Next Article in Journal
Treatment with Akebia Saponin D Ameliorates Aβ1–42-Induced Memory Impairment and Neurotoxicity in Rats
Next Article in Special Issue
Porphyrin Macrocycle Modification: Pyrrole Ring-Contracted or -Expanded Porphyrinoids
Previous Article in Journal
Cloning, Expression Profiling and Functional Analysis of CnHMGS, a Gene Encoding 3-hydroxy-3-Methylglutaryl Coenzyme A Synthase from Chamaemelum nobile
Previous Article in Special Issue
Recent Advances in Solid Catalysts Obtained by Metalloporphyrins Immobilization on Layered Anionic Exchangers: A Short Review and Some New Catalytic Results
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Porphyrins as Catalysts in Scalable Organic Reactions

by
Juan C. Barona-Castaño
,
Christian C. Carmona-Vargas
,
Timothy J. Brocksom
and
Kleber T. De Oliveira
*
Departamento de Química, Universidade Federal de São Carlos, São Carlos 13565-905, Brazil
*
Author to whom correspondence should be addressed.
These authors contributed equally to this work.
Molecules 2016, 21(3), 310; https://doi.org/10.3390/molecules21030310
Submission received: 29 January 2016 / Revised: 26 February 2016 / Accepted: 1 March 2016 / Published: 8 March 2016

Abstract

:
Catalysis is a topic of continuous interest since it was discovered in chemistry centuries ago. Aiming at the advance of reactions for efficient processes, a number of approaches have been developed over the last 180 years, and more recently, porphyrins occupy an important role in this field. Porphyrins and metalloporphyrins are fascinating compounds which are involved in a number of synthetic transformations of great interest for industry and academy. The aim of this review is to cover the most recent progress in reactions catalysed by porphyrins in scalable procedures, thus presenting the state of the art in reactions of epoxidation, sulfoxidation, oxidation of alcohols to carbonyl compounds and C–H functionalization. In addition, the use of porphyrins as photocatalysts in continuous flow processes is covered.

Graphical Abstract

1. A Brief History of Catalysis

Catalysis is a phenomenon observed since ancient times when Valerius Cordus (1514–1544) converted ethanol to ethyl ether using sulfuric acid as catalyst [1]. Formally, this term was proposed in 1835 by Berzelius (1779–1848), who defined catalysis as the ability of substances “to awaken affinities, which are asleep at a particular temperature, by their mere presence and not by their own affinity” [1,2]. Later, Ostwald (Nobel Prize in Chemistry in 1909) introduced a rational physical chemical definition of a catalyst in 1895 as a “substance that can change the reaction rate (accelerate or inhibit) without modification of the energy factors of the reaction” [2]. However, the complete molecular basis of the catalytic processes was established just one century later, [3] and now is widely applied on a large scale. One important and historical example in industry is the Haber–Bosch process for ammonia synthesis [3].
In fact, acid-base and metal catalysis have been the most recurrent procedures in industry and academy over the last centuries, allowing the growth of chemistry with many benefits for society. However, very complex systems have been developed more recently, including mixed metallo-organic materials and multiple catalysts [4].
Currently, almost all the chemical processes in industry are mediated by catalysts implying large annual expenses, involving green and energy-saving processes. In 2009 Noyori [5] stated “I personally consider that catalysis is the most important subject in chemistry and also technology—80 per cent of all commercial products are made by catalysis and the total market of these commercial products is $7 trillion [£4.3 trillion]”; in a very recent editorial Zhou affirms that it is now 90% [6].
Taking into account the challenges of the global economy and the needs for more sustainable chemical processes, catalysed reactions represent an important role for the development of new synthetic methods to generate target molecules in fewer steps and with less chemical waste. Certainly, the design, synthesis and applications of new catalysts are a hot research topic in the three scientific disciplines, considered essential for catalysis: chemical engineering, inorganic chemistry and organic chemistry.
In this context, it is important to highlight that porphyrin derivatives are special molecular scaffolds, which present relevant activities and cost-competitive transformations in these fields, thus deserving attention. In this review, we intend to cover the most relevant scalable porphyrin-catalysed procedures, showing how these compounds represent broad applications in chemistry.

2. General Concepts in Porphyrin Catalysis

Porphyrins and their derivatives are a class of naturally occurring macrocyclic compounds with intense colour, that have been extensively studied [7,8,9] due to the key role played in some life processes (Figure 1 and Figure 2) and involving their special aromatic structure (18π electrons). A suitable example that can be given is the heme group which contains an iron-porphyrin complex which coordinates oxygen and carbon dioxide for cellular carriage [10], responsible for cellular respiration, and contributing to the catalytic activities of many enzymes as cofactors [11]. This is one of the reasons why porphyrin-derived pigments are called “the Pigments of Life” (Figure 1 and Figure 2) [12,13].
The history of the study of metalloporphyrins’ activities began in 1747 when Menghini demonstrated for the first time the presence of iron in blood [14]. Hoppe and Seyler in 1871 isolated porphyrins from blood, describing these compounds as pyrrole derivatives, and in 1940 the structure and biological functions of iron porphyrin complexes were well established [15]. Nowadays, this kind of metalloporphyrins is well known since they are prosthetic group of an important class of proteins and enzymes called hemoproteins.
Metalloporphyrins have also been widely studied as bioinspired models of cytochrome P-450 (hemoproteins), and have exhibited catalytic activity for highly selective monooxygenation reactions [16], which proceed via formation of a high valence metal-oxygen complex intermediate. In 1979, Groves and co–workers developed the first oxidation system using a synthetic metalloporphyrin as a bioinspired catalyst [17].
Since then, the synthesis and study of these very robust macrocycles inspired by biological systems, have been incorporated into the development of sophisticated bulky, chiral, and surface linked catalysts, because of a variety of properties that turn them useful for organic synthesis (Scheme 1) [18].
The development of metalloporphyrin catalysts as synthetic tools has recently evolved significantly [19]. For instance, in catalytic C–H oxidation reactions, the simplest metalloporphyrins without substituents at the meso positions are not useful because they undergo fast oxidative degradation [20]. The degradation process (the catalase reaction) leads to hydroxylation at the meso position, followed by other processes that occur in natural heme degradation, in order to form meso-hydroxyporphyrin derivatives which then inactivate the catalytic activity [20]. Therefore, it has been demonstrated that the introduction of phenyl or related groups at the meso positions (Figure 3) is a good strategy which provides efficient catalytic oxidations by protecting these reactive sites [19].
In 1997, Dolphin and Traylor proposed a classification system for all the different metalloporphyrins used in catalysis based on their structural features (Figure 4) [21]. These authors classified the synthetic metalloporphyrins without substituents in the aryl moiety at the meso positions as first generation porphyrins, and one example is the meso-tetraphenylporphyrin iron(III) chloride ([Fe(TPP)]Cl) that Groves et al. employed in cytochrome P-450 bioinspired catalysis (Figure 4a) [17].
Nonetheless, meso–phenyl-substituted metalloporphyrins bearing electronegative and bulky groups were classified as second-generation porphyrins such as meso-tetrakis(pentafluorophenyl)porphyrin iron(III) chloride and meso-tetramesitylporphyrin iron(III) chloride (TMPFeCl) (Figure 4b). The introduction of electron-withdrawing groups (such as halogens) in the β-pyrrole positions of second-generation porphyrins yields the so-called third generation porphyrin derivatives (Figure 4c) [22], which provide better catalytic activity over the other porphyrin generations, according to Haber and co-workers [23].
Over the last few decades, a number of publications have covered the many catalytic activities of porphyrins; however, these studies were typically focused on methodology, converting only a few milli- or micrograms of substrates during the experiments. Herein, we intend to present porphyrin derivatives which are able to perform scaled-up transformations.

3. Oxidation Reactions with Porphyrin and Metalloporphyrin–Based Catalysts

Oxidation reactions are important synthetic tools, and a number of applications can be found in the chemical industry [24]. The manufacture of products obtained from oxidation of organic substrates and the production of pharmaceutical ingredients (APIs) are of major importance [25].
The inert nature of C–H bonds requires the use of highly reactive reagents and stoichiometric amounts of oxidizing agents such as strong inorganic acids, peroxyacids, or highly toxic oxo–metal oxidants; all yield products with low chemo, regio and stereoselectivity, and generate large amounts of toxic waste [26]. However, since the discovery of cytochrome P450, the use of metalloporphyrin catalysts has emerged as an alternative for controlled oxidation reactions in the above aspects (Figure 5) [27].
This cytochrome family of enzymes play a key role in aerobic oxidation reactions in biological systems under mild conditions, such as highly selective hydroxylation of alkanes (C–O bond formation via saturated C–H bond functionalization) (Scheme 2) [28,29]. Metalloporphyrins with ruthenium, iron, manganese, among other metals, constitute the family of catalysts which are efficient to mediate C–H oxidations with high selectivity and good yields [30].

