used. Photochemical and especially photocatalytic reactions have recently been recognized as being important tools of organic chemistry as they make compounds or compound families available that cannot be or are difficultly synthesized with conventional methods of organic synthesis. The present review gives a short overview with selected examples on photocatalytic reactions of biopolymers, carbohydrates, fatty acids and some biomass-derived platform chemicals such as furans or levoglucosenone. In this article, the focus is on application to organic synthesis.

Graphical Abstract

1. Introduction

The chemical transformation of biomass as a renewable feedstock as well as catalysis play a key role in sustainable chemistry [1]. In many cases, biomass is also an important alternative to fossils in the context of international economic and political dependencies and conflicts. In the context of chemical valorization, it should be pointed out that the molecular structure of biomass is different form corresponding fossil carbon compounds [2,3,4]. For example, carbohydrates which represent 75% of the biomass have a higher oxygen contain [5,6]. Also, lignin (about 20% of the biomass), which is the essential biomass-based source for aromatic compounds, contains higher amounts of oxygen [7]. The presence of numerous functional groups in such compounds makes them particular interesting for the production of new platform chemicals [8,9,10,11,12] for industry and also for many applications to fine chemistry and organic synthesis [13,14,15,16,17,18,19,20]. Catalysis is a further key element of green chemistry [21]. The catalytic transformation of biomass or biomass-derived platform or fine chemicals thus permits synergistic effects of sustainable chemistry as two requirements of green chemistry [1] are combined [22,23].
Photochemical reactions have been recognized as being at the origin of sustainable chemistry [24,25,26,27,28]. They also play an important role in organic synthesis [29,30,31,32]. In this context, photochemical reactions also became very interesting in the chemical or pharmaceutical industry [33]. Applications of photochemical reactions in heterocyclic chemistry are particularly interesting [34,35,36]. Heterocycles are the partial structure of numerous bioactive compounds and play an important role in material science, for example for organic semi-conducting materials. More recently, photocatalytic reactions have gained an enormous interest in organic synthesis in the academic domain as well as in industry [33,37]. Various kinds of photocatalysis are reported [38]. They are characterized by different elementary steps such as sensitization by energy transfer, electron transfer or hydrogen transfer. The latter are often coupled and described as proton-coupled electron transfer (PCET) [39,40,41]. In particular, photocatalysis with visible light is now investigated [42,43]. The systematic study of these reactions opens perspectives for organic synthesis [44,45]. The generally mild photochemical reaction conditions facilitate a large number of C-H activation processes. Additionally, classical thermal reactions such as the Diels–Alder reaction can be facilitated [46]. Often in these cases, the reaction mechanism changes. A further important aspect is the combination of photoredox with other kinds of catalysis [47,48,49,50,51,52] such as enzyme [53,54,55,56,57], organocatalysis [58,59,60,61,62] or with conventional organometallic catalysis [63,64,65,66,67,68]. All these forms of catalysis are currently applied to the transformation of biomass or biomass-derived platform chemicals.
Further synergies in connection with sustainable chemistry are currently developed by systematic investigations of photocatalytic transformations of lignocellulosic material and platform chemicals obtained from biomass. Various aspects of this domain have been recently reviewed [69,70,71]. Therefore, the present article mainly focuses on the application of such strategies in fine chemistry or organic synthesis.

2. Photocatalytic Transformations of Biopolymers

By far most of the biomass is composed of polymers. Polysaccharides and lignins are the most important compound families. In many cases, photocatalytic transformation of such material yields mixtures of monomers [9,69,72,73]. For example, the depolymerization of cellulose is photo-catalyzed by TiO2 [71,74,75]. Reactive radical species are generated, which further react with oxygen. Under acidic conditions, such a catalysis may support the formation of furan derivatives, which are important platform chemicals. In a similar way, the photochemically supported Fenton oxidation has been used for depolymerization of polysaccharides [76].
Similar photoredox catalytic reactions have been carried out with lignin [77,78,79,80]. Lignin is an important renewable source for aromatic compounds (Figure 1) [7,81]. For example, production processes of vanillin from lignin are frequently studied [82]. Another research approach with these material deals with the study of photochemical reactions [83]. Studies on the depolymerization with processed or native lignin are particular challenging. In order to obtain information on possibly efficient elementary steps, reactions of partial structures of lignin are investigated. Using Pd catalysis lignin fragments such as 1 are selectively oxidized to the corresponding acetophenone derivatives 2 (Scheme 1) [84]. Primary alkyl alcohol functions are not transformed under these conditions [85]. In a second photoredox catalytic reaction, C-O bond cleavage takes place yielding the acetophenone derivatives 3 and phenols 4. The iridium complex [Ir(ppy)2(dtbbpy)]PF6 was used as photoredox catalyst. The first process was separately performed and a workup without purification was carried out. This was necessary because different solvents were used in both steps. The resulting raw material was subjected to the photoredox catalytic step. The photoredox catalytic step was also carried out under similar conditions with model oligomers or polymers of lignin such as 5. The corresponding monomers 6 and 7 were isolated in high yields.
The following mechanism has been discussed for the photoredox catalytic part (Scheme 2). After excitation of the [Ir(ppy)2(dtbbpy)] photocatalyst, which is an IrIII species, electron transfer from the sacrificial amine donor occurs and an IrII species is generated. The latter is capable of reducing the acetophenone derivative 1, which leads to a radical anion 8. Fragmentation yields a neutral radical 9 and a phenolat 10. Protonation yields the final product 11. The neutral electrophilic radical undergoes hydrogen abstraction of hydrogen atom transfer (HAT), yielding the final acetophenone derivative 12. This step may also involve electron transfer followed by proton transfer [86]. Two catalytic processes are performed consecutively in the present transformation.
This transformation was also carried out with sunlight. Similar transformations of small model compounds have been reported [87,88,89]. Such studies have also been performed in the context of the sustainable production of vanillin [90,91]. In the context of organic photochemical reactions, for example cycloaddition with alkenes, derivatives of this compound are particularly interesting because in such transformations a high degree of molecular complexity and diversity is generated without chemical activation and thus reducing waste production [27,92,93,94].

3. Carbohydrates

Carbohydrates are versatile platform compounds and synthons in organic synthesis. As enantiomerically pure compounds they or their derivatives are key intermediates in asymmetric synthesis [95,96]. The systematic arrangement of chiral centers is particularly interesting in this regard [97,98,99,100]. Photocatalytic reactions have been applied in this domain [101]. In the present example, the selective oxidation of the primary alcohol function (13) into an aldehyde (14) plays a key role (Scheme 3) [102]. No secondary functions should be oxidized. Thus many syntheses with carbohydrate synthons can be facilitated since the number of protecting and deprotecting steps is reduced. The reaction is well known and can be carried out with galactose oxidase and molecular oxygen [103]. Since this enzyme activity is limited to galactose derivatives, various biomimetic strategies have been developed in order to extend the reactivity to further hexoses such as glucose or mannose derivatives [104,105,106,107]. Aside other methods, the Semmelhack reaction [108] can be used for the oxidation of primary alcohols to aldehydes [109]. Molecular oxygen is used as oxidant and copper salts and TEMPO (2,2,6,6-Tetramethylpiperidinyloxyl) are used as catalysts or as co-catalysts. Cu(II) acts as an oxidant (Scheme 4) [110]. This species (15) is generated by the addition of TEMPO and release of TEMPOH. In this ligand exchange process, the primary alcohol is added (16). The release of the carbonyl product regenerates the Cu(I) catalyst. The co-catalyst TEMPO is regenerated by the oxidation of TEMPOH by molecular oxygen. It has been observed that when the transformation is carried out with visible light irradiation, the reaction becomes efficient and can be applied to organic synthesis (Scheme 3) [102]. This effect can be explained by the fact that the Cu-O bond is weakened when such complexes are photochemically excited via a ligand to metal charge transfer (LMCT). Thus, ligand exchange steps in the mechanism are accelerated [111]. The oxidation product is stabilized by the formation of an anhydro derivative such as 17. Isolation and purification of these products were difficult. However, in situ transformation by a reductive amination efficiently yielded the corresponding glucose and mannose derivatives 19 and 22. Elimination of the benzylic protecting group followed by an intramolecular reductive amination yields the tetrahydroxyazepanes 20 and 23. Using the enzyme catalytic oxidation with galactose oxidase, tetrahydroxyazempanes with glucose- or mannose-related configurations are not accessible. Tetrahydroxyazepanes are azasugars. Such compounds possess various biological activities such as the inhibition of glycosidases and glycosyltranferases or anti-cancer activities [112,113,114]. In the present synthesis many steps of protecting and deprotecting different hydroxyfunctions and the utilization of corresponding reagents are avoided. As the irradiation was performed with visible light with visible light, the use of sunlight can be projected [28,115].
Glycosylation plays a central role in carbohydrate chemistry. During recent years various photocatalytic methods have been developed for this purpose [116,117]. Many of these reactions are carried out with visible light and with corresponding LED’s as light source. Such reaction conditions are particularly mild and energy consumption is low. Carboxonium intermediates 25 play an essential role in such transformations (Scheme 5) [118,119]. These species are generated from precursors 24 with corresponding leaving groups. Glycosylation then occurs by a nucleophilic attack. As these intermediates carry a positive charge, protonating agents or Lewis acidic compounds are often used as promotors. Oxidation such as photochemical electron transfer induced release of the leaving group is an alternative [120]. In this context, photoredox catalytic reactions have been systematically studied. Thus, compound 26 carrying an electron rich thioanisyl as the leaving group at the anomeric center was efficiently transformed with 28 into the corresponding glycosylation product 27 (Scheme 5) [121]. Additionally, more complex compounds such as the serine derivative 29 were successfully transformed. In most cases, the formation of β-anomers was favored. Tetrahalogenmethanes were added to the reaction mixture. They play a key role in the initiation step of the catalytic cycle. The following mechanism has been discussed. The reaction starts with the photochemical excitation of the iridium photocatalyst (Scheme 6). The IrIII species is oxidized by a tetrahalogen methane. The resulting IrIV complex is reduced by the anisylthioether function (PMPS) in 26. The photocatalyst is thus regenerated. After the release of the PMPS radical from 31 and the nucleophilic, the attack of the alcohol compound takes place (32). The photochemically excited IrIII complex can also be oxidized in the reaction with intermediates of the PMPS leaving group. As it has already been pointed out, acidic activation of a leaving group leads to the formation of the carboxonium ion intermediate 25 (Scheme 5). Electronic excitation can significantly increase the acidity [122,123] and thus induce the release of a leaving group [124]. Thus, the trichloroimidate group was activated by photochemically excited phenols or naphthols [125]. Thiourea derivatives have also been used for the same purpose [126]. In some cases, both electron and proton transfer processes are involved in this activation [127].
C-Glycosids play an important role as bioactive compounds [128]. In such compounds, an acetal C-O bond is replaced by a more stable C-C bond. Thus, many of these products possess enzyme inhibitor properties. For example, glucosidases or many nucleoside transformations can be inhibited. In the latter case, a C-N bond is replaced by a C-C bond [129]. In the reaction of the 1-β-d-glucofruanose derivative 33 with p-methoxyphenylacetylene 34, only one diastereoisomer of the C-glucosid 35 was formed (Scheme 7) [130]. The photoredox catalytic part is carried out with an iridium catalyst. Compound 33 is decarboxylated and the intermediate 36 is formed. Such photodecarboxylation reactions are now frequently studied [131]. In the copper catalytic cycle, the complex 37 is formed, which reacts with the radical species 36 leading to the final product. The copper catalysis part is activated by addition of the ligand L2. In this case organic photocatalysts were less efficient. It cannot be completely excluded that the copper catalytic step is not affected by light absorption. In the present report, 28 C-glucosides have been synthesized with yields up to 96%. Heteroaromatic and alkyl acetylene derivatives have also been added to furanosyl radicals.
Chemical reactions can efficiently be induced by photosensitization. In this case, a chemical compound absorbs light and transfers its excitation energy to the substrate, which is electronically excited in this way. The latter then undergoes a photochemical reaction. The ground state of photosensitizer is regenerated and a new sensitizing process may take place. The photosensitizer acts as catalyst [38]. Often ketones are used as photosensitizer that efficiently transfer their triplet energies to a substrate. For example, this technique is applied to the photochemical transformation of α,β-unsaturated lactones. Often, these compounds possess a high excitation energy and low wavelength light irradiation is necessary for direct excitation to the S1 state. Such conditions leads to unselective reactions. Since triplet energies are lower, triplet energy transfer from a sensitizer often leads to more selective reactions. More convenient reaction conditions such as room temperature instead of low temperature can be applied. Such a sensitization of a α,β-unsaturated butyrolactone or furanone is depicted in Scheme 8. Acetone absorbs light and is transferred to its singlet state S1 38 possessing nπ* character. After intersystem crossing to the corresponding energy lower triplet state, T1 39 is reached. Triplet energy transfer now occurs to the furanone 40 that is thus transferred to its triplet state T1 41 with ππ* character. The latter species undergoes photochemical reactions. Acetone also absorbs at lower wavelengths (λmax (nπ*) = 278 nm) [132]. However, irradiation at longer wavelengths such as 300 nm generates sufficient excited molecules for sensitization when acetone is used as a solvent or co-solvent.
Under these conditions, furanone compounds such as 42 carrying a glucosyl moiety have been electronically excited to their triplet state 43 (Scheme 9) [133]. Sugar derivatives with their secondary alcohol functions are excellent hydrogen donors for hydrogen abstraction or hydrogen atom transfer (HAT) processes [134,135]. Such reaction steps are also observed in non-catalytic photochemical transformations. For some examples, see refs. [136,137,138]. Hydrogen abstraction is observed with many triplet exited compounds [94,135,137,139,140]. In the photosensitized reaction of glucosylated furanone derivative 42, the spirocyclic compounds 44 and 45 are formed. In contrast to many conventional radical addition reactions of such α,β-unsaturated lactones, a C-C bond is formed in the α-position. The electronic excitation occurs via triplet energy transfer from acetone to 42. The 3(ππ)* state (43) is characterized by the suppression of the C=C π-bond and an increased spin density in the β-position [135,141]. For this reason, hydrogen abstraction occurs from the anomeric center of the sugar moiety into this position and the intermediates 46 and 47 are formed, both in conformational equilibrium. Radical combination yields the final products 44 and 45. In this reaction, the regioselectivity of hydrogen abstraction depends on the relative configuration of the chiral centers in the anomeric position and the chiral center at the furanone moiety. When the corresponding α-anomer diastereoisomer 48 is transformed under the same conditions, hydrogen abstraction occurs at the position 5′ of the glucosyl substituent and the macrocyclic product 49 is formed. Similar reactions have been carried out with direct excitation of a furanone moiety to the singlet state and subsequent intersystem crossing to the triplet state. The transformation under these conditions needed lower wavelength irradiation and low reaction temperature [142]. It should further be pointed out that for the triplet sensitized of C-H activation, no particular reagents are needed, which often leads to the production of additional waste. Furthermore, acetone is used as a co-solvent and no chromatographic separation of the sensitizer or a photocatalyst is necessary.

