8
Bio-Based Nitriles from the Heterogeneously Catalyzed Oxidative Decarboxylation of Amino Acids Laurens Claes, [a] Roman Matthessen, [a] Ine Rombouts, [b] Ivo Stassen, [a] Trees De Baerdemaeker, [a] Diederik Depla, [c] Jan A. Delcour, [b] Bert Lagrain, [a] and Dirk E. De Vos* [a] Introduction Decarboxylation, or the selective elimination of a carboxylate as CO 2 , is an excellent strategy to reduce excess functionality in biomass-derived compounds, particularly amino acids. [1] Although pure amino acids are produced mainly by microbial fermentation or biocatalytic synthesis, [2] protein-rich side prod- ucts from the agro-industry or biorefinery provide, after hydrol- ysis, cheap alternative sources. [3] Examples are wheat gluten from starch isolation, [4] dried distiller’s grains with solubles (DDGS) from bio-ethanol production, [5] sugarbeet vinasse, poul- try feather meal, [6] and press cake from jatropha and castor oil production, [7] in which the average protein content accounts for at least 20–40 wt % of the dry mass. The isolated proteins are often of too low quality for food/feed applications and, therefore, they are attractive as a feedstock to produce bio- based N-containing chemicals, such as amine or nitrile inter- mediates. [8] A valorization approach based on protein hydroly- sis and amino acid oxidative decarboxylation would be highly complementary to the current efforts in depolymerization and chemical modification of (hemi)cellulose, lignin, or lipids for bio-based fuels and chemicals production. [9] In addition, the isolation of nitrile products after oxidative decarboxylation is much easier than the isolation of zwitterionic amino acids from protein hydrolyzates by electrodialysis. [10] Synthetic methods for the oxidative decarboxylation of amino acids often rely on hypohalite species, which can be produced from commercial reagents such as N-bromosuccini- mide, [11] chloramine-T, [12] trichloroisocyanuric acid, [13] and NaOCl, [14] but the atom efficiency of these reagents is low and product purification may be complicated by organic residues or salts. Alternatively, halides such as bromide have been oxi- dized in situ either electrochemically [15] or by using bromoper- oxidase enzymes and H 2 O 2 as the terminal oxidant. [16] However, the long-term stability of such enzymes, and particularly the ni- trile versus aldehyde selectivity in amino acid conversion remain problematic. We previously showed that peroxotungstate [W(O 2 ) n (O) 4n 2 , n = 1, 2] immobilized on a layered double hydroxide (LDH) support acts as a catalyst for in situ bromide oxidation using H 2 O 2 and enables bromide-assisted olefin epoxidation as well as the oxidative bromination of aromatics. [17] LDH materials— generally represented as [M II 1x M III x (OH) 2 ](A z ) x/z (M = di- or triva- lent metal cation, e.g., Mg 2 + , Ni 2 + , or Al 3 + ;A = exchanged anion)—are excellent supports for (peroxo)tungstate species as they exhibit an anion-exchange capacity (AEC) as a result of the isomorphous substitution and, hence, the permanent posi- tive charges on the edge-sharing M II and M III hydroxide octahe- dra in the brucite-type layers. Here we show that this heterogeneous catalyst for halide ox- idation can perform the oxidative decarboxylation of a broad range of amino acids under mild conditions, with remarkably high nitrile selectivity and with the preservation of other func- The oxidative decarboxylation of amino acids to nitriles was achieved in aqueous solution by in situ halide oxidation using catalytic amounts of tungstate exchanged on a [Ni,Al] layered double hydroxide (LDH), NH 4 Br, and H 2 O 2 as the terminal oxi- dant. Both halide oxidation and oxidative decarboxylation were facilitated by proximity effects between the reactants and the LDH catalyst. A wide range of amino acids was con- verted with high yields, often > 90 %. The nitrile selectivity was excellent, and the system is compatible with amide, alcohol, and in particular carboxylic acid, amine, and guanidine func- tional groups after appropriate neutralization. This heterogene- ous catalytic system was applied successfully to convert a pro- tein-rich byproduct from the starch industry into useful bio- based N-containing chemicals. [a] L. Claes, R. Matthessen, I. Stassen, T. De Baerdemaeker, Dr. B. Lagrain, Prof. D. E. De Vos Centre for Surface Chemistry and Catalysis Department of Microbial and Molecular Systems KU Leuven Kasteelpark Arenberg 23, 3001 Heverlee (Belgium) E-mail: [email protected] [b] Dr. I. Rombouts, Prof. J. A. Delcour Laboratory of Food Chemistry and Biochemistry Department of Microbial and Molecular Systems KU Leuven Kasteelpark Arenberg 20, 3001 Heverlee (Belgium) [c] Prof. D. Depla Research Group DRAFT Department of Solid State Sciences Ghent University Krijgslaan 281(S1), 9000 Gent (Belgium) Supporting Information for this article is available on the WWW under http://dx.doi.org/10.1002/cssc.201402801. ChemSusChem 2015, 8, 345 – 352 # 2015 Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim 345 Full Papers DOI: 10.1002/cssc.201402801