3.1. Epoxidation

The epoxidation of alkenes is one of the processes of great importance in the fine chemical industry from an economic point of view, because epoxides are useful intermediates in the production and manufacture of high–value commercial polymers like polyurethane, polyamides, resins, and polyesters [31]. In addition, this transformation is used to carry out bioinspired oxidations [32] to produce drug candidates or metabolites (Scheme 3).
Iron porphyrins are effective catalysts for epoxidation of olefins by a number of oxidants, such as iodosylbenzene (PhIO) and 2,6-dichloropyridine-N-oxide (2,6-Cl2pyNO) [15]. Due to the relevance of cytochrome P-450 mechanism, catalytic epoxidation by iron porphyrins has gained great attention [33]. Since the Groves and co-workers’ first report in 1983 [34], several research groups have employed chiral metalloporphyrins [35] to catalyse the oxidation of various organic substrates, and new synthetic routes have been created to improve the catalytic performance of these heterocyclic complexes.
The complexes FeTHAPP and FeTCBCPP (Figure 6) are examples of metalloporphyrins which are able to perform asymmetric induction in the presence of either PhIO or iodosylmesitylene as oxygen donors, achieving products with enantioselectivities up to 50% and a turnover number (TON) near to 100 [36].
Optically active epoxides obtained from the catalytic enantioselective epoxidation of alkenes are important intermediates in asymmetric synthesis, since these compounds are very useful for the synthesis of chirons with up to two contiguous stereogenic centres [37,38]. Chiral epoxidation of allylic alcohols provides interesting building blocks for the asymmetric synthesis of biologically active compounds, and consequently, asymmetric epoxidation of allylic alcohols has been extensively developed [39]. Chiral non-racemic iron porphyrins with binaphthyl moieties linked to the macrocycle allow the enantioselective epoxidation of non-functionalized terminal olefins with good enantioselectivity (>97% for styrene) and high turnover numbers (>16,000) (Scheme 4) [35,40].
In general, iron porphyrins present good reactivity to catalyse two primary reactions, namely the enantioselective transfer of an oxygen atom from the hydrogen peroxide to the substrate, and the decomposition of the peroxide in water and oxygen [41]. On the other hand, the use of manganese porphyrins as catalysts with hydrogen peroxide as oxidant, yield low enantiomeric excesses for epoxidations in organic solvents or biphasic medium [42].
However, Simonneaux and co-workers reported in 2012 the enantioselective epoxidation of styrene derivatives using H2O2 (ee up to 68%), in water-methanol solutions using chiral water-soluble manganese and iron porphyrins as catalysts (Scheme 5) [43,44]. They also studied various factors which affect the catalytic epoxidation of olefins and found that the water present in the methanol is quite useful [44].
Besides the hydrogen peroxide decomposition [45] and other factors like the structural features of the porphyrin catalyst [46], the substrate concentration and the solvent [47] affect strongly the activity and the selectivity of these catalysts.
Scheme 6 shows how the reaction conditions affect the total conversion of alkenes when an iron porphyrin is employed as catalyst in a mixture of dichloromethane and methanol. In this case, direct oxygen transfer to obtain the epoxide is observed from the high-valent Fe(IV)-porphyrin radical cation complex (path A) [37]. On the other hand, when an aprotic solvent such as acetonitrile is used, the reactive intermediate is more likely to be the iron hydroperoxide complex, leading to the generation of competing radicals, thus promoting low selectivity and yield even though an imidazole is added to help the stabilization of the intermediate (path B) [37].
Despite numerous studies on the metalloporphyrin–catalysed epoxidation of olefins, the substrate types in such oxidation systems reported in the literature were rather limited and were usually confined to electron–rich alkenes such as styrenes, norbornene, cyclohexene and cyclooctene. However, employing dendritic ruthenium porphyrins [48] as catalysts, and 2,6-Cl2pyNO, t-BuOOH or O2, some authors have described efficient epoxidations of different alkenes in high yields and turnover numbers (>700).
Che and co-workers investigated the catalytic properties of the catalyst for the epoxidation of unsaturated cholesteryl esters with 2,6-Cl2pyNO in dichloromethane using just 0.1 mol % of the dendritic ruthenium porphyrin catalyst (Scheme 7) [48].

3.2. Sulfides to Sulfoxides

Since the pioneering work of Oae and co-workers (Scheme 8) [49] on the bioinspired oxidation of sulfides to sulfoxides, this transformation has gained much attention. Historically, sulfur-containing compounds have figured as targets for their importance to the pharmaceutical industry as antibacterial agents [25]. Therefore, the development of catalytic systems for the preparation of optically active sulfoxides is important, because they are chiral synthons [50] in the synthesis of bioactive compounds [51].
Sulfoxidation is commonly presented as being a direct pathway for generating sulfoxides, however, most of the reagents used for this reaction such as iodosylbenzene, peroxyacids, and stoichiometric oxo-metal oxidants are unsatisfactory due to their high toxicity and low chemoselectivity between sulfoxide and sulfone products [52]. One successful example of a green protocol was described by Baciocchi and co-workers, who reported the oxidation of sulfides with an ethanolic solution of H2O2 and iron tetrakis(pentafluorophenyl)porphyrin as catalyst, thus giving the corresponding sulfoxides on a gram-scale (Scheme 9) [53].
Among the methods to obtain sulfoxides [54], the enantioselective oxidation of sulfides with small amounts of metal-organic catalysts is one of the most attractive routes to optically active sulfoxides [55]. For example, Simonneaux and co-workers reported a small scale enantioselective sulfide oxidation with 35% aqueous hydrogen peroxide and a chiral water-soluble iron-porphyrin as catalyst [56], thus obtaining sulfoxides with enantioselectivities between 78% and 90% from alkyl aryl sulfides, bearing electron-withdrawing groups in the phenyl ring (Scheme 10a).
The advantages of the process is supported by the absence of excess of oxidant, small reaction time, reaction at room temperature, and protection against destruction of the macrocycle by the two bulky norbornane groups joined to the central benzene ring [56]; however, this procedure was not scaled-up, and only a few micrograms of sulfide were converted.
Another example of an enantioselective oxidation of sulfides (small scale) catalysed by a chiral manganese porphyrin in an aqueous methanol solution in the presence of H2O2 furnished the non-steroidal anti-inflammatory drug Sulindac® (Banyu Pharmaceutical Co., Ltd., Tokyo, Japan) (Scheme 10b) [57].
Nonetheless, the formation of sulfoxides using metalloporphyrins as catalyst can be achieved by other methodologies, but in this case, on a superior scale. The use of oxygen as the oxidant with NaBH4 and Me4NOH, [58] is an example to obtain the oxidized product of diarylsulfide compounds (Scheme 11).

3.3. Hydroxylation

Once the mechanism by which the cytochrome P-450 acts as catalyst in C–H functionalization became known [59], with an iron porphyrin core as the active site, many synthetic metalloporphyrins have been studied as catalyst for this purpose with different organic substrates [60,61]. In fact, this is one of the most difficult transformations for the petrochemical industry, and one of the major challenge to convert petroleum derivatives into chemicals of higher value [62].
The first report of metalloporphyrin-catalysed hydroxylation of saturated C–H bonds was published by Groves and co-workers in 1979 [17], where they showed the catalytic activity of [Fe(TPP)Cl] towards oxidation of cyclohexane and adamantane with iodosylbenzene to give the corresponding alcohols. Approximately 10 years later, Groves and Viski reported for the first time, the enantioselective hydroxylation of ethylbenzenes catalysed by a chiral iron porphyrin with up to 77% ee using PhIO as oxidant (Scheme 12) [63,64]. When manganese was used, higher yields were obtained, but the enantioselectivity decreased and ketones were also observed as by-products in the catalytic reactions.
In 1999, Gross and Ini reported the first example of ruthenium-catalysed enantioselective (ee up to 38%) hydroxylation of racemic tertiary alkanes, using a chiral porphyrin ligand and 2,6-Cl2pyNO as oxidant (Scheme 13) [65].
Che and co-workers [66] reported the Ru-catalysed enantioselective hydroxylation of aromatic hydrocarbons with benzylic C–H bonds. Using 2,6-Cl2pyNO as oxidant, the chiral ruthenium porphyrin was shown to be an effective catalyst for hydroxylating a series of aromatic hydrocarbons to form the corresponding secondary alcohols. Although the conversion of the substrates was incomplete, the Ru-based hydroxylation gave yields up to 76% ee (Scheme 14).
In 2012 Simonneaux and co-workers reported examples of enantioselective hydroxylation of alkanes, with a chiral iron porphyrin as catalyst and using hydrogen peroxide as oxidant in methanol and water, to give optically active secondary alcohols (ee up to 63%) [43]. Nonetheless, one limitation for this system is the requirement of an excess of alkane versus the oxidant, and consequently the asymmetric hydroxylation of alkanes using these conditions is still difficult to proceed without electron-deficient chiral metalloporphyrins [44]. The treatment of ethylbenzene with H2O2 and catalysis by the complex (Scheme 15) afforded a mixture of 1–phenylethanol (57%) and acetophenone (43%) with 88% of conversion [43].
Other types of porphyrins have been explored for this hydroxylation reaction, and in 2007, Idemori and co–workers studied Mn(III)–tetrapyridylporphyrin as catalyst in the hydroxylation of cyclohexene [67]. Iida and co-workers have shown that the combination of tert-butyl hydroperoxide (TBHP) and the osmium(II) carbonyl complex of meso-tetramesitylporphyrin [Os(TMP)(CO)] as oxygen donor and catalyst respectively (Scheme 16), is an efficient and versatile oxidation system for the functionalization of bioactive molecules, like bile acids [68] and terpenoids [69].
The oxy-functionalization system with osmium-porphyrins as catalyst regioselectively hydroxylates one of the inactivated tertiary C–H bonds in dihydrofaradiol diacetate to form the corresponding alcohol in 50% yield (Scheme 17). The functionalization of triterpenoids with the oxidant system afforded a variety of novel oxygenated derivatives in one-step with good isolated yields [69].