4. Fatty Acids

Fatty acids obtained from triglycerides play an important role in the chemical valorization of biomass. Their use for the production of surfactants belong to the basic culture techniques of humanity. More recently their application as part of biofuel is reported [143]. Fatty acid methyl esters (FAME) can be obtained by photocatalytic esterification to produce biodiesel [144,145,146].
Photodecarboxylation is an interesting transformation both for biofuel production and in organic synthesis [147,148,149] as it leads to radical intermediates that can form alkanes or further react with a variety of substrates. Decarboxylation is also carried out using photochemical electron transfer [150,151,152,153,154]. Such reaction steps also play an important role in many metal catalyzed reactions [131,150,155,156]. In this context, but also for application to organic synthesis, decarboxylation is an interesting transformation as it leads to a variety of radical intermediates.
Hatanaka et al. have transformed palmitic acid into the corresponding n-pentadecane [157]. This method consists of the photoelectron transfer of the carboxylate with a photogenerated cation radical of phenantrene followed by decarboxylation. The hydrogen abstraction from a thiol results in the formation of the corresponding alkane. Nicewicz et al. used of an acridinium photooxidant in the presence of an organic base for a decarboxylation [158]. The mechanism involves a photochemical electron transfer to the acridinium from the carboxylate followed by CO2 release. The generated carbon radical abstracts a hydrogen from the in situ generated thiophenol. Wang et al. reported a route for photocatalytic decarboxylation for alkane production using fatty acids. Their system uses semiconductor Pt/TiO2 and irradiation using 365 nm LEDs, and a hydrogen atmosphere resulted to be crucial for the high yields (>90%), while several fatty acids were successfully transformed in the corresponding alkanes [159].
Enzyme-catalyzed reactions play a key role in the transformation of biomass or biomass-derived compounds. Recently, an algal photoenzyme named fatty acid photodecarboxylase (CvFAP) was discovered. This photoenzyme transforms fatty acids into the corresponding alkanes upon light irradiation [160,161,162]. Among the different uses of this photo-enzyme, several methods were reported to obtain, for example, bio-propane and bio-butane from fatty acids [163,164], and also bio-diesel [165] depending on the substrate that was used.
Concerning the application of such reactions to organic synthesis, radical intermediates are often generated by decarboxylation of carboxylic acids or their derivatives [166,167,168,169,170]. These radicals undergo addition to various acceptor molecules. In many of these reactions, photocatalysis involving electron transfer is applied. The transformations are often also carried out with corresponding fatty acids or their derivatives.
To perform the photodecarboxylation of fatty acids followed by radical reactions, a strategy using an activated form of the fatty acid was used. König et al. used the N-hydroxyphtalimide activated esters such as 50 of palmitic acid to perform the photodecarboxylation reaction (Scheme 10) [171]. They used 10 mol% of eosin Y, an organic and unexpensive sensitizer, and irradiation with green light (λ = 528 nm) to start the photodecarboxylation, and the resulting radical underwent coupling with alkynes such as 51 carrying a sulfone group. Diisopropylethylamine (DIPEA) was used as an electron donor in the photoredox catalytic process. The same reaction was successfully carried out with corresponding derivatives of lauric acid, myristic acid or stearic acid. Corresponding derivatives of unsaturated acids such as oleic or linoleic acid were also efficiently transformed into the corresponding coupling products. Glorius et al. [172] reported the C-H alkylation of heteroarenes using aliphatic carboxylic acids. Their method used an iridium photocatalyst in 0.5 mol% to perform the process. The group tolerance is large and two fatty acids were also used in this transformation.
Photodecarboxylation can also be used to further functionalize organic molecules; for example, a decarboxylative borylation was described by Li et al. to obtain the corresponding alkyl boronates (Scheme 11) [173]. They performed the reaction using the activated N-acyloxyphtalimides such as 53 obtained from oleic acid. Irradiation using visible light and an iridium photocatalyst resulted in the photodecarboxylation and the radicals were successfully trapped by a B2pin2 derived intermediate to obtain the corresponding borylated product 54. The photoredox catalysis mechanism depicted in Scheme 12 has been suggested. The photochemically excited Ir(III) catalyst is able to reduce the N-acyloxyphtalimide 53 yielding the radical anion 55 and an Ir(IV) species. Fragmentation of this intermediate yields the carboxyl radical 56, which immediately releases carbon dioxide. Concomitantly, the phthalimide anion 57 is formed. Protonation of the latter by water leads to hydroxide ions, which react with B2pin2 (58). Reaction of 58 with the alkyl radicals (R·) yields the final product 54. The boron ester-derived radical anion 59 is oxidized by the Ir(IV) species, thus regenerating the photocatalyst.
The same reaction was also successfully carried out with a musristic acid derivative leading to a corresponding borylated product. In a modified procedure, corresponding derivatives carrying a BF3K group have been synthesized. In a different report, Aggarwal et al. [174] reported the conversion of the same activated N-acyloxyphtalimides into boronic esters by simple irradiation with blue LEDs, no additional additives or metal catalysts were needed in this case. The photogenerated radicals were trapped by bis(catecholato)diboron. In this report, among the several substrates used, three fatty acids were used with yields up to 90%.
Glorius et al. have efficiently transformed N-(acyloxy)-phthalimides of fatty acids such as 50 (Scheme 10) into corresponding terminal alkenes (Figure 2) using photocatalysis with an organic photocatalyst (dihydrophenazine derivative) combined with copper catalysis [175]. Compound 60 was synthesized on gram-scale.
Radicals generated by decarboxylation of fatty acids are also involved in complex radical tandem reactions. Such a reaction with oleic acid has been described by Zhu et al. (Scheme 13) [176]. Radicals are generated from this acid involving photoredox catalysis with visible light and, an iridium catalyst and PhI(OAc)2. Michael addition to the metacrylanilide 61 and cyclization (62) yields a quaternary oxindole 63. In the present example, two C-C bonds are formed under the same photocatalytic conditions, thus avoiding a further activation.

5. Biomass-Derived Platform Chemicals

Levoglucosenone is an interesting platform chemical, which is obtained from pyrolysis of cellulose under acidic reaction conditions [177,178,179]. Recently, it gained additional interest as a precursor of a biobased dipolar solvent [180]. Levoglucosenone is now produced on the industrial scale. Different functional groups are localized on a relatively small enantiomerically pure molecule. Thus, it becomes a very attractive synthon for organic synthesis [181,182,183,184,185]. Levoglucosenone can also be transformed into further platform chemicals such as furanone derivatives [186,187].
Photocatalytic reactions have also been performed with this compound. Using tetra-n-butylammonium decatungstate (TBADT) as a photocatalyst, various radical species have been added to levoglucosenone 64 (Scheme 14) [188]. Thus, alkanes (65a) or cyclic ethers (65b) have been added. The addition of formamide (65c) was particular efficient. The method was also used to the addition of aromatic (65d), heteroaromatic (65e) and aliphatic aldehydes (65f). The scope for the addition of aldehydes is large [189]. From these precursors radical intermediates are generated by hydrogen atom transfer (HAT) to the photochemically excited TBADT (Scheme 15) [190,191,192,193]. The radicals add stereospecifically anti with respect to the (-CH2-O-) moiety in 64. The energy difference for both diastereotopic attacks of the radical intermediates is 5 kcal∙mol−1. The configuration of further chiral centers generated in the reaction as in 65b was not controlled. The electrophilic radical intermediate 66 is reduced by [HW10O32]4−, which leads to the final product 65. In the context of sustainable chemistry, it should further be mentioned that such reactions have been carried out with sunlight and under very simple apparatus conditions [191].
Furans are valuable platform chemicals obtained from carbohydrates [194,195,196]. They are formed by the dehydration of sugars. Thus, dehydration of pentoses or hemicellososes yields furfural while corresponding transformations of hexoses lead to hydroxymethylfurfural (HMF). These compounds are currently studied in connection with the production bulk products such as new polymers [194,197] or surfactants [198,199,200].
One of the most-used photocatalytic reactions of furans is the photooxygenation of these compounds involving singlet oxygen. Oxygen is one of the rather rare molecules possessing a triplet multiplicity at the ground state [201,202]. The reactivity of both oxygen species is significantly different. While singlet oxygen undergoes preferentially polar reactions, for example with alkenes, triplet oxygen is characterized by its radical reactivity [203,204,205]. Most conveniently singlet oxygen is produced by photosensitization or photocatalysis (Scheme 16). In this case, a sensitizer is photochemically excited, generally to the singlet state. After intersystem crossing (isc) to the triplet state, energy transfer to triplet oxygen occurs by interaction of two species with the same spin multiplicity. The sensitizer returns to the singlet ground state and the oxygen molecule is excited to the singlet state. As the excitation energy of oxygen is rather low (23 kcal∙mol−1), dyes absorbing visible light such as rose bengale 67 or methylene blue 68 are often used. These dyes are soluble in protic solvents such as alcohols. For transformations in un-polar solvents, tetramethylporphyrine 69 is often used. Additionally, coordination compounds such as ruthenium complexes or nanoparticles as they are used for photoredox catalysis can be used for singlet oxygen production [206,207]. Oxidations with singlet oxygen generally occur under particularly mild reaction conditions and often with high selectivity and yields. Thus, numerous applications to organic synthesis have been reported [208]. Recently, an industrial process has been developed for the production of artemisinin where the photooxygenation is used as a key step [209]. This compound plays an important role in the treatment of malaria [210,211]. In this natural compound, a trioxane moiety is the pharmacophore. The photooxygenation of allylalcohols has been used to synthesize structural analogues [212,213,214]. Terpenes represent an important biomass-derived compound family [215,216,217,218]. The photooxygenation of these compounds is also widely applied, for example, to the industrial production of fragrances and flavors [219,220] as well as to the preparation of many other compounds [208,221,222]. Various catalytic techniques such as heterogeneous catalysts or micro and flow reactors have been developed in order to optimize the reaction [223,224,225,226,227,228].
Furans are excellent substrates for photooxygenation with singlet oxygen [229,230]. As pointed out above, furfural 70 is a platform chemical obtained from hemicellulose or pentoses by dehydration. Its photooxygenation with singlet oxygen yields hydroxyfuranone (Scheme 17) [231,232,233]. In most cases of photooxygenation of furan derivative, an endoperoxide such as 71 is formed. As the reaction is carried out in an alcohol as solvent, this intermediate undergoes a nucleophilic attack of an alcohol molecule leading to the final products 5-hydroxyfuran-2[5H]-one 72 and a formic ester 73. This very convenient oxidation can be carried out in the laboratory at 100 g scale [234]. Hydroxyfuranone 72 is itself a platform chemical with numerous applications to organic synthesis [235,236]. Chiral synthons can be obtained from this compound and many efficient applications to asymmetric synthesis have been reported [237,238,239,240]. At the electronically excited state, reactions are generally less stereoselective due to the suppression of the C-C π-bond [141]. Hydorxyfuranone 72 was also used for the preparation of biodegradable surfactants (Figure 3) [241]. A photoredox catalytic process was also applied for the preparation of such compounds [242,243].
Photooxygenation was also applied as a key step to the synthesis compound 74 (Scheme 18, also compare Scheme 9). Using conventional carbohydrate chemistry methodology, it is often difficult to prepare α-anomeric derivatives of hexose pyranoses such as derived from glucose for example. Isomaltulose is a disaccharide obtained by enzymatic isomerization of saccharose [244,245]. In the isomaltulose molecule, a fructose moiety in its furanose form is connected in the α-anomeric position of glucose. Selective dehydration on the fructose furanoside under acidic conditions leads to the formation a hydroxymethylfurfural derivative (75) [246]. Acylation (76) and photooxygenation yields the product 77. Finally, reduction of the hydroperoxide 77 leads to 74 [133]. The present synthesis also provides a very convenient access to 5-hydroxymethylfuranones such as 74. Using conventional methods of carbohydrate chemistry, these compounds are prepared with multi-step syntheses from sugars or amino acids [247,248].