Bio-Based Nitriles from the Heterogeneously Catalyzed Oxidative … · 2017-08-17 · cilitates both bromide oxidation and amino acid oxi-dative decarboxylation. By analogy with what

  • Upload
    others

  • View
    5

  • Download
    0

Embed Size (px)

Citation preview

Page 1: Bio-Based Nitriles from the Heterogeneously Catalyzed Oxidative … · 2017-08-17 · cilitates both bromide oxidation and amino acid oxi-dative decarboxylation. By analogy with what

Bio-Based Nitriles from the Heterogeneously CatalyzedOxidative Decarboxylation of Amino AcidsLaurens Claes,[a] Roman Matthessen,[a] Ine Rombouts,[b] Ivo Stassen,[a]

Trees De Baerdemaeker,[a] Diederik Depla,[c] Jan A. Delcour,[b] Bert Lagrain,[a] andDirk E. De Vos*[a]

Introduction

Decarboxylation, or the selective elimination of a carboxylateas CO2, is an excellent strategy to reduce excess functionalityin biomass-derived compounds, particularly amino acids.[1]

Although pure amino acids are produced mainly by microbialfermentation or biocatalytic synthesis,[2] protein-rich side prod-ucts from the agro-industry or biorefinery provide, after hydrol-ysis, cheap alternative sources.[3] Examples are wheat glutenfrom starch isolation,[4] dried distiller’s grains with solubles(DDGS) from bio-ethanol production,[5] sugarbeet vinasse, poul-try feather meal,[6] and press cake from jatropha and castor oilproduction,[7] in which the average protein content accountsfor at least 20–40 wt % of the dry mass. The isolated proteinsare often of too low quality for food/feed applications and,therefore, they are attractive as a feedstock to produce bio-based N-containing chemicals, such as amine or nitrile inter-mediates.[8] A valorization approach based on protein hydroly-sis and amino acid oxidative decarboxylation would be highly

complementary to the current efforts in depolymerization andchemical modification of (hemi)cellulose, lignin, or lipids forbio-based fuels and chemicals production.[9] In addition, theisolation of nitrile products after oxidative decarboxylation ismuch easier than the isolation of zwitterionic amino acids fromprotein hydrolyzates by electrodialysis.[10]

Synthetic methods for the oxidative decarboxylation ofamino acids often rely on hypohalite species, which can beproduced from commercial reagents such as N-bromosuccini-mide,[11] chloramine-T,[12] trichloroisocyanuric acid,[13] andNaOCl,[14] but the atom efficiency of these reagents is low andproduct purification may be complicated by organic residuesor salts. Alternatively, halides such as bromide have been oxi-dized in situ either electrochemically[15] or by using bromoper-oxidase enzymes and H2O2 as the terminal oxidant.[16] However,the long-term stability of such enzymes, and particularly the ni-trile versus aldehyde selectivity in amino acid conversionremain problematic.

We previously showed that peroxotungstate [W(O2)n(O)4�n2�,

n = 1, 2] immobilized on a layered double hydroxide (LDH)support acts as a catalyst for in situ bromide oxidation usingH2O2 and enables bromide-assisted olefin epoxidation as wellas the oxidative bromination of aromatics.[17] LDH materials—generally represented as [MII

1�xMIII

x(OH)2](Az�)x/z (M = di- or triva-lent metal cation, e.g. , Mg2+ , Ni2 + , or Al3+ ; A = exchangedanion)—are excellent supports for (peroxo)tungstate species asthey exhibit an anion-exchange capacity (AEC) as a result ofthe isomorphous substitution and, hence, the permanent posi-tive charges on the edge-sharing MII and MIII hydroxide octahe-dra in the brucite-type layers.