3.4. Oxidation of Alcohols to Carbonyl Compounds

The transformation of primary alcohols to the corresponding aldehydes or carboxylic acids is important for organic synthesis, since these are versatile functional groups which are present in many building blocks [71] (Scheme 18).
Conventionally, stoichiometric or even over-stoichiometric amounts of metal oxides and metal complexes are used for these oxidations [72], but the catalytic oxidation with O2 is of sustainability interest [73].
Metalloporphyrins have been widely used as catalysts for various oxidations of alcohols with PhIO [74], 2,6-Cl2PyNO [75], m-chloroperbenzoic acid [76], Bu4NHSO5 (tetrabutylammonium peroxymonosulfate) [77] and oxygen as oxidants. Woo and co-workers reported the aerobic homogeneous oxidation of benzyl alcohol with oxo-titanium porphyrin ((TPP)Ti=O), which gave benzaldehyde in 50% yield [78]. It has been found that (TPP)Ti=O reacts with free diols to yield diolato complexes [79], and these complexes undergo oxidative cleavage reactions at high temperatures to release carbonyl compounds.
Ji and co-workers reported a selective oxidation of alcohols to carbonyl compounds by molecular oxygen with isobutyraldehyde as oxygen acceptor in the presence of ruthenium (III) meso-tetraphenylporphyrin chloride (Ru(TPP)Cl) (Scheme 19) [80]. In addition, they found that the catalytic activity and selectivity of metalloporphyrins towards benzaldehydes appeared to be dependent on the nature of the central ion, and were influenced by the stability of different valences of the metal atoms and their respective electric potential.
Oxidation of aldehydes to the corresponding carboxylic acids is one of the ubiquitous transformation in organic synthesis, for the preparation of numerous APIs, vitamins and fragrances [73]. Rebelo and co-authors investigated the oxidation of benzaldehyde with hydrogen peroxide using a Mn(III) porphyrin/ammonium acetate system to give benzoic acid in 93% yield [81].
The oxidation of α,β-enones at the γ position is a relevant transformation, with few known methodologies. Che and co-workers reported the Ru-porphyrin catalysed oxidation of ketosteroids to form diketosteroids using meso-tetrakis(2,6-dichlorophenyl)porphyrin (Ru(TDClPP)Cl2) as catalyst and 2,6–Cl2pyNO as oxidant (Scheme 20) [82].

4. Porphyrin Derivatives Acting as Photocatalysts

Since 1930, when Kautsky first examined the generation of singlet oxygen {O2 (1Δg)}, by energy transfer from a photoexcited organic molecules to molecular oxygen, this kind of environmentally friendly methodology has been of interest [83]. The importance of singlet oxygen in many physical and biological processes has been recognized, with rich details concerning the physics and chemistry of this electronically excited molecule [84]. Generally, singlet oxygen can be generated by chemical or photosensitized reactions, but the photochemical approach is more cost-competitive. The mechanism of the photochemical singlet oxygen generation includes the promotion of the photosensitizer to the triplet state, passing through a singlet -excited state. At the triplet state many photosensitizers are able to transfer energy to molecular oxygen leading to the formation of reactive oxygen species (ROS) (Figure 7) [85].
In this context, porphyrins are prominent photosensitizers because of their high light absorption coefficient, exited state energy levels, and high photostability as compared to other dyes. For example, meso-tetraphenylporphyrin (TPP) has been chosen as one of the most highly effective photosensitizer [86] for singlet oxygen generation [87].
In 1999, methylene blue was used as photosensitizer in the photooxidation of 8-hydroxyquinoline to afford quinoline-5,8-quinone in 64%–70% yields [88]. However, Cossy and Belotti showed in 2001 an improved method using TPP for the photooxygenation of substituted 8-hydroxyquinolines in 50%–89% yield (Scheme 21) [85].
Spivey and co-workers reported in 2010 the synthesis of high-functionalized molecules using two photooxygenation reactions with oxygen, visible light and TPP as photocatalyst, including a base-catalysed Kornblum DeLaMare rearrangement. Also, Co(II)TPP was used three times as the oxidation catalyst (Scheme 22) showing its versatility in the entire synthetic approach [89].
By comparison, in 2011 Nicolaou presented the total synthesis of dithiodiketopiperazines (epicoccin G and 8,8’-epi-ent-rostratin B) including two TPP-photocatalysed endoperoxide formations, and base-catalysed Kornblum DeLaMare rearrangements (Scheme 23) [90]. Singlet oxygen generation by porphyrins can take place in reactions with amines [91]. In the case of secondary amines, there is an example with the dehydrogenation of N–neopentylallylic amine using TPP as photocatalyst to obtain the corresponding imine product [92]. Moreover, Che and co-workers [93] reported the highly efficient photooxidation of secondary benzylamines to imines in 90% yield using molecular oxygen and TPP as photosensitizer. They used directly the in situ formed imine products for further functionalization to develop an oxidative Ugi–type MCR (Scheme 24).
Another photocatalytic application of porphyrins is their encapsulation in microemulsions of ethyl acetate, water and surfactants to allow the use of TPP as photosensitizer for singlet oxygen generation [94]. For example, Oelgemöller and co-workers carried out the photooxygenation of 1,5-dihydroxynaphthalene to Juglone (5-hydroxy-1,4-naphthoquinone) as a reaction model in a micellar system, but only on a small-scale (Scheme 25) [95].

4.1. Immobilized Porphyrins as Photocatalysts

One approach to enhance the usefulness of metalloporphyrin catalysis, involves recyclability through heterogeneous catalytic systems [96] by using immobilized homogeneous catalysts [97]. In the case of asymmetric oxidation, immobilization can avoid the need for chiral ligand recovery. One of the simplest ways to prepare a polymer-immobilized catalyst is the direct reaction of a simple functionalized polymer, such as Merrifield`s resin with a derivative of the desired ligand and then insertion of the metal [98].
The covalent immobilization of porphyrins on polymeric matrices is very versatile to generate singlet oxygen [99]. Anchoring the photocatalyst onto a poly(ethylene glycol) allows the easy recovery of the catalyst from the reaction mixture, as Benaglia and co-workers [100] reported with a new poly(ethylene glycol)-supported porphyrin which exhibited high activity as catalyst, comparable to that of a non-anchored sensitizer (Scheme 26).
Che and co-workers reported a covalently attached carbonylruthenium(II)meso-tetrakis(pentafluorophenyl)porphyrin (Ru(TPFPP)(CO)) linked to hydrophilic PEG macromolecules [101]. The resulting PEG-supported Ru-porphyrin was shown to be an effective catalyst for oxidation of both secondary and tertiary C–H bonds (Scheme 27) [101].
In 2004, Shiragami and co-workers reported a scaled-up photocatalytic oxidation of cycloalkenes with O2 under visible-light using a SbTPP(OH)2\SiO2 catalyst (Scheme 28) [102].
Recently, Gonsalves and co-workers have synthesized supported materials with a non-symmetric halogenated tetra-arylporphyrin linked to Merrifield polymers, thus showing high efficiency in the scalable photooxygenations of different substrates using sunlight (Scheme 29) [103,104].

4.2. Continuous Flow Photocatalysis

Continuous flow reactors have recently emerged as a new technology in chemical synthesis with a wide variety of efficient applications [105,106]. The small inner dimensions of these devices in combination with their continuous flow operation make them especially attractive for photochemical reactions [107]. A combination of a microreactor system as the reaction medium and visible light offers a new and convenient approach towards green photochemistry, in which the production of by-products is also reduced. As a result, not only enhanced selectivity is achieved, but also the amount of environmentally problematic and expensive solvents are reduced, and the security is really improved [108].
Reactions with singlet oxygen (1O2) have not been used for large/industrial scale processes for the following reasons: (I) decreased photochemical efficiency of photosensitizers in batch conditions due to low light penetration in classic photochemical reactors); (II) the need for high substrate dilutions due to the safety requested by procedures with peroxides and endoperoxides formation; (III) major by-products formation; and (IV) difficulty to find suitable non-flammable and environmentally friendly solvents, with no incompatibility with singlet oxygen generation [109].
However, the use of supercritical carbon dioxide as solvent brings a solution to some of these issues, because it is neither flammable nor toxic, completely miscible with gases, and has a lower viscosity and higher diffusivity [110]. For example, George and co-workers reported the production of artemisinin (gram-scale) using liquid CO2 as solvent and an immobilized porphyrin as photocatalyst. They developed a simple method to anchor TPP or TPFPP onto the sulfonated cross-linked polystyrene ion-exchange resin (Amberlyst-15) [111]. This yielded a dual catalyst, with both the Bronsted acid and photo-catalytic functions needed to convert dihydroartemisinic acid (DHAA) into artemisinin, currently a very important drug for malaria treatment (Scheme 30). The reaction resulted in 98% of conversion using toluene as co-solvent.
George and co-workers have been the pioneers in using this kind of green solvent approach, since in 2011 they reported the use of supercritical carbon dioxide as solvent with different immobilized photosensitizers for the continuous flow synthesis of ascaridole from α-terpinene (Scheme 31) [112]. The authors coupled a porphyrin to amino functionalized PVC beads by covalent amide linkage, and achieved the continuous production of ascaridole and also citronellol hydroperoxide from citronellol [112].
Lapkin and co-workers[113] have also carried out a gram-scale photooxygenation/rearrangement of α-pinene using TPP as sensitizer in dichloromethane under continuous flow conditions, thus obtaining pinocarvone quantitatively. This oxygenation reaction was efficiently performed in a segmented gas-liquid flow condition in a very safe manner (Scheme 32) [113].
Seeberger and co-workers have also established gram-scale photoinduced singlet-oxygen generation under continuous flow by using TPP as photocatalyst, and a very simple photoreactor (perfluoroalkoxy—PFA—tube coiled around a cooled glass tube and a lamp) [114]. In this apparatus, the illumination of the entire solution is highly efficient because the tubing has a constant, narrow diameter, and it is positioned very close to the light source. Thus, an increasing quantity (space-time yield = 200g·day-1) of the product can be produced in long-term experiments (Scheme 33).
The above mentioned method has opened perspectives to carry out the crucial step in the synthesis of artemisinin (Scheme 33) [115]. The great importance of TPP as photosensitizer for singlet oxygen production can be highlighted by the green gram-scale synthesis of artemisinin recently established as an industrial process by the pharmaceutical company Sanofi [116].
To finish this section, it is important to highlight the recent and cost competitive protocol for photooxygenation of naphthols under continuous flow conditions, which was recently published by Oliveira, Miller and McQuade [117]. A careful methodological study was performed using four different porphyrins as photocatalysts. After optimisation of the reaction conditions, a number of substituted naphthols were converted into the corresponding naphthoquinones in high yields, as well as in gram-scale experiments (up to 1.4 g produced continuously in a 24 h experiment) (Scheme 34).