6. Conclusions

Biomass or biomass-derived platform chemicals have become an alternative feedstock for chemical industry. They may replace conventional fossil-based resources in two senses. On one hand biomass based compounds can be used for the synthesis of already existing products in the chemical industry as they are available mainly from petroleum. On the other hand, these biomass or biomass-derived compounds can be used to produce new and original compounds, which is facilitated by characteristic structure elements such as the presence of numerous hydroxyl functions in carbohydrates. Photochemical or photocatalytic reactions make compounds or compound families available that cannot or difficultly synthesized using conventional methods of organic synthesis. A high degree of molecular complexity and diversity is thus generated in a facile way. This is explained by the complementarity of ground state and excited state reactivity. In this same context, it should be pointed out that in the past, photochemical or photocatalytic reactions of biomass-derived products have rarely been investigated in a systematic way.
Photochemical and catalytic reactions on one hand and the utilization of biomass as a renewable feedstock on the other are part of sustainable chemistry. The combination of both opens perspectives for sustainable chemistry.

Funding

Our research work is currently funded by the ANR (French National Research agency, Projects IMPHOCHEM and NoPerox), the Communauté Urbaine Grand Reims and the Université de Reims Champagne-Ardenne. We acknowledge their support.

Data Availability Statement

No new data were created or analyzed in this study. Data sharing is not applicable to this article.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Anastas, P.T.; Kirchhoff, M.M. Origins, Current Status, and Future Challenges of Green Chemistry. ACC Chem. Res. 2002, 35, 686–694. [Google Scholar] [CrossRef] [PubMed]
  2. Wertz, J.-L.; Bédué, O. Lignocellulosic Biorefineries; EPFL Press: Lausanne, Switzerland, 2013. [Google Scholar]
  3. Barrault, J.; Kervennal, J.; Isnard, P. La chimie durable: Pour l’environement, l’économie notre socciété (Numéro thématique). Actual. Chim. 2018, 15–116. [Google Scholar]
  4. Marion, P.; Bernela, B.; Piccirilli, A.; Estrine, B.; Patoullard, N.; Guilbot, J.; Jérôme, F. Sustainable chemistry: How to produce better and more from less? Green Chem. 2017, 19, 4973–4989. [Google Scholar] [CrossRef]
  5. Lichtenthaler, F.W.; Peters, S. Carbohydrates as green raw materials for the chemical industry. Comptes Rendus Chim. 2004, 7, 65–90. [Google Scholar] [CrossRef]
  6. Lichtenthaler, F.W. Carbohydrate-based Product Lines. In Biorefineries–Industrial Processes and Products; Kamm, B., Gruber, P.R., Kamm, M., Eds.; Wiley-VCH: Weinheim, Germany, 2006; Volume 1, pp. 2–59. [Google Scholar]
  7. Heitner, C.; Dimmel, D.R.; Schmidt, J.A. (Eds.) Lignin and Lignans; CRC Press: Boca Raton, FL, USA, 2010. [Google Scholar]
  8. Farmer, T.J.; Mascal, M. Platfom molecules. In Introduction to Chemicals from Biomass, 2nd ed.; Clark, J., Deswarte, F., Eds.; Wiley: Chichester, UK, 2015; pp. 89–152. [Google Scholar]
  9. Cramail, H.; Bizet, B.O.; Durand, P.-L.; Hibert, G.; Grau, E. Bio-Sourced Polymers: Recent Advances In Avanced Green Chemistry, Part 2; Horváth, I.T., Malacria, M., Eds.; World Scientific: Singapore, 2020; pp. 167–328. [Google Scholar]
  10. Esposito, D.; Antonietti, M. Redefining biorefinery: The search for unconventional building blocks for materials. Chem. Soc. Rev. 2015, 44, 5821–5835. [Google Scholar] [CrossRef] [Green Version]
  11. Monet, E. La forêt: Un gisement privilege et durable de molecules biosourcées. Actual. Chim. 2023, 14–23. [Google Scholar]
  12. de Rezende Locatel, W.; Guilhaume, N.; Schuurman, Y.; Laurenti, D. La pyrolyse du bois et la conversion catalytique de ses vapeurs. Actual. Chim. 2023, 71–77. [Google Scholar]
  13. Gallezot, P. Conversion of biomass to selected chemical products. Chem. Soc. Rev. 2012, 41, 1538–1558. [Google Scholar] [CrossRef]
  14. Tuck, C.O.; Pérez, E.; Horvath, I.; Sheldon, R.A.; Poliakoff, M. Valorization of Biomass: Deriving More Value from Waste. Science 2012, 337, 695–699. [Google Scholar] [CrossRef]
  15. Ravelli, D.; Samorì, C. (Eds.) Biomass Valorisation; Wiley-VCH: Weinheim, Germany, 2021. [Google Scholar]
  16. Corma, A.; Iborra, S.; Velty, A. Chemical Routes for the Transformation of Biomass into Chemicals. Chem. Rev. 2007, 107, 2411–2502. [Google Scholar] [CrossRef]
  17. Chatterjee, C.; Pong, F.; Sen, A. Chemical conversion pathways for carbohydrates. Green Chem. 2015, 17, 40–71. [Google Scholar] [CrossRef]
  18. Alonso, D.M.; Wettstein, S.G.; Dumesic, J.A. Bimetallic catalysts for upgrading of biomass to fuels and chemicals. Chem. Soc. Rev. 2012, 41, 8075–8098. [Google Scholar] [CrossRef] [PubMed]
  19. Groß, J.; Kühlborn, J.; Opatz, T. Applications of xylochemistry from laboratory to industrial scale. Green Chem. 2020, 22, 4411–4425. [Google Scholar] [CrossRef]
  20. Kühlborn, J.; Groß, J.; Opatz, T. Making natural products from renewable feedstocks: Back to the roots? Nat. Prod. Rep. 2020, 37, 380–424. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  21. Sheldon, R.A.; Arends, I.; Hanefeld, U. (Eds.) Green Chemistry and Catalysis; Wiley-VCH: Weinheim, Germany, 2007. [Google Scholar]
  22. Centi, G.; van Santen, R.A. (Eds.) Catalysis for Renewables; Wiley-VCH: Weinheim, Germany, 2007. [Google Scholar]
  23. Imhof, P.; van der Waal, J.C. (Eds.) Catalytic Process Development for Renewable Materials; Wiley-VCH: Weinheim, Germany, 2013. [Google Scholar]
  24. Protti, S.; Manzini, S.; Fagnoni, M.; Albini, A. RSC Green Chemistry Book Series, Eco-Friendly Synthesis of Fine Chemicals; Ballini, R., Ed.; Royal Society of Chemistry: London, UK, 2009; pp. 80–111. [Google Scholar]
  25. Ciamician, G. The Photochemistry of the Future. Science 1912, 36, 385–394. [Google Scholar] [CrossRef] [Green Version]
  26. Ciamician, G. Sur les actions de la lumière. Bull. Soc. Chim. Fr. 1908, 3, i. [Google Scholar]
  27. Hoffmann, N. Photochemical reactions of aromatic compounds and the concept of the photon as a traceless reagent. Photochem. Photobiol. Sci. 2012, 11, 1613–1641. [Google Scholar] [CrossRef]
  28. Oelgemöller, M.; Jung, C.; Mattay, J. Green photochemistry: Production of fine chemicals with sunlight. Pure Appl. Chem. 2007, 79, 1939–1947. [Google Scholar] [CrossRef]
  29. Turro, N.J.; Schuster, G. Photochemical Reactions as a Tool in Organic Syntheses. Science 1975, 187, 303–312. [Google Scholar] [CrossRef]
  30. Hoffmann, N. Photochemical Key Steps in organic Synthesis. Chem. Rev. 2008, 108, 1052–1103. [Google Scholar] [CrossRef]
  31. Bach, T.; Hehn, J.P. Photochemical Reactions as Key Steps in Natural Product Synthesis. Angew. Chem. Int. Ed. 2011, 50, 1000–1052. [Google Scholar] [CrossRef] [PubMed]
  32. Kärkäs, M.D.; Porco, J.A.; Stephenson, C.R.J. Photochemical Approaches to Complex Chemotypes: Applications in Natural Product Synthesis. Chem. Rev. 2016, 116, 9683–9747. [Google Scholar] [CrossRef] [PubMed]
  33. Bonfield, H.E.; Knauber, T.; Lévesque, F.; Moschetta, E.G.; Susanne, F.; Edwards, L.J. Photons as a 21st century reagent. Nat. Commun. 2020, 11, 804. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  34. Lefebvre, C.; Fortier, L.; Hoffmann, N. Photochemical Rearrangements in Heterocyclic Chemistry. Eur. J. Org. Chem. 2020, 2020, 1393–1404. [Google Scholar] [CrossRef]
  35. Latrache, M.; Hoffmann, N. Photochemical radical cyclization reactions with imines, hydrazones, oximes and related compounds. Chem. Soc. Rev. 2021, 50, 7418–7435. [Google Scholar] [CrossRef]
  36. Hoffmann, N. Enantioselective synthesis of heterocyclic compounds using photochemical reactions. Photochem. Photobiol. Sci. 2021, 20, 1657–1674. [Google Scholar] [CrossRef]
  37. Michelin, C.; Hoffmann, N. Photocatalysis applied to organic synthesis—A green chemistry approach. Curr. Opin. Green Sustain. Chem. 2018, 10, 40–45. [Google Scholar] [CrossRef]
  38. Michelin, C.; Hoffmann, N. Photosensitization and Photocatalysis—Perspectives in Organic Synthesis. ACS Catal. 2018, 8, 12046–12055. [Google Scholar] [CrossRef]
  39. Tyburski, R.; Liu, T.; Glover, S.D.; Hammarström, L. Proton-Coupled Electron Transfer Guidelines, Fair and Square. J. Am. Chem. Soc. 2021, 143, 560–576. [Google Scholar] [CrossRef]
  40. Miller, D.C.; Tarantino, K.T.; Knowles, R.R. Proto-Coupled Electron Transfer in Organic Synthesis: Fundamentals, Applications, and Opportunities. Top. Curr. Chem. 2016, 374, 30. [Google Scholar] [CrossRef] [Green Version]
  41. Hoffmann, N. Proton-Coupled Electron Transfer in Photoredox Catalytic Reactions. Eur. J. Org. Chem. 2017, 2017, 1982–1992. [Google Scholar] [CrossRef] [Green Version]
  42. Stephenson, C.R.J.; Yoon, T.P.; MacMillan, D.W.C. (Eds.) Visible Light Photocatalysis in Organic Synthesis; Wiley-VCH: Weinheim, Germany, 2018. [Google Scholar]
  43. König, B. (Ed.) Chemical Photocatalysis, 2nd ed.; De Gruyter: Berlin, Germany, 2020. [Google Scholar]
  44. Marzo, L.; Pagire, S.K.; Reiser, O.; König, B. Visible-Light Photocatalysis: Does It Make a Difference in Organic Synthesis? Angew. Chem. Int. Ed. 2018, 57, 10034–10072. [Google Scholar] [CrossRef] [PubMed]
  45. Barata-Vallejo, S.; Yerien, D.E.; Postigo, A. Bioinspired Photocatalyzed Organic Synthetic Transformations. The Use of Natural Pigments and Vitamins in Photocatalysis. ChemCatChem 2022, e202200623. [Google Scholar] [CrossRef]
  46. Mattay, J. Charge Transfer and Radical Ions in Photochemistry. Angew. Chem. Int. Ed. 1987, 26, 825–845. [Google Scholar] [CrossRef]
  47. Zeitler, K.; Neumann, M. Synergistic visible light photoredox catalysis. In Chemical Photocatalysis, 2nd ed.; König, B., Ed.; De Gruyter: Berlin, Germany, 2020; pp. 245–283. [Google Scholar]