Here we show that this heterogeneous catalyst for halide ox-idation can perform the oxidative decarboxylation of a broadrange of amino acids under mild conditions, with remarkablyhigh nitrile selectivity and with the preservation of other func-

The oxidative decarboxylation of amino acids to nitriles wasachieved in aqueous solution by in situ halide oxidation usingcatalytic amounts of tungstate exchanged on a [Ni,Al] layereddouble hydroxide (LDH), NH4Br, and H2O2 as the terminal oxi-dant. Both halide oxidation and oxidative decarboxylationwere facilitated by proximity effects between the reactantsand the LDH catalyst. A wide range of amino acids was con-

verted with high yields, often >90 %. The nitrile selectivity wasexcellent, and the system is compatible with amide, alcohol,and in particular carboxylic acid, amine, and guanidine func-tional groups after appropriate neutralization. This heterogene-ous catalytic system was applied successfully to convert a pro-tein-rich byproduct from the starch industry into useful bio-based N-containing chemicals.

[a] L. Claes, R. Matthessen, I. Stassen, T. De Baerdemaeker, Dr. B. Lagrain,Prof. D. E. De VosCentre for Surface Chemistry and CatalysisDepartment of Microbial and Molecular SystemsKU LeuvenKasteelpark Arenberg 23, 3001 Heverlee (Belgium)E-mail : [email protected]

[b] Dr. I. Rombouts, Prof. J. A. DelcourLaboratory of Food Chemistry and BiochemistryDepartment of Microbial and Molecular SystemsKU LeuvenKasteelpark Arenberg 20, 3001 Heverlee (Belgium)

[c] Prof. D. DeplaResearch Group DRAFTDepartment of Solid State SciencesGhent UniversityKrijgslaan 281(S1), 9000 Gent (Belgium)

Supporting Information for this article is available on the WWW underhttp://dx.doi.org/10.1002/cssc.201402801.

ChemSusChem 2015, 8, 345 – 352 � 2015 Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim345

Full PapersDOI: 10.1002/cssc.201402801

Page 2: Bio-Based Nitriles from the Heterogeneously Catalyzed Oxidative … · 2017-08-17 · cilitates both bromide oxidation and amino acid oxi-dative decarboxylation. By analogy with what

tional groups in the amino acids(Figure 1). Moreover, a crudeamino acid mixture, obtained di-rectly from the acidic hydrolysisof wheat gluten, was convertedsuccessfully, which demonstratesthe applicability and robustnessof this powerful system forwaste protein valorization.

Results and Discussion

In the present study, a typical[Ni,Al]-LDH-WO4 catalyst wasprepared by the co-precipitationof Ni2 + and Al3 + nitrates inaqueous alkaline solution(pH 8.5) under ambient condi-tions, followed by anion-ex-change with Na2WO4. The crys-tallographic structure of the ta-kovite-type LDH was confirmedby powder X-ray diffraction(PXRD) (Figure 2 a).[18] SEM re-veals that the catalyst consists ofvery small crystallites that areaggregated into larger porousgranules in the range of 1–2 mmwithout a clear morphology (Fig-ure 2 b). This observation corre-sponds well with the broadpeaks observed in the diffractionpatterns. The BET surface area of72 m2 g�1, determined by N2

physisorption at 77 K, suggeststhat only the outer surface areaof the catalyst is accessible (Fig-ure 2 c). The chemical composi-tion and textural properties ofthe catalyst and its precursor areshown in Table 1. Elementalanalysis by inductively coupled

plasma atomic emission spectroscopy (ICP-AES) showed thatthe catalytically active material has a tungstate loading of0.18 mmol g�1, which corresponds to 12 % of the total AEC. Asno precautions were taken to exclude CO2 during catalystpreparation, the final material was rich in carbonate, which islocated preferentially in the interlayer gallery. This evidencesthat tungstate is located mainly at accessible sites on the ex-ternal surface, for example, at the outer basal planes or at crys-tal edges as confirmed by the comparison of ICP-AES and X-ray photoelectron spectroscopy (XPS). Additional details oncatalyst characterization are given in the Supporting Informa-tion (Figures S1–S4, Tables S1 and S2).

The catalytic system was optimized for the oxidative decar-boxylation of phenylalanine (2 a ; Table 2). Efficient bromide oxi-

Figure 1. Heterogeneous catalytic system for the oxidative decarboxylationof amino acids, for example, 2 a to 2 a’.