5. Perspectives

Over the last few decades, synthetic organic chemists have taken advantage of the catalytic activity of porphyrin derivatives to develop novel laboratory procedures and to enhance existing protocols, but normally on a small scale. Recently some scaled-up reactions have been described showing that these macrocycles are very useful for increasing the scope of available multiple catalysts. The challenge now is to explore the broad potential and applications of these catalysts for the synthesis of important products from an economic point of view, for the industrial and pharmaceutical industries.
In addition, another challenge is to introduce simple and cost-competitive synthetic routes [118] to access porphyrinods, since historically these compounds are considered rare due to low synthetic yields and tedious purification protocols. Therefore, it is necessary to incorporate enabling techniques for the optimisation of the production and utilization of these fascinating naturally inspired catalysts, and chemists must now consider these robust compounds for many new purposes.

Acknowledgments

The authors would like to thank the São Paulo Research Foundation (FAPESP 2015/21110-4, 2013/07276-1 and 2011/13993-2) for the financial support and also CAPES and CNPq for fellowships.

Author Contributions

Juan C. Barona-Castaño and Christian C. Carmona-Vargas contributed equally to this review. They accomplished the literature collection, designed the figures and schemes, and also wrote the first draft of the manuscript. Timothy J. Brocksom and Kleber T. de Oliveira wrote the final version of the manuscript, suggested modifications in schemes and figures, and checked the cited literature.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Wisniak, J. The history of catalysis. From the beginning to Nobel prizes. Educ. Quim. 2010, 21, 60–69. [Google Scholar]
  2. Roberts, M.W. Birth of the catalytic concept (1800–1900). Catal. Lett. 2000, 67, 1–4. [Google Scholar] [CrossRef]
  3. Van Santen, R. Catalysis in perspective: Historic review. In Catalysis: From Principles to Applications, 1st ed.; Beller, M., Renken, A., van Santen, R., Beller, M., Renken, A., Van Santen, R., Eds.; Wiley-VCH: Weinheim, Germany, 2012; pp. 3–19. [Google Scholar]
  4. Drain, C.M.; Hupp, J.T.; Suslick, K.S.; Wasielewski, M.R.; Chen, X. A perspective on four new porphyrin-based functional materials and devices. J. Porphyrins Phthalocyanines 2002, 06, 243–258. [Google Scholar] [CrossRef]
  5. Chalmers, M.; Notman, N. Living the nobel life. Chem. World 2009, 6. Available online: http://www.rsc.org/chemistryworld/Issues/2009/September/LivingtheNobellife.asp (accessed on 3 March 2016). [Google Scholar]
  6. Zhou, Q.L. Transition-metal catalysis and organocatalysis: Where can progress be expected? Angew. Chem. 2015, 2–4. [Google Scholar] [CrossRef] [PubMed]
  7. Carvalho, C.M.B.; Brocksom, T.J.; de Oliveira, K.T. Tetrabenzoporphyrins: Synthetic developments and applications. Chem. Soc. Rev. 2013, 42, 3302–3317. [Google Scholar] [CrossRef] [PubMed]
  8. Betoni Momo, P.; Pavani, C.; Baptista, M.S.; Brocksom, T.J.; de Oliveira, K.T. Chemical transformations and photophysical properties of meso-tetrathienyl-substituted porphyrin derivatives. Eur. J. Org. Chem. 2014, 2014, 4536–4547. [Google Scholar] [CrossRef]
  9. Carvalho, C.M.B.; Fujita, M.A.; Brocksom, T.J.; de Oliveira, K.T. Synthesis and photophysical evaluations of β-fused uracil-porphyrin derivatives. Tetrahedron 2013, 69, 9986–9993. [Google Scholar] [CrossRef]
  10. Perutz, M.F. Stereochemical mechanism of oxygen transport by haemoglobin. Proc. R. Soc. Lond. B. Biol. 1980, 208, 135–162. [Google Scholar] [CrossRef]
  11. Senge, M.O.; MacGowan, S.A.; O’Brien, J.M. Conformational control of cofactors in nature—The influence of protein-induced macrocycle distortion on the biological function of tetrapyrroles. Chem. Commun. 2015, 51, 17031–17063. [Google Scholar] [CrossRef] [PubMed]
  12. Milgrom, L.R. The Colours of Life—An Introduction to the Chemistry of Porphyrins and Related Compounds; Oxford University Press, Inc.: New York, NY, USA, 1997; Volume 22. [Google Scholar]
  13. Battersby, A.R. Tetrapyrroles: The pigments of life. Nat. Prod. Rep. 2000, 17, 507–526. [Google Scholar] [CrossRef] [PubMed]
  14. Sheftel, A.D.; Mason, A.B.; Ponka, P. The long history of iron in the Universe and in health and disease. Biochim. Biophys. Acta—Gen. Subj. 2012, 1820, 161–187. [Google Scholar] [CrossRef] [PubMed]
  15. Sheldon, R. Metalloporphyrins in Catalytic Oxidations; Sheldon, R., Ed.; Marcel Dekker: New York, NY, USA, 1994. [Google Scholar]
  16. Collman, J.; Zhang, X.; Lee, V.; Uffelman, E.; Brauman, J. Regioselective and enantioselective epoxidation catalysed by metalloporphyrins. Science 1993, 261, 1404–1411. [Google Scholar] [CrossRef] [PubMed]
  17. Groves, J.T.; Nemo, T.E.; Myers, R.S. Hydroxylation and epoxidation catalysed by iron-porphine complexes. Oxygen transfer from iodosylbenzene. J. Am. Chem. Soc. 1979, 101, 1032–1033. [Google Scholar] [CrossRef]
  18. Jiang, G.; Liu, Q.; Guo, C. Biomimetic oxidation of hydrocarbons with air over metalloporphyrins. In Biomimetic Based Applications; George, A., Ed.; InTech: Rijeka, Croatia, 2011; pp. 31–58. [Google Scholar]
  19. Meunier, B. Metalloporphyrins as versatile catalysts for oxidation reactions and oxidative DNA cleavage. Chem. Rev. 1992, 92, 1411–1456. [Google Scholar] [CrossRef]
  20. Chang, C.; Kuo, M. Reaction of iron(III) porphyrins and iodosoxylene. The active oxene complex of cytochrome P-450. J. Am. Chem. Soc. 1979, 101, 3413–3415. [Google Scholar] [CrossRef]
  21. Dolphin, D.; Traylor, T.G.; Xie, L.Y. Polyhaloporphyrins: Unusual ligands for metals and metal-catalysed oxidations. Acc. Chem. Res. 1997, 30, 251–259. [Google Scholar] [CrossRef]
  22. Traylor, T.G.; Tsuchiya, S. Perhalogenated tetraphenylhemins: Stable catalysts of high turnover catalytic hydroxylations. Inorg. Chem. 1987, 26, 1338–1339. [Google Scholar] [CrossRef]
  23. Haber, J.; Matachowski, L.; Pamin, K.; Poltowicz, J. The effect of peripheral substituents in metalloporphyrins on their catalytic activity in Lyons system. J. Mol. Catal. A Chem. 2003, 198, 215–221. [Google Scholar] [CrossRef]
  24. Srour, H.; le Maux, P.; Chevance, S.; Simonneaux, G. Metal-catalysed asymmetric sulfoxidation, epoxidation and hydroxylation by hydrogen peroxide. Coord. Chem. Rev. 2013, 257, 3030–3050. [Google Scholar] [CrossRef] [Green Version]
  25. Caron, S.; Dugger, R.W.; Ruggeri, S.G.; Ragan, J.A.; Brown Ripin, D.H. Large-scale oxidations in the pharmaceutical industry. Chem. Rev. 2006, 106, 2943–2989. [Google Scholar] [CrossRef] [PubMed]
  26. Lenoir, D. Selective oxidation of organic compounds-sustainable catalytic reactions with oxygen and without transition metals? Angew. Chem. Int. Ed. Engl. 2006, 45, 3206–3210. [Google Scholar] [CrossRef] [PubMed]
  27. Denisov, I.G.; Makris, T.M.; Sligar, S.G.; Schlichting, I. Structure and chemistry of cytochrome P450. Chem. Rev. 2005, 105, 2253–2277. [Google Scholar] [CrossRef] [PubMed]
  28. Che, C.-M.; Kar-yan lo, V.; Zhou, C.-Y.; Huang, J.-S. Selective functionalisation of saturated C–H bonds with metalloporphyrin catalysts. Chem. Soc. Rev. 2011, 40, 1950–1975. [Google Scholar] [CrossRef] [PubMed]