  48. Skubi, K.L.; Blum, T.R.; Yoon, T.P. Dual Catalysis Strategies in Photochemical Synthesis. Chem. Rev. 2016, 116, 10035–10074. [Google Scholar] [CrossRef]
  49. Hopkinson, M.N.; Sahoo, B.; Li, J.-L.; Glorius, F. Dual Catalysis Sees the Light: Combining Photoredox with Organo-, Acid, and Transition-Metal Catalysis. Chem. Eur. J. 2014, 20, 3874–3886. [Google Scholar] [CrossRef]
  50. Ooi, T. Molecular Technology: Energy Innovation; Yamamoto, H., Kato, T., Eds.; Wiley-VCH: Weinheim, Germany, 2018; pp. 131–163. [Google Scholar]
  51. Mastandrea, M.M.; Pericàs, M. Photoredox Dual Catalysis: A Fertile Playground for the Discovery of New reactivities. Eur. J. Inorg. Chem. 2021, 2021, 3421–3431. [Google Scholar] [CrossRef]
  52. Lévêque, C.; Levernier, E.; Corcé, V.; Fensterbank, L.; Malacria, M.; Ollivier, C. Photoredox Catalysis, an Opportunity for Sustainable Radical Chemistry. In Avanced Green Chemistry, Part 2; Horváth, I.T., Malacria, M., Eds.; World Scientific: Singapore, 2020; pp. 49–121. [Google Scholar]
  53. Yang, N.; Tian, Y.; Zhang, M.; Peng, X.; Li, F.; Li, J.; Li, Y.; Fan, B.; Wang, F.; Song, H. Photocatalyst-enzyme hybrid systems for light-driven biotransformations. Biotechnol. Adv. 2022, 54, 107808. [Google Scholar] [CrossRef]
  54. Lee, S.H.; Choi, D.S.; Kuk, S.K.; Park, C.B. Photobiocatalysis: Activating Redox Enzymes by Direct or Indirect Transfer of Photoinduced Electrons. Angew. Chem. Int. Ed. 2018, 57, 7958–7985. [Google Scholar] [CrossRef]
  55. Rudorff, F.; Mihovilovic, M.D.; Gröger, H.; Snajdrova, R.; Iding, H.; Bornscheuer, U.T. Opportunities and challenges for combining chemo- and biocatalysis. Nat. Catal. 2018, 1, 12–22. [Google Scholar]
  56. Schmermund, L.; Jurkas, V.; Özgen, F.F.; Barone, G.D.; Büchsenschütz, H.C.; Winkler, C.K.; Schmidt, S.; Kourist, R.; Kroutil, W. Photo-Biocatalysis: Biotransformations in the Resence of Light. ACS Catal. 2019, 9, 4115–4144. [Google Scholar] [CrossRef]
  57. Zhang, W.; Hollmann, F. Noncoventional regeneration of redox enzymes—A practical approach for organic synthesis. Chem. Commun. 2018, 54, 7281–7289. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  58. Nicewicz, D.A.; MacMillan, D.W.C. Merging Photoredox Catalysis with Organocatalysis: The Direct Asymmetric Alkylation of Aldehydes. Science 2008, 322, 77–80. [Google Scholar] [CrossRef] [Green Version]
  59. Saha, D. Catalytic Enantioselective Radical Transformations Enabled by Visible Light. Chem. Asian J. 2020, 15, 2129–2152. [Google Scholar] [CrossRef] [PubMed]
  60. Brimioulle, R.; Lenhart, D.; Maturi, M.M.; Bach, T. Enantioselective Catalysis of Photochemical Reactions. Angew. Chem. Int. Ed. 2015, 54, 3872–3890. [Google Scholar] [CrossRef] [PubMed]
  61. Jiang, C.; Chen, W.; Zheng, W.-H.; Lu, H. Advances in asymmetric visible-light photocatalysis, 2015–2019. Org. Biomol. Chem. 2019, 17, 8673–8689. [Google Scholar] [CrossRef]
  62. Fidaly, K.; Ceballos, C.; Falguières, A.; Sylla-Iyarreta Veitia, M.; Guy, A.; Ferroud, C. Visible light photoredox organocatalysis: A fully transition metal-free direct asymmetric α-alkylation of aldehydes. Green Chem. 2012, 14, 1293–1297. [Google Scholar] [CrossRef]
  63. Cheung, K.P.S.; Sarkar, S.; Gevorgyan, V. Visible Light-Induced Transition Metal Catalysis. Chem. Rev. 2022, 122, 1543–1625. [Google Scholar] [CrossRef]
  64. Lipp, A.; Badir, S.O.; Molander, G.A. Stereoinduction in Metallaphotoredox Catalysis. Angew. Chem. Int. Ed. 2021, 60, 1714–1726. [Google Scholar] [CrossRef]
  65. Hoffmann, N. Combining Photoredox and Metal Catalysis. ChemCatChem 2015, 7, 393–394. [Google Scholar] [CrossRef]
  66. Fabry, D.C.; Rueping, M. Merging Visible Light Photoredox Catalysis with Metal Catalyzed C-H Activations: On the Role of Oxygen and Superoxide Ions as Oxidants. ACC Chem. Res. 2016, 49, 1969–1979. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  67. Sadek, O.; Abdellaoui, M.; Millanvois, A.; Ollivier, C.; Fensterbank, L. Organometallic catalysis under visible light activation: Benefits and preliminary rationals. Photochem. Photobiol. Sci. 2022, 21, 585–606. [Google Scholar] [CrossRef] [PubMed]
  68. Huo, H.; Shen, X.; Wang, C.; Zhang, L.; Röse, P.; Chen, L.A.; Harms, K.; Marsch, M.; Hilt, G.; Meggers, E. Asymmetric photoredox transition-metal catalysis ativated by visible light. Nature 2014, 515, 100–103. [Google Scholar] [CrossRef] [PubMed]
  69. Wu, X.; Luo, N.; Xie, S.; Zhang, H.; Zhang, Q.; Wang, F.; Wang, Y. Photocatalytic transformations of lignocellulosic biomass into chemicals. Chem. Soc. Rev. 2020, 49, 6198–6223. [Google Scholar] [CrossRef] [PubMed]
  70. Chen, H.; Wan, K.; Zheng, F.; Zhang, Z.; Zhang, H.; Zhang, Y.; Long, D. Recent Advances in Photocatalytic Transformation on Carbohydrates into Valuable Platform Chemicals. Front. Chem. Eng. 2021, 3, 615309. [Google Scholar] [CrossRef]
  71. Rao, V.N.; Malu, T.J.; Cheralathan, K.K.; Sakar, M.; Pitchaimuthu, S.; Rodríguez-González, V.; Kumari, M.M.; Shankar, M.V. Light-driven transformation of biomass into chemicals using photocatalysts—Vistas and challenges. J. Environ. Manag. 2021, 284, 111983. [Google Scholar]
  72. Colmenares, J.C.; Luque, R. Heterogeneous photocatalytic nanomaterials: Prospects and challenges in selective transformations of biomass-derived compounds. Chem. Soc. Rev. 2014, 43, 765–778. [Google Scholar] [CrossRef]
  73. Granone, L.I.; Sieland, F.; Zheng, N.; Dillert, R.; Bahnemann, D.W. Photocatalytic conversion of biomass into valuable products: A meaningful approach? Green Chem. 2018, 20, 1169–1192. [Google Scholar] [CrossRef]
  74. Fan, H.; Li, G.; Yang, F.; Yang, L.; Zhang, S. Photodegradation of cellulose under UV light catalysed by TiO2. J. Chem. Technol. Biotechnol. 2011, 86, 1107–1112. [Google Scholar] [CrossRef]
  75. Wang, L.; Zhang, Z.; Zhang, L.; Xue, S.; Doherty, W.O.S.; O’Hara, I.M.; Ke, X. Sustainable conversion of cellulosic biomass to chemicals under visible-light irradiation. RSC Adv. 2015, 5, 85242–85247. [Google Scholar] [CrossRef]
  76. Laugel, C.; Estrine, B.; Le Bras, J.; Hoffmann, N.; Marinkovic, S.; Muzart, J. Visible Light-Accelerated Depolymerisation of Starch under Fenton Conditions and Preparation of Calcium Sequestering Compounds. Catal. Lett. 2014, 144, 1674–1680. [Google Scholar] [CrossRef]
  77. Li, S.-H.; Liu, S.; Colmenares, J.C.; Xu, Y.-J. A sustainable approach for lignin valorization by heterogeneous photocatalysis. Green Chem. 2016, 18, 594–607. [Google Scholar] [CrossRef]
  78. Zakzeski, J.; Bruijnincx, P.C.A.; Jongerius, A.L.; Weckhuysen, B.M. The catalytic valorization of lignin for the production of renewable chemicals. Chem. Rev. 2010, 110, 3552–3599. [Google Scholar] [CrossRef] [PubMed]
  79. Rahimi, A.; Ulbrich, A.; Coon, J.J.; Stahl, S.S. Formic-acid-induced depolymerization of oxidized lignin to aromatics. Nature 2014, 515, 249–252. [Google Scholar] [CrossRef]
  80. Rinaldi, R.; Jastrzebski, R.; Clough, M.T.; Ralph, J.; Kennema, M.; Bruijnincx, P.C.A.; Weckhuysen, B.M. Paving the Way for Lignin Valorisation: Recent Advances in Bioengineering, Biorefining and Catalysis. Angew. Chem. Int. Ed. 2016, 55, 8164–8215. [Google Scholar] [CrossRef] [Green Version]
  81. Vanholme, R.; Demedts, B.; Morreel, K.; Ralph, J.; Boerjan, W. Lignin biosynthesis and structure. Plant Physiol. 2010, 153, 895–905. [Google Scholar] [CrossRef] [Green Version]
  82. Nayak, J.; Basu, A.; Dey, P.; Kumar, R.; Upadhaya, A.; Ghosh, S.; Bishayee, B.; Mishra, S.R.; Tripathy, S.K.; Banerjee, S.; et al. Transformation of agro-biomass into vanillin through novel membrane integrated value-addition process: A state-of-the-art review. Biomass Conv. Bioref. 2022, 1–24. [Google Scholar] [CrossRef]
  83. Kärkäs, M.D.; Matsuura, B.S.; Monos, T.M.; Magallanes, G.; Stephenson, C.R.J. Transition-metal catalyzed valorization of lignin: The key to a sustainable carbon-neutral future. Org. Biomol. Chem. 2016, 14, 1853–1914. [Google Scholar] [CrossRef] [Green Version]
  84. Magallanes, G.; Kärkäs, M.D.; Bosque, I.; Lee, S.; Maldonado, S.; Sephenson, C.R.J. Selective C–O Bond Cleavage of Lignin Systems and Polymers Enabled by Sequential Palladium-Catalyzed Aerobic Oxidation and Visible-Light Photoredox Catalysis. ACS Catal. 2019, 9, 2252–2260. [Google Scholar] [CrossRef]
  85. Muzart, J. Palladium-catalysed oxidation of primary and secondary alcohols. Tetrahedron 2003, 59, 5789–5816. [Google Scholar] [CrossRef]
  86. Hoffmann, N. Electron and hydrogen transfer in organic photochemical reactions. J. Phys. Org. Chem. 2015, 28, 121–136. [Google Scholar] [CrossRef]
  87. Zhang, J. Conversion of Lignin Models by Photoredox Catalysis. ChemSusChem 2018, 11, 3071–3080. [Google Scholar] [CrossRef] [PubMed]
  88. Das, A.; König, B. Transition metal- and photoredox-catalyzed valorisation of lignin subunits. Green Chem. 2018, 20, 4844–4852. [Google Scholar] [CrossRef]
  89. Shatskiy, A.; Kärkäs, M.D. Biomass Processing via Photochemical Means. In Biomass Valorisation; Ravelli, D., Samorì, C., Eds.; Wiley-VCH: Weinheim, Germany, 2021; pp. 265–288. [Google Scholar]
  90. Martín-Perales, A.I.; Rodríguez-Padron, D.; García Coleto, A.; Len, C.; de Miguel, G.; Muñoz-Batista, M.J.; Rafael Luque, R. Photocatalytic Production of Vanilline over CeOx and ZrO2 Modified Biomass-Templated Titania. Ind. Eng. Chem. Res. 2020, 59, 17085–17093. [Google Scholar] [CrossRef]
  91. Camera-Roda, G.; Augugliaro, V.; Cardillo, A.; Loddo, V.; Palmisano, G.; Palmisano, L. A pervaporation reactor for the green Synthesis of Vanillin. Chem. Eng. J. 2013, 224, 136–143. [Google Scholar] [CrossRef] [Green Version]
  92. Desvals, A.; Baudron, S.A.; Bulach, V.; Hoffmann, N. Photocycloadditions of Arenes Derived from Lignin. J. Org. Chem. 2021, 86, 13310–13321. [Google Scholar] [CrossRef]