Figure 2. Characterization of the tungstate-exchanged [Ni,Al]-LDH catalyst : a) PXRD pattern, b) SEM image, andc) N2 physisorption curve at 77 K: as-synthesized catalyst (c), spent catalyst after oxidative decarboxylation of2 a (a).

ChemSusChem 2015, 8, 345 – 352 www.chemsuschem.org � 2015 Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim346

Full Papers

Page 3: Bio-Based Nitriles from the Heterogeneously Catalyzed Oxidative … · 2017-08-17 · cilitates both bromide oxidation and amino acid oxi-dative decarboxylation. By analogy with what

dation and hence a high yield of phenylacetonitrile (2 a’) wereonly achieved by using the [Ni,Al]-LDH-supported tungstatecatalyst (Table 2, entries 1–4). This is consistent with the hy-pothesis that the LDH exerts a charge-shielding effect, whichthus allows the efficient oxidation of bromide anions by anion-ic peroxotungstate species. Moreover, electrostatic attractionbetween the carboxylate groups and the positively chargedLDH surface brings the amino acids into close proximity to thesurface. For an amino acid that contains a hydrophobic sidechain such as 2 a, the accumulation of carboxylatesnear the surface was evidenced by a clear increase incatalyst volume (Figure 3 a and b), even if no true in-tercalation in the interlayer gallery takes place, asshown by PXRD (Figure 2 a).[19] Upon addition of H2O2

and near-complete conversion of 2 a, the catalyst re-verts to its original volume (Figure 3 c and d), whichshows that the solid has a much lower affinity for theneutral 2 a’ product than for the amino acid reactant.Thus, the LDH exerts a double proximity effect and fa-cilitates both bromide oxidation and amino acid oxi-dative decarboxylation.

By analogy with what is known for enzyme-[16a] andNaOBr-mediated oxidative decarboxylation,[20] a mech-

anism can be proposed (Scheme 1). First, an N-bromo aminoacid derivative is formed, which reacts further through twoparallel pathways. Product selectivity is strongly dependent onthe pH: in neutral or slightly acidic conditions a second bromi-nation occurs, and nitriles are produced through the consecu-tive release of two bromide ions from an N,N-dibromo aminoacid intermediate. However, if the conditions are more alkaline,bromide elimination from the N-bromo amino acid is preferredover a second bromination, and aldehydes are obtained by thehydrolysis of the a-imino acid intermediate. Our catalyticsystem is much more selective for nitrile formation from 2 athan NaOCl-[21] and enzyme-mediated reactions.[16b] In the lattercase, the aldehyde accounts for at least 17 % of the products,whereas our nitrile selectivity of >99 % keeps virtually all nitro-gen in a useful product. Here, favorable, rather neutral condi-tions and, hence, a high nitrile selectivity (>99 %) were ach-

Table 1. Chemical composition and textural properties of the as-synthesized [Ni,Al]-LDH materials.

Material Ni/Al/W Chemical composition AEC[a] S[b] V[c] D[d]

ICP-AES XPS [meq g�1] [m2 g�1] [cm3 g�1] [nm]

[Ni,Al]-LDH 1.74/1/– 1.69/1/– [Ni0.64Al0.36(OH)2](NO3)0.22(CO3)0.07 3.0 73�1 0.41 25�2[Ni,Al]-LDH-WO4 1.76/1/0.061 1.67/1/0.083 [Ni0.64Al0.36(OH)2](NO3)0.04(CO3)0.14(WO4)0.02 3.0 72�1 0.34 21�2

[a] Anion-exchange capacity. [b] BET surface area. [c] Barrett–Joyner–Halenda (BJH) pore volume. [d] BJH average pore diameter.

Table 2. Optimization of the catalytic system for the oxidative decarboxy-lation of 2 a.[a]

Entry Catalyst Bromide salt Yield[b]

[%]

1 – NH4Br 02 Na2WO4 NH4Br 03 [Ni,Al]-LDH + Na2WO4 NH4Br 154 [Ni,Al]-LDH-WO4 NH4Br 885 [Mg,Al]-LDH-WO4 NH4Br 486 [Ni,Al]-LDH-WO4 NaBr 307 [Ni,Al]-LDH-WO4 – 48 [Ni,Al]-LDH-WO4 NH4Br (1.5 equiv.) 899 [Ni,Al]-LDH-WO4 NH4Br (2 equiv.) 7410[c] [Ni,Al]-LDH-WO4 NH4Br 81

[a] Reaction conditions: 2 a (0.5 mmol), bromide salt (0.5 mmol), catalyst(0.1 g; 18 mmol WO4

2�), CH3CN (4 mL), room temperature. H2O2 (5 mmol)diluted in CH3CN (total volume: 4 mL); addition rate: 0.33 mmol h�1.[b] Yield of 2 a’ determined by GC using benzonitrile as the internal stan-dard. The selectivity to phenylacetaldehyde is always <1 %. [c] Same as[a], except for NH4Br (0.125 mmol), [Ni,Al]-LDH-WO4 (25 mg), CH3CN(2 mL); substrate: W = 110; addition rate of H2O2: 0.66 mmol h�1.