  29. Guo, C.-C.; Liu, X.-Q.; Liu, Q.; Liu, Y.; Chu, M.-F.; Lin, W.-Y. First industrial-scale biomimetic oxidation of hydrocarbon with air over metalloporphyrins as cytochrome P-450 monooxygenase model and its mechanistic studies. J. Porphyr. Phthalocyanines 2009, 13, 1250–1254. [Google Scholar] [CrossRef]
  30. Costas, M. Selective C–H oxidation catalysed by metalloporphyrins. Coord. Chem. Rev. 2011, 255, 2912–2932. [Google Scholar] [CrossRef]
  31. Nakagaki, S.; Ferreira, G.; Ucoski, G.; Dias de Freitas Castro, K. Chemical reactions catalysed by metalloporphyrin-based metal-organic frameworks. Molecules 2013, 18, 7279–7308. [Google Scholar] [CrossRef] [PubMed]
  32. Akagah, B.; Lormier, A.T.; Fournet, A.; Figadère, B. Oxidation of antiparasitic 2-substituted quinolines using metalloporphyrin catalysts: Scale-up of a biomimetic reaction for metabolite production of drug candidates. Org. Biomol. Chem. 2008, 6, 4494–4497. [Google Scholar] [CrossRef] [PubMed]
  33. Che, C.-M.; Huang, J.-S. Metalloporphyrin-based oxidation systems: From biomimetic reactions to application in organic synthesis. Chem. Commun. 2009, 27, 3996–4015. [Google Scholar] [CrossRef] [PubMed]
  34. Groves, J.T.; Myers, R.S. Catalytic asymmetric epoxidations with chiral iron porphyrins. J. Am. Chem. Soc. 1983, 105, 5791–5796. [Google Scholar] [CrossRef]
  35. Rose, E.; Andrioletti, B.; Zrig, S.; Quelquejeu-Ethève, M. Enantioselective epoxidation of olefins with chiral metalloporphyrin catalysts. Chem. Soc. Rev. 2005, 34, 573–583. [Google Scholar] [CrossRef] [PubMed]
  36. Gopalaiah, K. Chiral iron catalysts for asymmetric synthesis. Chem. Rev. 2013, 4, 3248–3296. [Google Scholar] [CrossRef] [PubMed]
  37. Lane, B.S.; Burgess, K. Metal-catalysed epoxidations of alkenes with hydrogen peroxide. Chem. Rev. 2003, 103, 2457–2473. [Google Scholar] [CrossRef] [PubMed]
  38. De Faveri, G.; Ilyashenko, G.; Watkinson, M. Recent advances in catalytic asymmetric epoxidation using the environmentally benign oxidant hydrogen peroxide and its derivatives. Chem. Soc. Rev. 2011, 40, 1722–1760. [Google Scholar] [CrossRef] [PubMed]
  39. Adam, W.; Stegmann, V.R.; Saha-mo, C.R. Regio- and diastereoselective epoxidation of chiral allylic alcohols catalysed by manganese (salen) and iron (porphyrin) complexes. J. Am. Chem. Soc. 1999, 121, 1879–1882. [Google Scholar] [CrossRef]
  40. Rose, E.; Ren, Q.Z.; Andrioletti, B. A Unique binaphthyl strapped iron-porphyrin catalyst for the enantioselective epoxidation of terminal olefins. Chem. Eur. J. 2004, 10, 224–230. [Google Scholar] [CrossRef] [PubMed]
  41. Stephenson, N.A.; Bell, A.T. Mechanistic insights into iron porphyrin-catalysed olefin epoxidation by hydrogen peroxide: Factors controlling activity and selectivity. J. Mol. Catal. A Chem. 2007, 275, 54–62. [Google Scholar] [CrossRef]
  42. Vilain-Deshayes, S.; Robert, A.; Maillard, P.; Meunier, B.; Momenteau, M. Enantioselective epoxidation of olefins by single-oxygen atom donors catalysed by manganese-glycoconjugated porphyrins. J. Mol. Catal. A Chem. 1996, 113, 23–34. [Google Scholar] [CrossRef]
  43. Le Maux, P.; Srour, H.F.; Simonneaux, G. Enantioselective water-soluble iron–porphyrin-catalysed epoxidation with aqueous hydrogen peroxide and hydroxylation with iodobenzene diacetate. Tetrahedron 2012, 68, 5824–5828. [Google Scholar] [CrossRef]
  44. Srour, H.; Le Maux, P.; Simonneaux, G. Enantioselective manganese-porphyrin-catalysed epoxidation and C–H hydroxylation with hydrogen peroxide in water/methanol solutions. Inorg. Chem. 2012, 51, 5850–5856. [Google Scholar] [CrossRef] [PubMed]
  45. Cunningham, I.D.; Danks, T.N.; Hay, J.N.; Hamerton, I.; Gunathilagan, S. Evidence for parallel destructive, and competitive epoxidation and dismutation pathways in metalloporphyrin-catalysed alkene oxidation by hydrogen peroxide. Tetrahedron 2001, 57, 6847–6853. [Google Scholar] [CrossRef]
  46. Nam, W.; Jin, S.W.; Lim, M.H.; Ryu, J.Y.; Kim, C. Anionic ligand effect on the nature of epoxidizing intermediates in iron porphyrin complex-catalysed epoxidation reactions. Inorg. Chem. 2002, 41, 3647–3652. [Google Scholar] [CrossRef] [PubMed]
  47. Nam, W.; Oh, S.; Sun, Y.J.; Kim, J.; Kim, W.; Woo, S.K. Factors affecting the catalytic epoxidation of olefins by iron porphyrin complexes and H2O2 in protic solvents. J. Org. Chem. 2003, 68, 7903–7906. [Google Scholar] [CrossRef] [PubMed]
  48. Zhang, J.L.; Zhou, H.B.; Huang, J.S.; Che, C.M. Dendritic ruthenium porphyrins: A new class of highly selective catalysts for alkene epoxidation and cyclopropanation. Chem. Eur. J. 2002, 8, 1554–1562. [Google Scholar] [CrossRef]
  49. Oae, S.; Watanabe, Y.; Fujimori, K. Biomimetic oxidation of organic sulfides with TPPFe(III)Cl/Imidazole/Hydrogen Peroxide. Tetrahedron Lett. 1982, 23, 1189–1192. [Google Scholar] [CrossRef]
  50. Kowalski, P.; Mitka, K.; Ossowska, K.; Kolarska, Z. Oxidation of sulfides to sulfoxides. Part 1: Oxidation using halogen derivatives. Tetrahedron 2005, 61, 1933–1953. [Google Scholar] [CrossRef]
  51. Legros, J.; Dehli, J.R.; Bolm, C. Applications of catalytic asymmetric sulfide oxidations to the syntheses of biologically active sulfoxides. Adv. Synth. Catal. 2005, 347, 19–31. [Google Scholar] [CrossRef]
  52. Wojaczyńska, E.; Wojaczyński, J. Enantioselective synthesis of sulfoxides: 2000–2009. Chem. Rev. 2010, 110, 4303–4356. [Google Scholar] [PubMed]
  53. Baciocchi, E.; Gerini, M.F.; Lapi, A. Synthesis of sulfoxides by the hydrogen peroxide induced oxidation of sulfides catalysed by iron tetrakis(pentafluorophenyl)porphyrin: Scope and chemoselectivity. J. Org. Chem. 2004, 69, 3586–3589. [Google Scholar] [CrossRef] [PubMed]
  54. Kaczorowska, K.; Kolarska, Z.; Mitka, K.; Kowalski, P. Oxidation of sulfides to sulfoxides. Part 2: Oxidation by hydrogen peroxide. Tetrahedron 2005, 61, 8315–8327. [Google Scholar] [CrossRef]
  55. Carreño, M.C.; Hernández-Torres, G.; Ribagorda, M.; Urbano, A. Enantiopure sulfoxides: Recent applications in asymmetric synthesis. Chem. Commun. 2009, 7345, 6129–6144. [Google Scholar] [CrossRef] [PubMed]
  56. Le Maux, P.; Simonneaux, G. First enantioselective iron-porphyrin-catalysed sulfide oxidation with aqueous hydrogen peroxide. Chem. Commun. 2011, 47, 6957–6959. [Google Scholar] [CrossRef] [PubMed]
  57. Srour, H.; Jalkh, J.; le Maux, P.; Chevance, S.; Kobeissi, M.; Simonneaux, G. Asymmetric oxidation of sulfides by hydrogen peroxide catalysed by chiral manganese porphyrins in water/methanol solution. J. Mol. Catal. A Chem. 2013, 370, 75–79. [Google Scholar] [CrossRef]
  58. Mori, T.; Santa, T.; Higuchi, T.; Mashino, T.; Hirobe, M. Oxygen activation by Iron(III)-porphyrin/NaBH4/Me4N·OH system as cytochrome P-450 Model. Oxygenation of olefin, N-dealkylation of tertiary amine, oxidation of sulfide, and oxidative cleavage of ether bond. Chem. Pharm. Bull. 1993, 41, 292–295. [Google Scholar] [CrossRef]
  59. Ortiz De Montellano, P.R. Hydrocarbon hydroxylation by cytochrome P450 enzymes. Chem. Rev. 2010, 110, 932–948. [Google Scholar] [CrossRef] [PubMed]
  60. Vinhado, F.S.; Prado-Manso, C.M.C.; Sacco, H.C.; Iamamoto, Y. Cationic manganese(III) porphyrins bound to a novel bis-functionalised silica as catalysts for hydrocarbons oxygenation by iodosylbenzene and hydrogen peroxide. J. Mol. Catal. A Chem. 2001, 174, 279–288. [Google Scholar] [CrossRef]