  93. Desvals, A.; Hoffmann, N. Photocycloadditions of benzene derivatives and their systematic application to organic synthesis. Aus. J. Chem. 2023, 76, 117–129. [Google Scholar] [CrossRef]
  94. Remy, R.; Bochet, C.G. Arene-Alkene Cycloaddition. Chem. Rev. 2016, 116, 9816–9849. [Google Scholar] [CrossRef]
  95. Chapleur, Y.; Chrétien, F. Sugars as Chiral Starting Materials in Enantiospecific Synthesis. In The Organic Chemistry of Sugars; Levy, D.E., Fügedi, P., Eds.; CRC Taylor & Francis: Boca Raton, FL, USA, 2006; pp. 489–573. [Google Scholar]
  96. Vogel, P. Total Asymmetric Synthesis of Monosaccharides and Analogs. In The Organic Chemistry of Sugars; Levy, D.E., Fügedi, P., Eds.; CRC Taylor & Francis: Boca Raton, FL, USA, 2006; pp. 629–725. [Google Scholar]
  97. Lichtenthaler, F.W. Emil Fischer’s Proof of the Configuration of Sugars: A Centennial Tribute. Angew. Chem. Int. Ed. 1992, 31, 1541–1556. [Google Scholar] [CrossRef]
  98. Stick, R.V. Carbohydrates—The Sweet Molecules of Life; Academic Press: San Diego, CA, USA, 2001. [Google Scholar]
  99. Toshima, K. Synthesis of Carbohydrate Containing Complex Natural Compounds. In The Organic Chemistry of Sugars; Levy, D.E., Fügedi, P., Eds.; CRC Taylor & Francis: Boca Raton, FL, USA, 2006; pp. 575–627. [Google Scholar]
  100. Kuszmann, J. Ingtoduction to Carbohydrates. In The Organic Chemistry of Sugars; Levy, D.E., Fügedi, P., Eds.; CRC Taylor & Francis: Boca Raton, FL, USA, 2006; pp. 25–52. [Google Scholar]
  101. Shatskiy, A.; Stepanova, E.V.; Kärkäs, M.D. Exploiting photoredox catalysis for carbohydrate modification through C-H and C-C activation. Nat. Rev. Chem. 2022, 6, 782–805. [Google Scholar] [CrossRef]
  102. Gassama, A.; Hoffmann, N. Selective Light Supported Oxidation of Hexoses Using Air as Oxidant—Synthesis of Tetrahydroxyazepanes. Adv. Synth. Catal. 2008, 350, 35–39. [Google Scholar] [CrossRef]
  103. Ito, N.; Phillips, S.E.V.; Stevens, C.; Ogel, Z.B.; McPhersonn, M.J.; Keen, J.N.; Yadav, K.D.S.; Knowles, P.F. Novel thioether bond revealed by a 1.7 Å crystal structure of galactose oxidase. Nature 1991, 350, 87–90. [Google Scholar] [CrossRef] [PubMed]
  104. Thomas, F. Ten Years of a Biomimetic Approach to the Copper(II) Radical Site of Galactose Oxidase. Eur. J. Inorg. Chem. 2007, 2007, 2379–2404. [Google Scholar] [CrossRef]
  105. Pierre, J.-L. One electron at a time oxidations and enzymatic paradigms: From metallic to non-metallic redox centers. Chem. Soc. Rev. 2000, 29, 251–257. [Google Scholar] [CrossRef]
  106. Berkessel, A.; Dousset, M.; Bulat, S.; Glaubitz, K. Combinatorial approaches to functional models for galactose oxidase. Biol. Chem. 2005, 386, 1035–1041. [Google Scholar] [CrossRef] [PubMed]
  107. Chaudhuri, P.; Hess, M.; Hildenbrand, K.; Bill, E.; Weyhermüller, T.; Wieghardt, K. Aerobic Oxidation of Primary Alcohols (Including Methanol) by Copper(II)− and Zinc(II)−Phenoxyl Radical Catalysts. J. Am. Chem. Soc. 1999, 121, 9599–9610. [Google Scholar] [CrossRef]
  108. Semmelhack, M.F.; Schmid, C.R.; Cortés, D.A.; Chou, C.S. Oxidation of alcohols to aldehydes with oxygen and cupric ion, mediated by nitrosonium ion. J. Am. Chem. Soc. 1984, 106, 3374–3376. [Google Scholar] [CrossRef]
  109. Ryland, B.L.; Stahl, S.S. Practical Aerobic Oxidations of Alcohols and Amines with Homogeneous Copper/TEMPO and Related Catalyst Systems. Angew. Chem. Int. Ed. 2014, 53, 8824–8838. [Google Scholar] [CrossRef] [Green Version]
  110. Dijksman, A.; Arends, I.W.C.E.; Sheldon, R.A. Cu(ii)-nitroxyl radicals as catalytic galactose oxidase mimics. Org. Biomol. Chem. 2003, 1, 3232–3237. [Google Scholar] [CrossRef]
  111. Abderrazak, Y.; Bhattacharyya, A.; Reiser, O. Visible-Light-Induced Homolysis of Earth-Abundant Metal-Substrate Complexes: A Complementary Activation Strategy in Photoredox Catalysis. Angew. Chem. Int. Ed. 2021, 60, 21100–21115. [Google Scholar] [CrossRef]
  112. Compain, P.; Martin, O.R. (Eds.) Iminosugars; Wiley & Sons: Chichester, UK, 2007. [Google Scholar]
  113. Li, H.; Marcelo, F.; Bello, C.; Vogel, P.; Butters, T.D.; Rauter, A.P.; Zhang, Y.; Sollogoub, M.; Blériot, Y. Design and synthesis of acetamido tri- and tetra-hydroxyazepanes: Potent and selective β-N-acetylhexosaminidase inhibitors. Bioorg. Med. Chem. 2009, 17, 5598–5604. [Google Scholar] [CrossRef] [PubMed]
  114. Joseph, C.C.; Regeling, H.; Zwanenburg, B.; Chittenden, G.J.F. Syntheses of (3R,4R,5R,6R)-tetrahydroxyazepane (1,6-dideoxy-1,6-imino-d-mannitol) and (3S,4R,5R,6R)-tetrahydroxyazepane (1,6-dideoxy-1,6-imino-d-glucitol). Tetrahedron 2002, 58, 6907–6911. [Google Scholar] [CrossRef]
  115. Oelgemöller, M. Solar Photochemical Synthesis: From the Beginnings of Organic Photochemistry to the Solar Manufacturing of Commodity Chemicals. Chem. Rev. 2016, 116, 9664–9682. [Google Scholar] [CrossRef]
  116. Sangwan, R.; Mandal, P.K. Recent advances in photoinduced glycosylation: Oligosaccharides, glycoconjugates and their synthetic applications. RSC Adv. 2017, 7, 26256–26321. [Google Scholar] [CrossRef] [Green Version]
  117. Ling, J.; Bennett, C.S. Recent Developments in Stereoselective Chemical Glycosylation. Asian J. Org. Chem. 2019, 8, 802–813. [Google Scholar] [CrossRef] [PubMed]
  118. Fügedi, P. Glycosylation Methods. In The Organic Chemistry of Sugars; Levy, D.E., Fügedi, P., Eds.; CRC Taylor & Francis: Boca Raton, FL, USA, 2006; pp. 89–179. [Google Scholar]
  119. Collins, P.; Ferrier, R. Monosaccharides; Wiley: Chichester, UK, 1995. [Google Scholar]
  120. Hashimoto, S.; Kurimoto, I.; Fujii, Y.; Noyori, R. Novel nucleophilic substitution reaction by radical cation intermediates. Photosensitized transacetalization via SON1 mechanism. J. Am. Chem. Soc. 1985, 107, 1427–1429. [Google Scholar] [CrossRef]
  121. Wever, W.J.; Cinelli, M.A.; Bowers, A.A. Visible Light Mediated Activation and O-Glycosylation of Thioglycosides. Org. Lett. 2013, 15, 30–33. [Google Scholar] [CrossRef]
  122. Tolbert, L.M.; Solntsev, K.M. Excited-State Proton Transfer:  From Constrained Systems to “Super” Photoacids to Superfast Proton Transfer. ACC Chem. Res. 2002, 35, 19–27. [Google Scholar] [CrossRef]
  123. Zivic, N.; Kuroishi, P.K.; Dumur, F.; Gigmes, D.; Dove, A.P.; Sardon, H. Recent Advances and Challenges in the Design of Organic Photoacid and Photobase Generators for Polymerizations. Angew. Chem. Int. Ed. 2019, 58, 10410–10422. [Google Scholar] [CrossRef]
  124. Saway, J.; Salem, Z.M.; Badillo, J.J. Recent Advances in Photoacid Catalysis for Organic Synthesis. Synthesis 2021, 53, 489–497. [Google Scholar]
  125. Iwata, R.; Uda, K.; Takahashi, D.; Toshima, K. Photo-induced glycosylation using reusable organophotoacids. Chem. Commun. 2014, 50, 10695–10698. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  126. Kimura, T.; Eto, T.; Takahashi, D.; Toshima, K. Stereocontrolled Photoinduced Glycosylation Using an Aryl Thiourea as an Organo photoacid. Org. Lett. 2016, 18, 3190–3193. [Google Scholar] [CrossRef] [PubMed]
  127. Zhao, G.; Li, J.; Wang, T. Visible-light-induced photoacid catalysis: Application in glycosylation with O-glycosyl trichloroacetimidates. Chem. Commun. 2021, 57, 12659–12662. [Google Scholar] [CrossRef] [PubMed]
  128. Levy, D.E. Strategies towards C-Glycosides. In The Organic Chemistry of Sugars; Levy, D.E., Fügedi, P., Eds.; CRC Taylor & Francis: Boca Raton, FL, USA, 2006; pp. 269–348. [Google Scholar]
  129. Zhang, M.; Kong, L.; Gong, R.; Iorio, M.; Donatio, S.; Deng, Z.; Sosio, M.; Chen, W. Biosynthesis of C-nucleoside antibiotics in actinobacteria: Recent advances and future developments. Microbiol. Cell Factories 2022, 21, 2. [Google Scholar] [CrossRef]
  130. Zhu, M.; Messoudi, S. Diastereoselective Decarboxylative Alkynylation of Anomeric Carboxylic Acids Using Cu/Photoredox Dual Catalysis. ACS Catal. 2021, 11, 6334–6342. [Google Scholar] [CrossRef]
  131. Schwarz, J.; König, B. Decarboxylative reactions with and without light—A comparison. Green Chem. 2018, 20, 323–361. [Google Scholar] [CrossRef]
  132. Martinez, R.D.; Buitrago, A.A.; Howell, N.W.; Hearn, C.H.; Joens, J.A. The near UV absorption spectra of several aliphatic aldehydes and ketones at 300 K. Atmos. Environ. 1992, 26A, 785–792. [Google Scholar] [CrossRef]
  133. Jahjah, R.; Gassama, A.; Bulach, V.; Suzuki, C.; Abe, M.; Hoffmann, N.; Martinez, A.; Nuzillard, J.-M. Stereoselective Triplet-Sensitised Radical Reactions of Furanone Derivatives. Chem. Eur. J. 2010, 16, 3341–3354. [Google Scholar] [CrossRef]
  134. Oelgemöller, M.; Hoffmann, N.; Shvydkiv, O. From ‘Lab & Light on a Chip’ to Parallel Microflow Photochemistry. Aust. J. Chem. 2014, 67, 337–342. [Google Scholar]
  135. Hoffmann, N. Photochemical Electron and Hydrogen Transfer in Organic Synthesis: The Control of Selectivity. Synthesis 2016, 48, 1782–1802. [Google Scholar] [CrossRef]
  136. Wessig, P.; Mühling, O. Abstraction of (γ ± n)-Hydrogen by Excited Carbonyls. In Synthetic Organic Photochemistry; Giesbeck, A.G., Mattay, J., Eds.; Marcel Dekker: New York, NY, USA, 2005; pp. 41–87. [Google Scholar]
  137. Martín, A.; Suárez, E. Carbohydrate Spiro-heterocycles via Radical Chemistry. Top. Heterocycl. Chem. 2019, 57, 51–104. [Google Scholar]
  138. Thiering, S.; Sund, C.; Thiem, J.; Giesler, A.; Kopf, J. Syntheses of imido-substituted glycosans and their photocyclisation towards highly functionalised heterotricycles. Carbohydr. Res. 2001, 336, 271–282. [Google Scholar] [CrossRef] [PubMed]
  139. Capaldo, L.; Ravelli, D. Hydrogen Atom Transfer (HAT): A Versatile Strategy for Substrate Activation in Photocatalyzed Organic Synthesis. Eur. J. Org. Chem. 2017, 2017, 2056–2071. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  140. Ravelli, D.; Protti, S.; Fagnoni, M. Carbon–Carbon Bond Forming Reactions via Photogenerated Intermediates. Chem. Rev. 2016, 116, 9850–9913. [Google Scholar] [CrossRef]