Figure 3. Adsorption of 2 a on the [Ni,Al]-LDH-WO4 catalyst during oxidativedecarboxylation: a) initial reaction mixture that contained 2 a, NH4Br, and[Ni,Al]-LDH-WO4 in CH3CN, b) after 12 h stirring with evident swelling, c) afterthe addition of 3 equiv. H2O2, and d) after the addition of 10 equiv. H2O2

(near-complete conversion of 2 a).

Scheme 1. Mechanism for hypobromite-mediated oxidative decarboxylation of aminoacids to aldehydes and nitriles.

ChemSusChem 2015, 8, 345 – 352 www.chemsuschem.org � 2015 Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim347

Full Papers

Page 4: Bio-Based Nitriles from the Heterogeneously Catalyzed Oxidative … · 2017-08-17 · cilitates both bromide oxidation and amino acid oxi-dative decarboxylation. By analogy with what

ieved by combining a slightly basic [Ni,Al]-LDH-WO4 catalystand NH4Br, a weak acid (Table 2, entry 4). Strikingly, the yield of2 a’ was much lower if a [Mg,Al]-LDH support with more pro-nounced basic properties was used, as evidenced by acid–basetitrations (Table S3), or NaBr (Table 2, entries 5 and 6). This canbe explained by the spontane-ous disproportionation of perox-otungstate and the hypobro-mite-induced decomposition ofH2O2 to singlet oxygen (1O2),which both proceed faster athigher pH;[22] the slow additionof H2O2 is recommended tocompensate for the loss in oxi-dant efficiency. At an H2O2/phe-nylalanine ratio of 6, an accepta-ble compromise was reached be-tween a good yield of 2 a’ (83 %)and sufficient H2O2 efficiency(28 %) (Figure 4). Acidification ofthe medium, observed if N-bro-mosuccinimide is used for exam-ple, is avoided here by usingH2O2 as the terminal oxidant forbromide oxidation, and only(neutral) water is produced asa byproduct.

Although the preferred oxida-tion reaction requires two hypo-bromite equivalents, bromide re-oxidation by the [Ni,Al]-LDH-WO4

catalyst allows the system toachieve a high nitrile yield withcatalytic amounts of NH4Br(Table 2, entries 7–9). In addition,comparable yields were ob-tained if the 2 a/WO4

2� ratio wasincreased up to >100 while theNH4Br/WO4

2� ratio was kept con-stant (entry 10).

The heterogeneity of the[Ni,Al]-LDH-WO4 catalyst wasconfirmed by separating the

solid material at an intermediate conversion of 2 a : the filtrateshowed no residual activity. ICP-AES showed that leaching inacetonitrile was limited to 0.01 % of tungstate and to <0.05 %of the support constituents (Table S4). Moreover, the catalystretains its textural properties (Figures S5 and S6). The activitywas maintained in three successive runs with recycled catalystwith 2 a’ yields of 85, 81, and 82 %.

The substrate scope was expanded successfully to a widerange of naturally occurring amino acids, and many reactionswere run in water (Table 3). Amino acids with an aliphatic sidechain, such as alanine (3 a), valine (3 b), leucine (3 c), isoleucine(3 d), and norleucine (3 e) were converted with excellent yields,often >95 %. Chirality in the side chain of 3 d was preserved inthe product (S)-2-methylbutyronitrile (3 d’). The lack of reactivi-ty of b-alanine and glycine provides further support for theproposed mechanism: CO2 elimination requires a conjugateda-imino acid with an electron-donating substituent, which isonly formed from an a-substituted a-amino acid moietyduring amine oxidation (Scheme 1). Next, amino acids that

Figure 4. Oxidative decarboxylation of 2 a using various H2O2/amino acidratios.