  61. Campestrini, S. Highly efficient cascade-oxygen-transfer from H2O2 to olefins mediated by halogenated carbonyl compounds and metalloporphyrins. J. Mol. Catal. A Chem. 2001, 171, 37–42. [Google Scholar] [CrossRef]
  62. Shilov, A.E.; Shul, G.B. Activation of C–H bonds by metal complexes. Chem. Rev. 1997, 97, 2879–2932. [Google Scholar] [CrossRef] [PubMed]
  63. Groves, J.T.; Viski, P. Asymmetric hydroxylation by a chiral iron porphyrin. J. Am. Chem. Soc. 1989, 111, 8537–8538. [Google Scholar] [CrossRef]
  64. Groves, J.; Viski, P. Asymmetric hydroxylation, epoxidation, and sulfoxidation catalysed by vaulted binaphthyl metalloporphyrins. J. Org. Chem. 1990, 55, 3628–3634. [Google Scholar] [CrossRef]
  65. Gross, Z.; Ini, S. Asymmetric catalysis by a chiral ruthenium porphyrin: Epoxidation, hydroxylation, and partial kinetic resolution of hydrocarbons. Org. Lett. 1999, 1, 2077–2080. [Google Scholar] [CrossRef]
  66. Zhang, R.; Yu, W.-Y.; Che, C.-M. Catalytic enantioselective oxidation of aromatic hydrocarbons with D4-symmetric chiral ruthenium porphyrin catalysts. Tetrahedron Asymmetry 2005, 16, 3520–3526. [Google Scholar] [CrossRef]
  67. De Freitas Silva, G.; da Silva, D.C.; Guimarães, A.S.; do Nascimento, E.; Rebouças, J.S.; de Araujo, M.P.; de Carvalho, M.E.M.D.; Idemori, Y.M. Cyclohexane hydroxylation by iodosylbenzene and iodobenzene diacetate catalysed by a new β-octahalogenated Mn–porphyrin complex: The effect of meso-3-pyridyl substituents. J. Mol. Catal. A Chem. 2007, 266, 274–283. [Google Scholar] [CrossRef]
  68. Ogawa, S.; Hosoi, K.; Iida, T.; Wakatsuki, Y.; Makino, M.; Fujimoto, Y.; Hofmann, A.F. Osmiumporphyrin-catalysed oxyfunctionalization and isomerization of natural (5β)-bile acids with tert-butyl hydroperoxide. Eur. J. Org. Chem. 2007, 2007, 3555–3563. [Google Scholar] [CrossRef]
  69. Ogawa, S.; Wakatsuki, Y.; Makino, M.; Fujimoto, Y.; Yasukawa, K.; Kikuchi, T.; Ukiya, M.; Akihisa, T.; Iida, T. Oxyfunctionalization of unactivated C-H bonds in triterpenoids with tert-butylhydroperoxide catalysed by meso-5,10,15,20-tetramesitylporphyrinate osmium(II) carbonyl complex. Chem. Phys. Lipids 2010, 163, 165–171. [Google Scholar] [CrossRef] [PubMed]
  70. Iida, T.; Ogawa, S.; Hosoi, K.; Makino, M.; Fujimoto, Y.; Goto, T.; Mano, N.; Goto, J.; Hofmann, A.F. Regioselective oxyfunctionalization of unactivated carbons in steroids by a model of cytochrome P-450: Osmiumporphyrin complex/tert-butyl hydroperoxide system. J. Org. Chem. 2007, 72, 823–830. [Google Scholar] [CrossRef] [PubMed]
  71. Ren, Q.-G.; Chen, S.-Y.; Zhou, X.-T.; Ji, H.-B. Highly efficient controllable oxidation of alcohols to aldehydes and acids with sodium periodate catalysed by water-soluble metalloporphyrins as biomimetic catalyst. Bioorg. Med. Chem. 2010, 18, 8144–8149. [Google Scholar] [CrossRef] [PubMed]
  72. Muzart, J. Palladium-catalysed oxidation of primary and secondary alcohols. Tetrahedron 2003, 59, 5789–5816. [Google Scholar] [CrossRef]
  73. Ji, H.; Zhou, X. Biomimetic homogeneous oxidation catalysed by metalloporphyrins with green oxidants. In Biomimetic Learning from Nature; Mukherjee, A., Ed.; InTech: Vienna, Austria, 2010; pp. 137–166. [Google Scholar]
  74. Adam, W.; Prikhodovski, S.; Roschmann, K.J.; Saha-Moller, C.R. Oxidation of aryl-substituted allylic alcohols by an optically active Fe(III)(porph*) catalyst: Enantioselectivity, diastereoselectivity and chemoselectivity in the epoxide versus enone formation. Tetrahedron Asymmetry 2001, 12, 2677–2681. [Google Scholar] [CrossRef]
  75. Burri, E.; Öhm, M.; Daguenet, C.; Severin, K. Site-isolated porphyrin catalysts in imprinted polymers. Chem. Eur. J. 2005, 11, 5055–5061. [Google Scholar] [CrossRef] [PubMed]
  76. Han, J.H.; Yoo, S.; Seo, S.; Hong, J.; Kim, K.; Kim, C. Biomimetic alcohol oxidations by an iron(III) porphyrin complex: Relevance to cytochrome P-450 catalytic oxidation and involvement of the two-state radical rebound mechanism. Dalton Trans. 2005, 2, 402–406. [Google Scholar] [CrossRef] [PubMed]
  77. Rezaeifard, A.; Jafarpour, M.; Moghaddam, G.K.; Amini, F. Cytochrome P-450 model reactions: Efficient and highly selective oxidation of alcohols with tetrabutylammonium peroxymonosulfate catalysed by Mn-porphyrins. Bioorg. Med. Chem. 2007, 15, 3097–3101. [Google Scholar] [CrossRef] [PubMed]
  78. Du, G.; Keith Woo, L. Alcohol oxidation with dioxygen mediated by oxotitanium porphyrin and related transition metal complexes. J. Porphyr. Phthalocyanines 2005, 9, 206–213. [Google Scholar] [CrossRef]
  79. Du, G.; Woo, L.K. Synthesis, characterization, and reactivity of group 4 metalloporphyrin diolate complexes. Organometallics 2003, 22, 450–455. [Google Scholar] [CrossRef]
  80. Ji, H.-B.; Yuan, Q.-L.; Zhou, X.-T.; Pei, L.-X.; Wang, L.-F. Highly efficient selective oxidation of alcohols to carbonyl compounds catalysed by ruthenium (III) meso-tetraphenylporphyrin chloride in the presence of molecular oxygen. Bioorg. Med. Chem. Lett. 2007, 17, 6364–6368. [Google Scholar] [CrossRef] [PubMed]
  81. Rebelo, S.L.; Simões, M.M.; Neves, M.G.P.M.; Cavaleiro, J.A. Oxidation of alkylaromatics with hydrogen peroxide catalysed by manganese(III) porphyrins in the presence of ammonium acetate. J. Mol. Catal. A Chem. 2003, 201, 9–22. [Google Scholar] [CrossRef]
  82. Zhang, J.L.; Che, C.M. Dichlororuthenium(IV) complex of meso-Tetrakis(2,6-dichlorophenyl) porphyrin: Active and robust catalyst for highly selective oxidation of arenes, unsaturated steroids, and electron-deficient alkenes by using 2,6-dichloropyridine N-oxide. Chem. Eur. J. 2005, 11, 3899–3914. [Google Scholar] [CrossRef] [PubMed]
  83. Twarowski, A.J. Porphyrin-sensitized formation of a polymer-bound endoperoxide in the solid state. J. Phys. Chem. 1988, 92, 6580–6584. [Google Scholar] [CrossRef]
  84. Slavětínská, L.; Mosinger, J.; Kubát, P. Supramolecular carriers of singlet oxygen: Photosensitized formation and thermal decomposition of endoperoxides in the presence of cyclodextrins. J. Photochem. Photobiol. A 2008, 195, 1–9. [Google Scholar] [CrossRef]
  85. Cossy, J.; Belotti, D. Efficient synthesis of substituted quinoline-5,8-quinones from 8-hydroxyquinolines by photooxygenation. Tetrahedron Lett. 2001, 42, 4329–4331. [Google Scholar] [CrossRef]
  86. Faustino, M.A.; Neves, M.G.; Cavaleiro, J.A.; Neumann, M.; Brauer, H.D.; Jori, G. meso-Tetraphenylporphyrin dimer derivatives as potential photosensitizers in photodynamic therapy. Part 2. Photochem. Photobiol. 2000, 72, 217–225. [Google Scholar] [CrossRef]
  87. Ogilby, P.R. Singlet oxygen: There is indeed something new under the sun. Chem. Soc. Rev. 2010, 39, 3181–3209. [Google Scholar] [CrossRef] [PubMed]
  88. Amarasekara, A.S. A new synthesis of quinoline-5,8-quinone. Synth. Commun. 1999, 29, 3063–3066. [Google Scholar] [CrossRef]
  89. Paddock, V.L.; Phipps, R.J.; Conde-Angulo, A.; Blanco-Martin, A.; Giró-Mañas, C.; Martin, L.J.; White, A.J.P.; Spivey, A.C. (±)-trans,cis-4-Hydroxy-5,6-di-O-isopropylidenecyclohex-2-ene-1-one: Synthesis and facile dimerization to decahydrodibenzofurans. J. Org. Chem. 2011, 76, 1483–1486. [Google Scholar] [CrossRef] [PubMed]
  90. Nicolaou, K.C.; Totokotsopoulos, S.; Giguère, D.; Sun, Y.; Sarlah, D. Total synthesis of epicoccin G. J. Am. Chem. Soc. 2011, 133, 8150–8153. [Google Scholar] [CrossRef] [PubMed]