  141. Fréneau, M.; de Sainte-Claire, P.; Abe, M.; Hoffmann, N. The structure of electronically excited α,β-unsaturated lactones. J. Phys. Org. Chem. 2016, 29, 718–724. [Google Scholar] [CrossRef]
  142. Lejeune, G.; Font, J.; Parella, T.; Alibés, R.; Figueredo, M. Intramolecular Photoreactions of (5S)-5-Oxymethyl-2(5H)-furanones as a Tool for the Stereoselective Generation of Diverse Polycyclic Scaffolds. J. Org. Chem. 2015, 80, 9437–9445. [Google Scholar] [CrossRef]
  143. Tabanelli, T.; Cavani, F. Industrial Perspectives of Biomass Processing. In Biomass Valorisation; Ravelli, D., Samorì, C., Eds.; Wiley-VCH: Weinheim, Germany, 2021; pp. 369–410. [Google Scholar]
  144. Welter, R.A.; Santana, H.; de la Torre, L.G.; Robertson, M.; Taranto, O.P.; Oelgemöller, M. Methyl Oleate Synthesis by TiO2 Photocatalytic Esterification of Oleic Acid: Optimisation by Response Surface Quadratic Methodology, Reaction Kinetics and Thermodynamics. ChemPhotoChem 2022, 6, e202200007. [Google Scholar] [CrossRef]
  145. Huang, J.; Jian, Y.; Zhu, P.; Abdelaziz, O.; Li, H. Research Progress on the Photo-Driven Catalytic Production of Biodiesel. Front. Chem. 2022, 10, 904251. [Google Scholar] [CrossRef]
  146. Belousov, A.S.; Suleimanov, E.V. Application of metal–organic frameworks as an alternative to metal oxide-based photocatalysts for the production of industrially important organic chemicals. Green Chem. 2021, 23, 6172–6204. [Google Scholar] [CrossRef]
  147. Liu, P.; Zhang, G.; Sun, P. Transition metal-free decarboxylative alkylation reactions. Org. Biomol. Chem. 2016, 14, 10763–10777. [Google Scholar] [CrossRef]
  148. Schwarz, J. Photocatalytic decarboxylations. Phys. Sci. Rev. 2018, 3, 20170186. [Google Scholar] [CrossRef]
  149. Tabandeh, M.; Cheng, C.K.; Centi, G.; Show, P.L.; Chen, W.-H.; Ling, T.C.; Ong, H.C.; Ng, E.-P.; Juan, J.C.; Lam, S.S. Recent advancement in deoxygenation of fatty acids via homogeneous catalysis for biofuel production. Mol. Catal. 2022, 523, 111207. [Google Scholar] [CrossRef]
  150. Patra, T.; Maiti, D. Decarboxylation as the Key Step in C−C Bond-Forming Reactions. Chem. Eur. J. 2017, 23, 7382–7401. [Google Scholar] [CrossRef]
  151. Griesbeck, A.G.; Kramer, W.; Oelgemöller, M. Photoinduced decarboxylation reactions. Radical chemistry in water. Green Chem. 1999, 1, 205–208. [Google Scholar] [CrossRef]
  152. Griesbeck, A.G.; Hoffmann, N.; Warzecha, K.-D. Photoinduced-Electron-Transfer Chemistry: From Studies on PET Processes to Applications in Natural Product Synthesis. ACC Chem. Res. 2007, 40, 128–140. [Google Scholar] [CrossRef] [PubMed]
  153. Maeda, K.; Saito, H.; Osaka, K.; Nishikawa, K.; Sugie, M.; Morita, T.; Takahashi, I.; Yoshimi, Y. Direct modification of tripeptides using photoinduced decarboxylative radical reactions. Tetrahedron 2015, 71, 1117–1123. [Google Scholar] [CrossRef]
  154. Kicatt, D.M.; Nicolle, S.; Lee, A.-L. Direct decarboxylative Giese reactions. Chem. Soc. Rev. 2022, 51, 1415–1453. [Google Scholar] [CrossRef]
  155. Rodríguez, N.; Goossen, L.J. Decarboxylative coupling reactions: A modern strategy for C–C-bond formation. Chem. Soc. Rev. 2011, 40, 5030–5048. [Google Scholar] [CrossRef] [Green Version]
  156. McMurray, L.; McGuire, T.M.; Howells, R.L. Recent Advances in Photocatalytic Decarboxylative Coupling Reactions in Medicinal Chemistry. Synthesis 2020, 52, 1719–1737. [Google Scholar] [CrossRef]
  157. Yoshimi, Y.; Itou, T.; Hatanaka, M. Decarboxylative reduction of free aliphatic carboxylic acids by photogenerated cation radical. Chem. Comm. 2007, 5244–5246. [Google Scholar] [CrossRef]
  158. Griffin, J.D.; Zeller, M.A.; Nicewicz, D.A. Hydrodecarboxylation of Carboxylic and Malonic Acid Derivatives via Organic Photoredox Catalysis: Substrate Scope and Mechanistic Insight. J. Am. Chem. Soc. 2015, 137, 11340–11348. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  159. Huang, Z.; Zhao, Z.; Zhang, C.; Liu, J.; Luo, N.; Zhang, J.; Wang, F. Enhanced photocatalytic alkane production from fatty acid decarboxylation via inhibition of radical oligomerization. Nat. Catal. 2020, 3, 170–178. [Google Scholar] [CrossRef]
  160. Sorigué, D.; Légeret, B.; Cuiné, S.; Blangy, S.; Moulin, S.; Billon, E.; Richaud, P.; Brugière, S.; Couté, Y.; Nurizzo, D.; et al. An algal photoenzyme converts fatty acids to hydrocarbons. Science 2017, 357, 903–907. [Google Scholar] [CrossRef] [Green Version]
  161. Sorigué, D.; Hadjidemetriou, K.; Blangy, S.; Gotthard, G.; Bonvalet, A.; Coquelle, N.; Samire, P.; Aleksandrov, A.; Antonucci, L.; Benachir, A.; et al. Mechanism and dynamics of fatty acid photodecarboxylase. Science 2021, 372, eabd5687. [Google Scholar] [CrossRef]
  162. Harrison, W.; Huang, X.; Zhao, H. Photobiocatalysis for Abiological Transformations. Acc. Chem. Res. 2022, 55, 1087–1096. [Google Scholar] [CrossRef] [PubMed]
  163. Amer, M.; Wojcik, E.Z.; Sun, C.; Hoeven, R.; Hughes, J.M.X.; Faulkner, M.; Yunus, I.S.; Tait, S.; Johannissen, L.O.; Hardman, S.J.O.; et al. Low carbon strategies for sustainable bio-alkane gas production and renewable energy. Energy Environ. Sci. 2020, 13, 1818–1831. [Google Scholar] [CrossRef]
  164. Zhang, W.; Ma, M.; Huijbers, M.M.E.; Filonenko, G.A.; Pidko, E.A.; van Schie, M.; de Boer, S.; Burek, B.O.; Bloh, J.Z.; van Berkel, W.J.H.; et al. Hydrocarbon Synthesis via Photoenzymatic Decarboxylation of Carboxylic Acids. J. Am. Chem. Soc. 2019, 141, 3116–3120. [Google Scholar] [CrossRef] [Green Version]
  165. Duong, H.T.; Wu, Y.; Sutor, A.; Burek, B.O.; Hollmann, F.; Bloh, J.Z. Intensification of Photobiocatalytic Decarboxylation of Fatty Acids for the Production of Biodiesel. ChemSusChem 2021, 14, 1053–1056. [Google Scholar] [CrossRef]
  166. Griesbeck, A.G.; Kramer, W.; Oelgemöller, M. Synthetic Applications of Photoinduced Electron Transfer Decarboxylation Reactions. Synlett 1999, 1169–1178. [Google Scholar] [CrossRef] [Green Version]
  167. Xuan, J.; Zhang, Z.G.; Xiao, W.J. Visible-Light-Induced Decarboxylative Functionalization of Carboxylic Acids and Their Derivatives. Angew. Chem. Int. Ed. 2015, 54, 15632–15641. [Google Scholar] [CrossRef]
  168. Zheng, Y.; Shao, X.; Ramadoss, V.; Tian, L.; Wang, Y. Recent Developments in Photochemical and Electrochemical Decarboxylative C(sp3)–N Bond Formation. Synthesis 2020, 52, 1357–1368. [Google Scholar]
  169. Mumtaz, S.; Robertson, M.J.; Oelgemöller, M. Recent advances in photodecarboxylations involving phthalimides. Aust. J. Chem. 2018, 71, 634–648. [Google Scholar] [CrossRef]
  170. Jin, Y.; Fu, H. Visible-Light Photoredox Decarboxylative Couplings. Asian J. Org. Chem. 2017, 6, 368–385. [Google Scholar] [CrossRef]
  171. Schwarz, J.; König, B. Decarboxylative Alkynylation of Biomass-Derived Compounds by Metal-Free Visible Light Photocatalysis. ChemPhotoChem 2017, 1, 237–242. [Google Scholar] [CrossRef]
  172. Garza-Sanchez, R.A.; Tlahuext-Aca, A.; Tavakoli, G.; Glorius, F. Visible Light-Mediated Direct Decarboxylative C–H Functionalization of Heteroarenes. ACS Catal. 2017, 6, 4057–4061. [Google Scholar] [CrossRef]
  173. Hu, D.; Wang, L.; Li, P. Decarboxylative Borylation of Aliphatic Esters under Visible-Light Photoredox Conditions. Org. Lett. 2017, 19, 2770–2773. [Google Scholar] [CrossRef]
  174. Fawcett, A.; Pradeilles, J.; Wang, Y.; Mutsuga, T.; Myers, E.L.; Aggarwal, V.K. Photoinduced decarboxylative borylation of carboxylic acids. Science 2017, 357, 283–286. [Google Scholar] [CrossRef] [Green Version]
  175. Tlahuext-Aca, A.; Candish, L.; Garza-Sanchez, R.A.; Glorius, F. Decarboxylative Olefination of Activated Aliphatic Acids Enabled by Dual Organophotoredox/Copper Catalysis. ACS Catal. 2018, 8, 1715–1719. [Google Scholar] [CrossRef]
  176. Xie, J.; Xu, P.; Li, H.; Xue, Q.; Jin, H.; Cheng, Y.; Zhu, C. A room temperature decarboxylation/C–H functionalization cascade by visible-light photoredox catalysis. Chem. Commun. 2013, 49, 5672–5674. [Google Scholar] [CrossRef]
  177. De bruyn, M.; Fan, J.; Budarin, V.L.; Macquarrie, D.J.; Gomez, L.D.; Simister, R.; Farmer, T.J.; Raverty, W.D.; McQueen-Mason, S.J.; Clark, J.H. A new perspective in bio-refining: Levoglucosenone and cleaner lignin from waste biorefinery hydrolysis lignin by selective conversion of residual saccharides. Energy Environ. Sci. 2016, 9, 2571–2574. [Google Scholar] [CrossRef] [Green Version]
  178. Halpern, Y.; Ritter, R.; Broido, A. Levoglucosenone (1,6-anhydro-3,4-dideoxy-.DELTA.3-.beta.-D-pyranosen-2-one). Major product of the acid-catalyzed pyrolysis of cellulose and related carbohydrates. J. Org. Chem. 1973, 38, 204–209. [Google Scholar] [CrossRef]
  179. He, J.; Liu, M.; Huang, K.; Walker, T.W.; Maravelias, C.T.; Dumesic, J.A.; Huber, G.W. Production of levoglucosenone and 5-hydroxymethylfurfural from cellulose in polar aprotic solvent–water mixtures. Green Chem. 2017, 19, 3642–3653. [Google Scholar] [CrossRef]
  180. Sherwood, J.; De bruyn, M.; Constantinou, A.; Moity, L.; McElroy, C.R.; Farmer, T.J.; Duncan, T.; Raverty, W.; Hunt, A.J.; Clark, J.H. Dihydrolevoglucosenone (Cyrene) as a bio-based alternative for dipolar aprotic solvents. Chem. Commun. 2014, 50, 9650–9652. [Google Scholar] [CrossRef]
  181. Comba, M.B.; Tsai, Y.; Sarotti, A.M.; Mangione, M.I.; Suárez, A.G.; Spanevello, R.A. Levoglucosenone and Its New Applications: Valorization of Cellulose Residues. Eur. J. Org. Chem. 2018, 2018, 590–604. [Google Scholar] [CrossRef]
  182. Awad, L.; Demange, R.; Zhu, Y.-H.; Vogel, P. The use of levoglucosenone and isolevoglucosenone as templates for the construction of C-linked disaccharides. Carbohydr. Res. 2006, 341, 1235–1252. [Google Scholar] [CrossRef] [PubMed]
  183. Sarotti, A.M.; Zanardi, M.M.; Spanevello, R.A.; Suárez, A.G. Recent Applications of Levoglucosenone as Chiral Synthon. Curr. Org. Synth. 2012, 9, 439–459. [Google Scholar] [CrossRef]