Table 3. Oxidative decarboxylation of various amino acids to nitriles.[a]

Amino acid Nitrile Yield (selectivity)[b]

[%]Amino acid Nitrile Yield (selectivity)[b]

[%]

99 (99) 99 (99)

91 (99) 72 (73)

99 (99) 99 (99)

94 (99) 45 (71)

99 (99) 90 (99)

94 (99) 89 (99)

60 (69) 72 (90)[c]

[a] Reaction conditions: amino acid (0.5 mmol), NH4Br (0.5 mmol), [Ni,Al]-LDH-WO4 (0.1 g; 18 mmol WO42�), H2O

(4 mL), room temperature. H2O2 (5 mmol) diluted in H2O (total volume: 4 mL); addition rate: 0.33 mmol h�1.[b] Yield and selectivity determined by 1H NMR spectroscopy. [c] The catalytic system was applied for S-oxida-tion and subsequent oxidative decarboxylation: (1) methionine (0.5 mmol), [Ni,Al]-LDH-WO4 (0.1 g; 18 mmolWO4

2�), CH3OH/H2O (90:10, 4 mL), room temperature. H2O2 (5 mmol) diluted in CH3OH (total volume: 4 mL); ad-dition rate: 0.33 mmol h�1. (2) Addition of NH4Br (0.5 mmol) to the reaction mixture. H2O2 (5 mmol) diluted inCH3OH (total volume: 4 mL); addition rate: 0.33 mmol h�1.

ChemSusChem 2015, 8, 345 – 352 www.chemsuschem.org � 2015 Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim348

Full Papers

Page 5: Bio-Based Nitriles from the Heterogeneously Catalyzed Oxidative … · 2017-08-17 · cilitates both bromide oxidation and amino acid oxi-dative decarboxylation. By analogy with what

contain an alcohol or an amide in the side chain, for example,serine (3 f), threonine (3 g), homoserine (3 h), and glutamine(3 i), were oxidized with good yields. For amino acids that bearpH-responsive functional groups, neutralization was needed topreserve the optimal pH conditions. The transformation of glu-tamic acid to its monosodium salt (3 j) improves the solubilityin water significantly, and 3-cyanopropionate (3 j’) was ob-tained with 99 % yield. This bifunctional nitrile is a bio-basedprecursor to acrylonitrile,[14] succinonitrile,[23] and adiponitri-le.[15b] However, the HCl salt of lysine (3 k) was converted withpoor selectivity because of hypochlorite-mediated side reac-tions. Neutralization by HBr is preferred over the use of HCl toavoid hypohalite scrambling. In this way, lysine (3 l) and argi-nine (3 m) were converted with 90 % yield. The selective syn-thesis of w-aminonitriles (e.g. , 3 l’) from amino acids is a majorachievement, as it is challenging to access these polymer pre-cursors through the metal-catalyzed hydrogenation of a,w-di-nitriles.[24] Finally, methionine (3 n) was transformed to a nitrile-sulfone derivative 3 n’ in 72 % yield through a two-step proce-dure using the same catalyst as the tungstate-exchanged[Ni,Al]-LDH catalyzes both S-oxidation[25] and oxidative decar-boxylation.

Although a variety of amino acids were converted, some lim-itations to the substrate scope have been identified. First, if anelectron-withdrawing group was present on the b-carbonatom, as in 3 g, and even more so in aspartic acid and aspara-gine, the reactivity of the a-amino acid moiety was diminished

and lower conversions were observed. In addition, moleculesthat contain a phenolic or heteroaromatic moiety, in particulartyrosine, tryptophan, and histidine, do not deliver the nitrilesbecause the electrophilic ‘Br+ ’ species are consumed in the ox-idative bromination of the aromatic moiety.[26] Such reactivity issimilar to that observed for bromoperoxidase enzymes.[27]

Competition experiments between 2 a and 4-ethylphenol orimidazole showed clearly that the 2 a’ yield was reduced drasti-cally in comparison with the 88 % yield obtained in the ab-sence of any additive (Table S5).