  91. Murahashi, S.-I. Synthetic aspects of metal-catalysed oxidations of amines and related reactions. Angew. Chem. Int. Ed. Engl. 1995, 34, 2443–2465. [Google Scholar] [CrossRef]
  92. Matsumoto, M.; Kitano, Y. Singlet Oxygenation of 1-aminomethyl-1-tert-butyl-2-methoxy-2-(3-methoxy-phenyl)ethylenes: Marked effect of allylic nitrogen on the reaction pathways and chemoselectivity. Tetrahedron Lett. 1996, 37, 8191–8194. [Google Scholar] [CrossRef]
  93. Jiang, G.; Chen, J.; Huang, J.-S.; Che, C.-M. Highly efficient oxidation of amines to imines by singlet oxygen and its application in Ugi-type reactions. Org. Lett. 2009, 11, 4568–4571. [Google Scholar] [CrossRef] [PubMed]
  94. Nardello, V.; Hervé, M.; Alsters, P.L.; Aubry, J.-M. “Dark” singlet oxygenation of hydrophobic substrates in environmentally friendly microemulsions. Adv. Synth. Catal. 2002, 344, 184–191. [Google Scholar] [CrossRef]
  95. Coyle, E.E.; Joyce, K.; Nolan, K.; Oelgemöller, M. Green photochemistry: The use of microemulsions as green media in photooxygenation reactions. Green Chem. 2010, 12, 1544–1547. [Google Scholar] [CrossRef]
  96. Lu, H.; Zhang, X.P. Catalytic C–H functionalization by metalloporphyrins: Recent developments and future directions. Chem. Soc. Rev. 2011, 40, 1899–1909. [Google Scholar] [CrossRef] [PubMed]
  97. Song, C.E.; Lee, S.G. Supported chiral catalysts on inorganic materials. Chem. Rev. 2002, 102, 3495–3524. [Google Scholar] [CrossRef] [PubMed]
  98. Leadbeater, N.E.; Marco, M. Preparation of polymer-supported ligands and metal complexes for use in catalysis. Chem. Rev. 2002, 102, 3217–3274. [Google Scholar] [CrossRef] [PubMed]
  99. Ribeiro, S.M.; Serra, A.C.; Rocha Gonsalves, A.M.D. Covalently immobilized porphyrins on silica modified structures as photooxidation catalysts. J. Mol. Catal. A Chem. 2010, 326, 121–127. [Google Scholar] [CrossRef]
  100. Benaglia, M.; Danelli, T.; Fabris, F.; Sperandio, D.; Pozzi, G. Poly(ethylene glycol)-supported tetrahydroxyphenyl porphyrin: A convenient, recyclable catalyst for photooxidation reactions. Org. Lett. 2002, 4, 4229–4232. [Google Scholar] [CrossRef] [PubMed]
  101. Zhang, J.-L.; Huang, J.-S.; Che, C.-M. Oxidation chemistry of poly(ethylene glycol)-supported carbonylruthenium(II) and dioxoruthenium(VI)meso-Tetrakis(pentafluorophenyl)porphyrin. Chem. Eur. J. 2006, 12, 3020–3031. [Google Scholar] [CrossRef] [PubMed]
  102. Shiragami, T.; Makise, R.; Inokuchi, Y.; Matsumoto, J.; Inoue, H.; Yasuda, M. Efficient photocatalytic oxidation of cycloalkenes by dihydroxo(tetraphenylporphyrinato)antimony supported on silica gel under visible light irradiation. Chem. Lett. 2004, 33, 736–737. [Google Scholar] [CrossRef]
  103. Azenha, E.G.; Serra, A.C.; Pineiro, M.; Pereira, M.M.; Seixas de Melo, J.; Arnaut, L.G.; Formosinho, S.J.; Rocha Gonsalves, A.M.D. Heavy-atom effects on metalloporphyrins and polyhalogenated porphyrins. Chem. Phys. 2002, 280, 177–190. [Google Scholar] [CrossRef]
  104. Ribeiro, S.; Serra, A.C.; Gonsalves, A.R. Efficient solar photooxygenation with supported porphyrins as catalysts. ChemCatChem 2013, 5, 134–137. [Google Scholar] [CrossRef]
  105. Wegner, J.; Ceylan, S.; Kirschning, A. Flow chemistry—A key enabling technology for (multistep) organic synthesis. Adv. Synth. Catal. 2012, 354, 17–57. [Google Scholar] [CrossRef]
  106. Wiles, C.; Watts, P. Continuous flow reactors: A perspective. Green Chem. 2012, 14, 38–54. [Google Scholar] [CrossRef]
  107. Oelgemöller, M.; Shvydkiv, O. Recent advances in microflow photochemistry. Molecules 2011, 16, 7522–7550. [Google Scholar] [CrossRef] [PubMed]
  108. Griesbeck, A.G.; El-Idreesy, T.T. Solvent-free photooxygenation of 5-methoxyoxazoles in polystyrene nanocontainers doped with tetrastyrylporphyrine and protoporphyrine-IX. Photochem. Photobiol. Sci. 2005, 4, 205–209. [Google Scholar] [CrossRef] [PubMed]
  109. Bourne, R.A.; Han, X.; Poliakoff, M.; George, M.W. Cleaner continuous photo-oxidation using singlet oxygen in supercritical carbon dioxide. Angew. Chem. Int. Ed. 2009, 48, 5322–5325. [Google Scholar] [CrossRef] [PubMed]
  110. Bourne, R.A.; Han, X.; Chapman, A.O.; Arrowsmith, N.J.; Kawanami, H.; Poliakoff, M.; George, M.W. Homogeneous photochemical oxidation via singlet O2 in supercritical CO2. Chem. Commun. 2008, 37, 4457–4459. [Google Scholar] [CrossRef] [PubMed]
  111. Amara, Z.; Bellamy, J.F.B.; Horvath, R.; Miller, S.J.; Beeby, A.; Burgard, A.; Rossen, K.; Poliakoff, M.; George, M.W. Applying green chemistry to the photochemical route to artemisinin. Nat. Chem. 2015, 7, 489–495. [Google Scholar] [CrossRef] [PubMed]
  112. Han, X.; Bourne, R.A.; Poliakoff, M.; George, M.W. Immobilised photosensitisers for continuous flow reactions of singlet oxygen in supercritical carbon dioxide. Chem. Sci. 2011, 2, 1059–1067. [Google Scholar] [CrossRef]
  113. Loponov, K.N.; Lopes, J.; Barlog, M.; Astrova, E.V.; Malkov, A.V.; Lapkin, A.A. Optimization of a scalable photochemical reactor for reactions with singlet oxygen. Org. Process Res. Dev. 2014, 18, 1443–1454. [Google Scholar] [CrossRef]
  114. Lévesque, F.; Seeberger, P.H. Highly efficient continuous flow reactions using singlet oxygen as a “Green” reagent. Org. Lett. 2011, 13, 5008–5011. [Google Scholar] [CrossRef] [PubMed]
  115. 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]
  116. 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]
  117. De Oliveira, K.T.; Miller, L.Z.; McQuade, D.T. Exploiting photooxygenations mediated by porphyrinoid photocatalysts under continuous flow conditions. RSC Adv. 2016, 6, 12717–12725. [Google Scholar] [CrossRef]
  118. Momo, P.B.; Bellete, B.S.; Brocksom, T.J.; de Souza, R.O.M.A.; de Oliveira, K.T. Exploiting novel process windows for the synthesis of meso-substituted porphyrins under continuous flow conditions. RSC Adv. 2015, 5, 84350–84355. [Google Scholar] [CrossRef]
Figure 1. Porphyrin-derived pigments that participate in life processes.
Figure 1. Porphyrin-derived pigments that participate in life processes.
Molecules 21 00310 g001
Figure 2. Metalloporphyrin derivatives involved in life processes.
Figure 2. Metalloporphyrin derivatives involved in life processes.
Molecules 21 00310 g002
Scheme 1. Bioinspired aerobic C–H oxidation of cyclohexane by a μ-oxo-iron porphyrin.
Scheme 1. Bioinspired aerobic C–H oxidation of cyclohexane by a μ-oxo-iron porphyrin.
Molecules 21 00310 g008
Figure 3. Common synthetic β- and meso-substituted porphyrins.
Figure 3. Common synthetic β- and meso-substituted porphyrins.
Molecules 21 00310 g003
Figure 4. Classification of metalloporphyrin catalysts. (a) First generation; (b) Second generation; (c) Third generation. R1–R5 represent EWG or bulky groups and R6 halogens.
Figure 4. Classification of metalloporphyrin catalysts. (a) First generation; (b) Second generation; (c) Third generation. R1–R5 represent EWG or bulky groups and R6 halogens.
Molecules 21 00310 g004
Figure 5. Some important transformations catalysed by porphyrins.
Figure 5. Some important transformations catalysed by porphyrins.
Molecules 21 00310 g005
Scheme 2. First industrial-scale bioinspired oxidation of cyclohexane with CoTPP.
Scheme 2. First industrial-scale bioinspired oxidation of cyclohexane with CoTPP.
Molecules 21 00310 g009