  184. Tsai, Y.; Borini Etichetti, C.M.; Di Benedetto, C.; Girardini, J.E.; Terra Martins, F.; Spanevello, R.A.; Suárez, A.; Sarotti, A.M. Synthesis of Triazole Derivatives of Levoglucosenone as Promising Anticancer Agents: Effective Exploration of the Chemical Space through retro-aza-Michael//aza-Michael Isomerizations. J. Org. Chem. 2018, 83, 3516–3528. [Google Scholar] [CrossRef]
  185. Camp, J.E.; Greatrex, B.W. Levoglucosenone: Bio-Based Platform for Drug Discovery. Front. Chem. 2022, 10, 902239. [Google Scholar] [CrossRef]
  186. Diot-Néant, F.; Rastoder, E.; Miller, S.A.; Allais, F. Chemo-Enzymatic Synthesis and Free Radical Polymerization of Renewable Acrylate Monomers from Cellulose-Based Lactones. ACS Sustain. Chem. Eng. 2018, 6, 17284–17293. [Google Scholar] [CrossRef]
  187. Bonneau, G.; Peru, A.A.M.; Flourat, A.L.; Allais, F. Organic solvent- and catalyst-free Baeyer–Villiger oxidation of levoglucosenone and dihydrolevoglucosenone (Cyrene®): A sustainable route to (S)-γ-hydroxymethyl-α,β-butenolide and (S)-γ-hydroxymethyl-γ-butyrolactone. Green Chem. 2018, 20, 2455–2458. [Google Scholar] [CrossRef]
  188. Lefebvre, C.; Van Gysel, T.; Michelin, C.; Rousset, E.; Djiré, D.; Allais, F.; Hoffmann, N. Photocatalytic Radical Addition to Levoglucosenone. Eur. J. Org. Chem. 2022, 2022, e202101298. [Google Scholar] [CrossRef]
  189. Raviola, C.; Protti, S.; Ravelli, D.; Fagnoni, M. Photogenerated acyl/alkoxycarbonyl/carbamoyl radicals for sustainable synthesis. Green Chem. 2019, 21, 748–764. [Google Scholar] [CrossRef]
  190. Taniellan, C. Decatungstate photocatalysis. Coord. Chem. Rev. 1998, 178–180, 1165–1181. [Google Scholar] [CrossRef]
  191. Ravelli, D.; Protti, S.; Fagnoni, M. Decatungstate Anion for Photocatalyzed “Window Ledge” Reactions. Acc. Chem. Res. 2016, 49, 2232–2242. [Google Scholar] [CrossRef] [PubMed]
  192. Yamada, K.; Fukuyama, T.; Fujii, S.; Ravelli, D.; Fagnoni, M.; Ryu, I. Cooperative Polar/Steric Strategy in Achieving Site-Selective Photocatalyzed C(sp3)−H Functionalization. Chem. Eur. J. 2017, 23, 8615–8618. [Google Scholar] [CrossRef]
  193. Ravelli, D.; Fagnoni, M.; Fukuyama, T.; Nishikawa, T.; Ryu, I. Site-Selective C–H Functionalization by Decatungstate Anion Photocatalysis: Synergistic Control by Polar and Steric Effects Expands the Reaction Scope. ACS Catal. 2018, 8, 701–713. [Google Scholar] [CrossRef] [Green Version]
  194. Mika, L.T.; Cséfalvay, E. Conversion of Carbohydrates to Chemicals. In Avanced Green Chemistry, Part 1; Horváth, I.T., Malacria, M., Eds.; World Scientific: Singapore, 2020; pp. 19–76. [Google Scholar]
  195. Zhu, J.; Yin, G. Catalytic Transformation of the Furfural Platform into Bifunctionalized Monomers for Polymer Synthesis. ACS Catal. 2021, 11, 10058–10083. [Google Scholar] [CrossRef]
  196. Martel, F.; Estrine, B.; Plantier-Royon, R.; Hoffmann, N.; Portella, C. Development of Agriculture Left-Overs: Fine Organic Chemicals from Wheat Hemicellulose-Derived Pentoses. Top. Curr. Chem. 2010, 294, 79–115. [Google Scholar]
  197. Bergman, J.A.; Kessler, M.R. Monomers and Resulting Polymers from Biomass. In Introduction to Chemicals from Biomass, 2nd ed.; Clark, J., Deswarte, F., Eds.; Wiley: Chichester, UK, 2015; pp. 157–204. [Google Scholar]
  198. Yue, X.; Queneau, Y. 5-Hydroxymethylfurfural Chemistry toward Biobased Surfactants. ChemSusChem 2022, 15, e202102660. [Google Scholar] [CrossRef]
  199. Velty, A.; Iborra, S.; Corma, A. Synthetic Routes for Designing Furanic and Non Furanic Biobased Surfactants from 5-Hydroxymethylfurfural. ChemSusChem 2022, 15, e202200181. [Google Scholar] [CrossRef]
  200. Girka, Q.; Hausser, N.; Estrine, B.; Hoffmann, N.; Le Bras, J.; Marinković, S.; Muzart, J. β-Amino acid derived gemini surfactants from diformylfuran (DFF) with particularly low critical micelle concentration (CMC). Green Chem. 2017, 19, 4074–4079. [Google Scholar] [CrossRef]
  201. Minaev, B.F. Electronic mechanisms of activation of molecular oxygen. Russ. Chem. Rev. 2007, 76, 988–1010. [Google Scholar] [CrossRef]
  202. Schweitzer, C.; Schmidt, R. Physical Mechanisms of Generation and Deactivation of Singlet Oxygen. Chem. Rev. 2003, 103, 1685–1757. [Google Scholar] [CrossRef] [PubMed]
  203. Espada, J. (Ed.) Reactive Oxygen Species; Springer Science + Business Media: Berlin, Germany, 2021. [Google Scholar]
  204. Krumova, K.; Cosa, G. Overview of Reactive Oxygen Species. In Singlet Oxygen, Applications in Biosciences and Nanosciences; Nonell, S., Flors, C., Eds.; The Royal Society of Chemistry: Cambridge, UK, 2016; Volume 1, pp. 3–21. [Google Scholar]
  205. Boix-Garriga, E.; Rodríguez-Amigo, B.; Planas, O.; Nonell, S. Properties of singelet oxygen. In Singlet Oxygen, Applications in Biosciences and Nanosciences; Nonell, S., Flors, C., Eds.; The Royal Society of Chemistry: Cambridge, UK, 2016; Volume 1, pp. 23–46. [Google Scholar]
  206. Gellé, A.; Price, G.D.; Voisard, F.; Brodusch, N.; Gauvin, R.; Amara, Z. Enhancing Singlet Oxygen Photocatalysis with Plasmonic Nanoparticles. ACS Appl. Mater. Interfaces 2021, 13, 35606–35616. [Google Scholar] [CrossRef]
  207. Tambosco, B.; Segura, K.; Seyrig, C.; Cabrera, D.; Port, M.; Ferroud, C.; Amara, Z. Outer-sphere effects in visible-light photochemical oxidations with immobilized and recyclable ruthenium bipyridyl salts. ACS Catal. 2018, 8, 4383–4389. [Google Scholar] [CrossRef]
  208. Ghogare, A.A.; Greer, A. Using Singlet Oxygen to Synthesize Natural Products and Drugs. Chem. Rev. 2016, 116, 9994–10034. [Google Scholar] [CrossRef]
  209. Turconi, J.; Griolet, F.; Guevel, R.; Oddon, G.; Villa, R.; Geatti, A.; Hvala, M.; Rossen, K.; Göller, R.; Burgard, A. Semisynthetic Artemisinin, the Chemical Path to Industrial Production. Org. Process Res. Dev. 2014, 18, 417–422. [Google Scholar] [CrossRef]
  210. Turconi, J.; Mackievicz, P. Paludisme et hémisynthèse industrielle de l’artémisinine: Du rêve à la réalité! Actual. Chim. 2018, 39–47. [Google Scholar]
  211. Tu, Y. Artemisinin—A Gift from Traditional Chinese Medicine to the World. Angew. Chem. Int. Ed. 2016, 55, 10210–10226. [Google Scholar] [CrossRef]
  212. Griesbeck, A.G.; Höinck, L.-O.; Neudörfl, J.M. Synthesis of spiroannulated and 3-arylated 1,2,4-trioxanes from mesitylol and methyl 4-hydroxytiglate by photooxygenation and peroxyacetalization. Beilstein J. Org. Chem. 2010, 6, 61. [Google Scholar] [CrossRef] [Green Version]
  213. Griesbeck, A.G.; El-Idreesy, T.T.; Fiege, M.; Brun, R. Synthesis of Antimalarial 1,2,4-Trioxanes via Photooxygenation of a Chiral Allylic Alcohol. Org. Lett. 2002, 4, 4193–4195. [Google Scholar] [CrossRef]
  214. Singh, C.; Varma, V.P.; Naikade, N.K.; Singh, A.S.; Hassam, M.; Pun, S.K. Novel Bis- and Tris-1,2,4-trioxanes: Synthesis and Antimalarial Activity against Multidrug-Resistant Plasmodium yoelii in Swiss Mice. J. Med. Chem. 2008, 51, 7581–7592. [Google Scholar] [CrossRef] [PubMed]
  215. Tsolakis, N.; Bam, W.; Srai, J.S.; Kumar, M. Renewable chemical feedstock supply network design: The case of terpenes. J. Cleaner Prod. 2019, 222, 802–822. [Google Scholar] [CrossRef]
  216. Behr, A.; Johnen, L. Myrcene as a Natural Base Chemical in Sustainable Chemistry: A Critical Review. ChemSusChem 2009, 2, 1072–1095. [Google Scholar] [CrossRef] [PubMed]
  217. Monica, F.D.; Kleij, A.W. From terpenes to sustainable and functional polymers. Polym. Chem. 2020, 11, 5109–5127. [Google Scholar] [CrossRef]
  218. Mewalal, R.; Rai, D.K.; Kainer, D.; Chen, F.; Külheim, C.; Peter, G.F.; Tuskan, G.A. Plant-Derived Terpenes: A Feedstock for Specialty Biofuels. Trends Biotechnol. 2017, 35, 227–240. [Google Scholar] [CrossRef] [Green Version]
  219. Rojahn, W.; Warnecke, H.U. Die photosensibilisierte Sauerstoffübertragung–eine Methode zur Herstellung hochwertiger Riechstoffe. Dragoco Report 1980, 27, 159–164. [Google Scholar]
  220. Surburg, H.; Panten, J. (Eds.) Common Fragrance and Flavor Materials, 6th ed.; Wiley-VCH: Weinheim, Germany, 2016. [Google Scholar]
  221. Park, C.Y.; Kim, Y.J.; Lim, H.J.; Park, J.H.; Kim, M.J.; Seo, S.W.; Park, C.P. Continuous flow photooxygenation of monoterpenes. RSC Adv. 2015, 5, 4233–4237. [Google Scholar] [CrossRef]
  222. Wau, J.S.; Robertson, M.J.; Oelgemöller, M. Solar Photooxygenations for the Manufacturing of Fine Chemicals—Technologies and Applications. Molecules 2021, 26, 1685. [Google Scholar] [CrossRef]
  223. Radjagobalou, R.; Blanco, J.-F.; Petrizza, L.; Le Bechec, M.; Dechy-Cabaret, O.; Lacombe, S.; Save, M.; Loubière, K. Efficient Photooxygenation Process of Biosourced α-Terpinene by Combining Controlled LED-Driven Flow Photochemistry and Rose Bengal-Anchored Polymer Colloids. ACS Sustain. Chem. Eng. 2020, 8, 18568–18576. [Google Scholar] [CrossRef]
  224. Lee, D.S.; Amara, Z.; Clark, C.A.; Xu, Z.; Kakimpa, B.; Morvan, H.P.; Pickering, S.J.; Poliakoff, M.; Gorge, M.W. Continuous Photo-Oxidation in a Vortex Reactor: Efficient Operations Using Air Drawn from the Laboratory. Org. Process Res. Dev. 2017, 21, 1042–1052. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  225. Radjagobalou, R.; Blanco, J.-F.; Dechy-Cabaret, O.; Oelgemöller, M.; Lobière, K. Photooxygenation in an advanced led-driven module: Experimental investigations and modelling. Chem. Eng. Process. Process Intensif. 2018, 130, 214–228. [Google Scholar] [CrossRef] [Green Version]
  226. Lévesque, F.; Seeberger, P.H. Continuous-Flow Synthesis of the Anti-Malaria Drug Artemisinin. Angew. Chem. Int. Ed. 2012, 51, 1706–1709. [Google Scholar] [CrossRef] [PubMed]
  227. Michelin, C.; Lefebvre, C.; Hoffmann, N. Les réactions photochimiques à l’échelle industrielle. Actual. Chim. 2019, 436, 19–26. [Google Scholar]
  228. Emmanuel, N.; Mendoza, C.; Winter, M.; Horn, C.R.; Vizza, A.; Dreesen, L.; Heinrichs, B.; Monbaliu, J.-C.M. Scalable Photocatalytic Oxidation of Methionine under Continious-Flow Conditions. Org. Process Res. Dev. 2017, 21, 1435–1438. [Google Scholar] [CrossRef]