The extension of this reaction to the valorization of real pro-tein side streams required only a minor adaptation to conven-tional protein hydrolysis (Supporting Information). As an exam-ple of a well-characterized protein source, wheat gluten witha high protein content (83 wt %)[28] was hydrolyzed using HBrfollowed by [Ni,Al]-LDH-WO4-catalyzed oxidative decarboxyla-tion (Table 4, Figure 5). Although the reaction resulted ina high conversion (>95 %) for most amino acids in the glutenhydrolyzate, the trends in reactivity observed earlier for pureamino acids (Table 3) were confirmed. In addition, although ty-rosine and histidine were present in the amino acid pool, in

Table 4. HBr-mediated hydrolysis of wheat gluten[a] and subsequent oxi-dative decarboxylation.[b]

Amino Hydrolysis by HBr Oxidative decarboxylation Conversionacid[c] [mmol g�1] [mmol g�1] [%]

Ala 302 0 >99Asx[d] 251 0 >99Cys[e] 35 0 >99Glx[f] 2287 0 >99Gly 424 325 23His 177 0 >99Ile 304 0 >99Leu 599 0 >99Lys 129 0 >99Met 45 0 >99Phe 335 0 >99Pro 1075 102 91Ser 240 19 92Thr 152 55 64Tyr 125 0 >99Val 343 17 95Total 6823 518 92

[a] Hydrolysis in 6 m HBr at 110 8C for 24 h, followed by evaporation, filtra-tion, dilution, and separation by HPAEC–IPAD; b-alanine was used as theinternal standard. [b] Reaction conditions: amino acids (0.5 mmol) ina mixture obtained by HBr-mediated gluten hydrolysis and neutralizationwith NH4OH; NH4Br (0.5 mmol), [Ni,Al]-LDH-WO4 (0.1 g; 18 mmol WO4

2�),H2O (4 mL), room temperature. H2O2 (5 mmol) diluted in H2O (totalvolume: 4 mL); addition rate: 0.33 mmol h�1. The reaction mixture was di-luted and analyzed by HPAEC–IPAD. [c] Arg is overestimated because ofthe co-elution of traces of HBr and therefore not shown. [d] Asx = asparticacid + asparagine. [e] Cystine = cysteine + cystine; cysteine is oxidizedduring HPAEC. [f] Glx = glutamic acid + glutamine.

Figure 5. Analysis of amino acids by high-performance anion-exchange chro-matography with integrated pulsed amperometric detection (HPAEC–IPAD):a) amino acids in gluten hydrolyzate, as prepared, b) amino acids in glutenhydrolyzate upon contact with the [Ni,Al]-LDH-WO4 catalyst before the addi-tion of H2O2. Note the clear decrease in the intensity of the peaks of theacidic fraction (glutamate 15 and aspartate 16), and c) residual amino acidsin gluten hydrolyzate upon near-complete oxidative decarboxylation[1 = Arg; 2 = Lys; 3 = b-alanine, internal standard; 4 = Ala; 5 = Thr; 6 = Gly;7 = Val; 8 = Ser; 9 = Pro; 10 = Ile ; 11 = Leu; 12 = Met; 13 = His; 14 = Phe;15 = Glu; 16 = Asp; 17 = cystine; 18 = Tyr] .

ChemSusChem 2015, 8, 345 – 352 www.chemsuschem.org � 2015 Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim349

Full Papers

Page 6: Bio-Based Nitriles from the Heterogeneously Catalyzed Oxidative … · 2017-08-17 · cilitates both bromide oxidation and amino acid oxi-dative decarboxylation. By analogy with what

practice they did not affect the oxidative decarboxylation be-cause of their low abundance, generally <5 wt %.

The identity of the products that resulted from the oxidativedecarboxylation of the wheat gluten hydrolyzate was demon-strated by various complementary analysis techniques. Typicalnitrile products were identified by 1H NMR spectroscopy of thecrude aqueous reaction mixture (Figure 6); the abundant prod-uct glutamate (3 j’) was recognized easily. The volatile com-pounds derived from valine 3 b’, leucine 3 c’, isoleucine 3 d’,and phenylalanine 2 a’ were extracted readily from the productpool using diethyl ether and analyzed by GC–MS; aldehydesand other derivatives were not detected (Figure S10). Asa result, the extracted aqueous phase was enriched in 3-cyano-propionate (3 j’; Figure S11). Although full product separationand isolation were not optimized, our results confirm the con-version of a biomass-derived pool of amino acids to a mixtureof nitrile products. The latter are at least easier to isolate andseparate than their precursors.