Scheme 3. Metalloporphyrin catalysed bioinspired oxidation of 2-propylquinoline.
Scheme 3. Metalloporphyrin catalysed bioinspired oxidation of 2-propylquinoline.
Molecules 21 00310 g010
Figure 6. Chiral iron porphyrin catalysts for enantioselective epoxidation of olefins.
Figure 6. Chiral iron porphyrin catalysts for enantioselective epoxidation of olefins.
Molecules 21 00310 g006
Scheme 4. Enantioselective epoxidation of styrene derivatives.
Scheme 4. Enantioselective epoxidation of styrene derivatives.
Molecules 21 00310 g011
Scheme 5. Epoxidation of styrenes using manganese porphyrins as catalyst and H2O2.
Scheme 5. Epoxidation of styrenes using manganese porphyrins as catalyst and H2O2.
Molecules 21 00310 g012
Scheme 6. Different mechanisms in catalytic epoxidation depending on the conditions.
Scheme 6. Different mechanisms in catalytic epoxidation depending on the conditions.
Molecules 21 00310 g013
Scheme 7. Dendritic Ru-porphyrin catalysed epoxidation of cholesteryl esters.
Scheme 7. Dendritic Ru-porphyrin catalysed epoxidation of cholesteryl esters.
Molecules 21 00310 g014
Scheme 8. First example of a bioinspired sulfoxidation.
Scheme 8. First example of a bioinspired sulfoxidation.
Molecules 21 00310 g015
Scheme 9. Synthesis of sulfoxides from sulfides.
Scheme 9. Synthesis of sulfoxides from sulfides.
Molecules 21 00310 g016
Scheme 10. (a) Enantioselective sulfoxidation of styrenes using iron porphyrins as catalyst and hydrogen peroxide; (b) Metalloporphyrin-catalysed epoxidation for the enantioselective synthesis of Sulindac®.
Scheme 10. (a) Enantioselective sulfoxidation of styrenes using iron porphyrins as catalyst and hydrogen peroxide; (b) Metalloporphyrin-catalysed epoxidation for the enantioselective synthesis of Sulindac®.
Molecules 21 00310 g017
Scheme 11. Synthesis of sulfoxides from sulfides using the MTPPCl/NaBH4/Me4N.OH system.
Scheme 11. Synthesis of sulfoxides from sulfides using the MTPPCl/NaBH4/Me4N.OH system.
Molecules 21 00310 g018
Scheme 12. Enantioselective hydroxylation of ethylbenzene by a chiral iron-porphyrin.
Scheme 12. Enantioselective hydroxylation of ethylbenzene by a chiral iron-porphyrin.
Molecules 21 00310 g019
Scheme 13. Catalytic enantioselective hydroxylation of a tertiary alkane.
Scheme 13. Catalytic enantioselective hydroxylation of a tertiary alkane.
Molecules 21 00310 g020
Scheme 14. Ruthenium catalysed asymmetric alkane hydroxylation.
Scheme 14. Ruthenium catalysed asymmetric alkane hydroxylation.
Molecules 21 00310 g021
Scheme 15. Hydroxylation of ethylbenzene derivatives using a water-soluble iron-porphyrin as catalyst.
Scheme 15. Hydroxylation of ethylbenzene derivatives using a water-soluble iron-porphyrin as catalyst.
Molecules 21 00310 g022
Scheme 16. Representation of the oxidation process by the osmium-porphyrin [70].
Scheme 16. Representation of the oxidation process by the osmium-porphyrin [70].
Molecules 21 00310 g023
Scheme 17. Hydroxylation of dihydrofaradiol diacetate with the osmium porphyrin.
Scheme 17. Hydroxylation of dihydrofaradiol diacetate with the osmium porphyrin.
Molecules 21 00310 g024
Scheme 18. Oxidation of benzylic alcohols to carbonyl compounds with a Mn-porphyrin.
Scheme 18. Oxidation of benzylic alcohols to carbonyl compounds with a Mn-porphyrin.
Molecules 21 00310 g025
Scheme 19. Gram-scale oxidation of benzylic alcohols with Ru(TPP)Cl.
Scheme 19. Gram-scale oxidation of benzylic alcohols with Ru(TPP)Cl.
Molecules 21 00310 g026
Scheme 20. Oxidation of ketosteroids catalysed by ruthenium 2,6-TDCPP dichloride.
Scheme 20. Oxidation of ketosteroids catalysed by ruthenium 2,6-TDCPP dichloride.
Molecules 21 00310 g027
Figure 7. Representation of the mechanism of porphyrin photosensitized reactions.
Figure 7. Representation of the mechanism of porphyrin photosensitized reactions.
Molecules 21 00310 g007
Scheme 21. Synthesis of substituted quinoline-5,8-quinones by photooxygenation.
Scheme 21. Synthesis of substituted quinoline-5,8-quinones by photooxygenation.
Molecules 21 00310 g028
Scheme 22. Synthesis of decahydrodibenzofurans by using porphyrins as catalysts.
Scheme 22. Synthesis of decahydrodibenzofurans by using porphyrins as catalysts.
Molecules 21 00310 g029
Scheme 23. Synthesis of bisendoperoxide intermediates in the total synthesis of (a) epicoccin G and (b) 8,8’-epi-ent-rostratin B using TPP as photosensitizer.
Scheme 23. Synthesis of bisendoperoxide intermediates in the total synthesis of (a) epicoccin G and (b) 8,8’-epi-ent-rostratin B using TPP as photosensitizer.
Molecules 21 00310 g030
Scheme 24. Photooxidation of secondary benzylamines to imines.
Scheme 24. Photooxidation of secondary benzylamines to imines.
Molecules 21 00310 g031
Scheme 25. Schematic representation of a “green” microemulsion Juglone synthesis.
Scheme 25. Schematic representation of a “green” microemulsion Juglone synthesis.
Molecules 21 00310 g032
Scheme 26. PEG-supported tetrahydroxyphenyl porphyrin for photooxidation reactions.
Scheme 26. PEG-supported tetrahydroxyphenyl porphyrin for photooxidation reactions.
Molecules 21 00310 g033
Scheme 27. PEG-supported ruthenium porphyrin for oxidation reactions.
Scheme 27. PEG-supported ruthenium porphyrin for oxidation reactions.
Molecules 21 00310 g034
Scheme 28. Different photocatalytic properties of antimony SbTPP(OH)2\SiO2 supported over silica.
Scheme 28. Different photocatalytic properties of antimony SbTPP(OH)2\SiO2 supported over silica.
Molecules 21 00310 g035
Scheme 29. Solar photooxygenation with supported porphyrins on Merrifield-modified polymers.
Scheme 29. Solar photooxygenation with supported porphyrins on Merrifield-modified polymers.
Molecules 21 00310 g036
Scheme 30. One-pot semi-synthesis of artemisinin using immobilized TPP on Amberlyst-15.
Scheme 30. One-pot semi-synthesis of artemisinin using immobilized TPP on Amberlyst-15.
Molecules 21 00310 g037
Scheme 31. Continuous flow reactions with PVC amino-functionalized covalently bonded porphyrin.
Scheme 31. Continuous flow reactions with PVC amino-functionalized covalently bonded porphyrin.
Molecules 21 00310 g038
Scheme 32. Scalable photochemical conversion of α-pinene to pinocarvone.
Scheme 32. Scalable photochemical conversion of α-pinene to pinocarvone.
Molecules 21 00310 g039
Scheme 33. Synthetic route to artemisinin from artemisinic acid.
Scheme 33. Synthetic route to artemisinin from artemisinic acid.
Molecules 21 00310 g040
Scheme 34. Photo-oxidation of naphthols under continuous flow conditions.
Scheme 34. Photo-oxidation of naphthols under continuous flow conditions.
Molecules 21 00310 g041

Share and Cite

MDPI and ACS Style

Barona-Castaño, J.C.; Carmona-Vargas, C.C.; Brocksom, T.J.; De Oliveira, K.T. Porphyrins as Catalysts in Scalable Organic Reactions. Molecules 2016, 21, 310. https://doi.org/10.3390/molecules21030310

AMA Style

Barona-Castaño JC, Carmona-Vargas CC, Brocksom TJ, De Oliveira KT. Porphyrins as Catalysts in Scalable Organic Reactions. Molecules. 2016; 21(3):310. https://doi.org/10.3390/molecules21030310

Chicago/Turabian Style

Barona-Castaño, Juan C., Christian C. Carmona-Vargas, Timothy J. Brocksom, and Kleber T. De Oliveira. 2016. "Porphyrins as Catalysts in Scalable Organic Reactions" Molecules 21, no. 3: 310. https://doi.org/10.3390/molecules21030310

Article Metrics

Back to TopTop