  229. Montagnon, T.; Kalaitzakis, D.; Sofadis, M.; Vassilikogiannakis, G. Chemoselective photooxygenations of furans bearing unprotected amines: Their use in alkaloid synthesis. Org. Biomol. Chem. 2016, 14, 8636–8640. [Google Scholar] [CrossRef]
  230. Gollnick, K.; Griesbeck, A. Singlet oxygen photooxygenation of furans: Isolation and reactions of (4+2)-cycloaddition products (unsaturated sec.-ozonides). Tetrahedron 1985, 41, 2057–2068. [Google Scholar] [CrossRef]
  231. Schenck, G.O. Photochemische Reaktionen II. Über die unsensibilisierte und photosensibilisierte Autoxydation von Furanen. Justus Liebigs Ann. Chem. 1953, 584, 156–176. [Google Scholar] [CrossRef]
  232. Cottier, L.; Descotes, G.; Nigay, H.; Parron, J.-C.; Grégoire, V. Photo-oxygénation des dérivés de l’hydroxyméthyl-5 furfural-2. Bull. Soc. Chim. Fr. 1986, 5, 844–850. [Google Scholar]
  233. Chauhan, D.K.; Battula, V.R.; Giri, A.; Patra, A.; Kailasam, K. Photocatalytic valorization of furfural to value-added chemicals via mesoporous carbon nitride: A possibility through a metal-free pathway. Catal. Sci. Technol. 2022, 12, 144–153. [Google Scholar] [CrossRef]
  234. Desvals, A.; Fortino, M.; Lefebvre, C.; Rogier, J.; Michelin, C.; Alioui, S.; Rousset, E.; Pedone, A.; Lemercier, G.; Hoffmann, N. Synthesis and characterization of polymethine dyes carrying thiobarbituric and carboxylic acid moieties. New J. Chem. 2022, 46, 8971–8980. [Google Scholar] [CrossRef]
  235. Miles, W.H. Synthetic Applications of γ-Hydroxybutenolides. Curr. Org. Synth. 2014, 11, 244–267. [Google Scholar] [CrossRef]
  236. Lefebvre, C.; Hoffmann, N. Les colorants et la lumière pour transformer la matière. Actual. Chim. 2019, 444–445, 38–43. [Google Scholar]
  237. Martel, J.; Tessier, J.; Demoute, J.-P. (Roussel Uclaf), Procédé de préparation de la 6,6-diméthyl-4-hydroxy-3-oxabicyclo(3.1.0)hexan-2-one et de ses éthers sous toutes leurs formes stéréoisomères. Eur. Pat. 1981, 0023454. [Google Scholar]
  238. Moradei, O.M.; Paquette, L.A. (5S)-(d-Menthyloxy)-2(5H)-furanone. Org. Synth. 2003, 80, 66–73. [Google Scholar]
  239. Marinković, S.; Brulé, C.; Hoffmann, N.; Prost, E.; Nuzillard, J.-M.; Bulach, V. Origin of Chiral Induction in Radical Reactions with the Diastereoisomers (5R)- and (5S)-5-l-Menthyloxyfuran-2[5H]-one. J. Org. Chem. 2004, 69, 1646–1651. [Google Scholar] [CrossRef] [PubMed]
  240. Feringa, B.L.; de Jong, J.C. New Strategies in Asymmetric Synthesis Based On γ-Alkoxybutenolides. Bull. Soc. Chim. Belg. 1992, 101, 627–640. [Google Scholar] [CrossRef] [Green Version]
  241. Gassama, A.; Ernenwein, C.; Youssef, A.; Agach, M.; Riguet, E.; Marinković, S.; Hoffmann, N. Sulfonated surfactants obtained from furfural. Green Chem. 2013, 15, 1558–1566. [Google Scholar] [CrossRef]
  242. Gassama, A.; Ernenwein, C.; Hoffmann, N. Photochemical Key Steps in the Synthesis of Surfactants from Furfural-Derived Intermediates. ChemSusChem 2009, 2, 1130–1137. [Google Scholar] [CrossRef]
  243. Hoffmann, N. Efficient photochemical electron transfer sensitization of homogeneous organic reactions. J. Photochem. Photobiol. C 2008, 9, 43–60. [Google Scholar] [CrossRef]
  244. Tan, J.-N.; Ahmar, M.; Queneau, Y. Isomaltulose Oxidation and Dehydration Products as Starting Materials towards Fine Chemicals. Curr. Org. Chem. 2014, 18, 1768–1787. [Google Scholar] [CrossRef]
  245. Tan, J.-N.; Ahmar, M.; Queneau, Y. Glucosyloxymethylfurfural (GMF): A creative renewable scaffold towards bioinspired architectures. Pure Appl. Chem. 2015, 87, 827–839. [Google Scholar] [CrossRef]
  246. Lichtenthaler, F.W.; Martin, D.; Weber, T.; Schiweck, H. 5-(α-D-Glucosyloxymethyl)furfural: Preparation from Isomaltulose and Exploration of Its Ensuing Chemistry. Liebigs Ann. Chem. 1993, 1993, 967–974. [Google Scholar] [CrossRef]
  247. Tadiparthi, K.; Venkatesh, S. Synthetic approaches toward butenolide-containing natural products. Dihydro-5-(hydroxymethyl)-2(3H)-furanone. J. Heterocycl. Chem. 2022, 59, 1285–1307. [Google Scholar] [CrossRef]
  248. Flourat, A.L.; Haudrechy, A.; Allais, F.; Renault, J.-H. (S)-γ-Hydroxymethyl-α,β-butenolide, a Valuable Chiral Synthon: Syntheses, Reactivity, and Applications. Org. Process Res. Dev. 2020, 24, 615–636. [Google Scholar] [CrossRef] [Green Version]
Figure 1. Structure of lignin. (Reproduced with the permission of the American Society of Plant Biologists [81]).
Figure 1. Structure of lignin. (Reproduced with the permission of the American Society of Plant Biologists [81]).
Molecules 28 04746 g001
Scheme 1. Catalytic cleavage of lignin model dimers and polymers involving photoredox catalysis with visible light.
Scheme 1. Catalytic cleavage of lignin model dimers and polymers involving photoredox catalysis with visible light.
Molecules 28 04746 sch001
Scheme 2. Mechanism of the photoredox catalytic cleavage of lignin model dimers.
Scheme 2. Mechanism of the photoredox catalytic cleavage of lignin model dimers.
Molecules 28 04746 sch002
Scheme 3. Convenient synthesis of tetrahydroxyazepanes form glucose and mannose derivatives.
Scheme 3. Convenient synthesis of tetrahydroxyazepanes form glucose and mannose derivatives.
Molecules 28 04746 sch003
Scheme 4. Mechanism of the Semmelhack reaction.
Scheme 4. Mechanism of the Semmelhack reaction.
Molecules 28 04746 sch004
Scheme 5. Photoredox catalytic glycosidation with anisylthioether derivatives. β-Anomeric glucosilation products are preferentially formed.
Scheme 5. Photoredox catalytic glycosidation with anisylthioether derivatives. β-Anomeric glucosilation products are preferentially formed.
Molecules 28 04746 sch005
Scheme 6. Mechanism of the Photoredox catalytic glycosidation with anisylthioether derivatives involving photochemical electron transfer with an iridium complex.
Scheme 6. Mechanism of the Photoredox catalytic glycosidation with anisylthioether derivatives involving photochemical electron transfer with an iridium complex.
Molecules 28 04746 sch006
Scheme 7. Combined photo and copper catalysis applied to the coupling reactions of carbohydrate derived acids with alkynes.
Scheme 7. Combined photo and copper catalysis applied to the coupling reactions of carbohydrate derived acids with alkynes.
Molecules 28 04746 sch007
Scheme 8. Electronic excitation of a furanone to the triplet state via triplet energy transfer.
Scheme 8. Electronic excitation of a furanone to the triplet state via triplet energy transfer.
Molecules 28 04746 sch008
Scheme 9. Intramolecular photosensitized radical addition of a carbohydrate moiety to a furanone. Acetone was used as sensitizer and cosolvent in a ratio of 1/4 with acetonitrile.
Scheme 9. Intramolecular photosensitized radical addition of a carbohydrate moiety to a furanone. Acetone was used as sensitizer and cosolvent in a ratio of 1/4 with acetonitrile.
Molecules 28 04746 sch009
Scheme 10. Photochemical decarboxylation of a fatty acid derivative applied to a coupling with alkynes.
Scheme 10. Photochemical decarboxylation of a fatty acid derivative applied to a coupling with alkynes.
Molecules 28 04746 sch010
Scheme 11. Photoredox catalytic borylation of a fatty acid via photocatalytic decarboxylation.
Scheme 11. Photoredox catalytic borylation of a fatty acid via photocatalytic decarboxylation.
Molecules 28 04746 sch011
Scheme 12. Suggested mechanism of the photoredox catalytic borylation with fatty acids (Scheme 11).
Scheme 12. Suggested mechanism of the photoredox catalytic borylation with fatty acids (Scheme 11).
Molecules 28 04746 sch012
Figure 2. Terminal alkenes obtained by decarboxylation of fatty acids. Combined photoredox and metal catalysis was applied.
Figure 2. Terminal alkenes obtained by decarboxylation of fatty acids. Combined photoredox and metal catalysis was applied.
Molecules 28 04746 g002
Scheme 13. Photoredox catalytic addition cyclisation reaction with radicals generated from oleic acid.
Scheme 13. Photoredox catalytic addition cyclisation reaction with radicals generated from oleic acid.
Molecules 28 04746 sch013
Scheme 14. Photocatalytic addition of radicals to levoglucosenone 64.
Scheme 14. Photocatalytic addition of radicals to levoglucosenone 64.
Molecules 28 04746 sch014
Scheme 15. Mechanism of the photocatalytic addition of radicals to levoglucosenone 64.
Scheme 15. Mechanism of the photocatalytic addition of radicals to levoglucosenone 64.
Molecules 28 04746 sch015
Scheme 16. Generation of singlet oxygen involving photochemical sensitization.
Scheme 16. Generation of singlet oxygen involving photochemical sensitization.
Molecules 28 04746 sch016
Scheme 17. Photooxygenation of furfural 70 to hydroxyfuranone 72.
Scheme 17. Photooxygenation of furfural 70 to hydroxyfuranone 72.
Molecules 28 04746 sch017
Figure 3. Surfactants obtained from hydroxyfuranone 72 (Scheme 17).
Figure 3. Surfactants obtained from hydroxyfuranone 72 (Scheme 17).
Molecules 28 04746 g003
Scheme 18. Synthesis of a α-connected glycosyl furanone 69 from isomaltulose.
Scheme 18. Synthesis of a α-connected glycosyl furanone 69 from isomaltulose.
Molecules 28 04746 sch018
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Gómez Fernández, M.A.; Hoffmann, N. Photocatalytic Transformation of Biomass and Biomass Derived Compounds—Application to Organic Synthesis. Molecules 2023, 28, 4746. https://doi.org/10.3390/molecules28124746

AMA Style

Gómez Fernández MA, Hoffmann N. Photocatalytic Transformation of Biomass and Biomass Derived Compounds—Application to Organic Synthesis. Molecules. 2023; 28(12):4746. https://doi.org/10.3390/molecules28124746

Chicago/Turabian Style

Gómez Fernández, Mario Andrés, and Norbert Hoffmann. 2023. "Photocatalytic Transformation of Biomass and Biomass Derived Compounds—Application to Organic Synthesis" Molecules 28, no. 12: 4746. https://doi.org/10.3390/molecules28124746

Article Metrics

Back to TopTop