Remarkably, upon exposure of the [Ni,Al]-LDH-WO4 catalystto the crude protein hydrolyzate, before the onset of H2O2 ad-dition, glutamate and aspartate are adsorbed preferentially be-cause of the AEC of the catalyst (Figure 5 b). In addition, the re-activity of glutamate in oxidative decarboxylation is enhancedstrongly because of its proximity to the catalytic sites for bro-mide oxidation; if substantial amounts of neutral compoundssuch as alanine, valine, leucine, isoleucine, or phenylalanine arestill present in the mixture, glutamate has already reachednear-complete conversion (Table S7, Figure S9). Thus, the selec-tive adsorption of glutamate, coupled to its high reactivity foroxidative decarboxylation, opens possibilities for a one-pot val-orization approach specifically for this abundant nonessentialamino acid in plant biomass. The remaining unreacted mixtureis then enriched in essential amino acids and, thus, in nutri-tional quality for feed applications, for example. Finally, thesystem is amenable to upscaling, as demonstrated by severalsuccessful larger-scale reactions (Scheme 2).

Figure 6. 1H NMR spectra of the product pool obtained after the oxidative decarboxylation of a crude amino acid mixture. Characteristic signals of abundantamino acid derived nitriles are marked: acetonitrile (>alanine), isobutyronitrile (>valine), isovaleronitrile (> leucine), 2-methylbutyronitrile (> isoleucine), phe-nylacetonitrile (>phenylalanine), 3-cyanopropionate (>glutamate), and glycolonitrile (> serine). The intense signal of the remaining glycine reflects the lowreactivity of this compound. b-Alanine was used as an internal standard.

ChemSusChem 2015, 8, 345 – 352 www.chemsuschem.org � 2015 Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim350

Full Papers

Page 7: Bio-Based Nitriles from the Heterogeneously Catalyzed Oxidative … · 2017-08-17 · cilitates both bromide oxidation and amino acid oxi-dative decarboxylation. By analogy with what

Conclusions

A heterogeneous catalyst for halide oxidation, based on tung-state immobilized on a solid [Ni,Al] layered double hydroxide(LDH) support was applied to the oxidative decarboxylation ofamino acids using H2O2. Electrostatic interactions between bro-mide, carboxylates, and the LDH surface are key to achieve cat-alytic activity. High and often excellent nitrile yields were com-bined with broad functional group compatibility. Bifunctionalnitriles, such as that derived from glutamic acid, are intermedi-ates to high-value building blocks and can even be obtainedfrom cheap protein-rich side products from the agro-industry.

Experimental Section

Catalyst synthesis

LDH supports were prepared by co-precipitation from the corre-sponding metal nitrate salts in alkaline medium under ambientconditions. Ni(NO3)2·6 H2O (76.8 mol) and Al(NO3)3·9 H2O (43.2 mol)were dissolved in deionized water (240 mL). This aqueous solutionwas added dropwise by using a B jBraun Perfusor Space pump todeionized water (200 mL) at pH 8.5 under magnetic stirring. ThepH was kept constant during the synthesis by the addition of anaqueous solution of NaOH (2 m). Afterwards, the aqueous suspen-sion was stirred for 24 h at RT. The green solid was precipitated bycentrifugation, washed several times with deionized water, anddried by lyophilization. The tungstate-loaded [Ni,Al]-LDH materialwas obtained through anion exchange. Therefore, the air-dry[Ni,Al]-LDH support (3 g) was brought into contact with an aque-ous solution (300 mL) that contained Na2WO4·2 H2O (1.5 mmol) andstirred for 24 h at RT. The washing and drying protocol was repeat-ed.

Catalyst characterization

PXRD measurements were performed by using a STOE StadiP dif-fractometer with CuKa radiation. SEM was performed by usinga Philips XL 30 FEG microscope after the samples were coated withAu. Nitrogen physisorption measurements were performed byusing a Micromeritics 3Flex surface analyzer at 77 K. Before the

measurement, the 100 mg samples were outgassed at 423 K for6 h under vacuum. XPS measurements were performed by usingan S-probe monochromatized XPS spectrometer (Surface ScienceInstruments VG) equipped with an AlKa X-ray (1486.6 eV) mono-chromatic source. The take-off angle was 458, and the voltage andpower of the source were 10 kV and 200 W, respectively. A basepressure of 3 � 10�9 mbar was obtained in the measuring chamber,and the pass energy was 107 eV. The analysis surface was 250 �1000 mm2 with a flood gun (neutralizer) setting of 4 eV on thesample with Ni grid. The accumulation time was approximately 7 hfor each spectrum. Elemental analysis was performed by usinga Jobin Yvon Ultima spectrometer. Therefore, the LDH material(50 mg) was degraded by heating at 110 8C in aqua regia (HNO3/HCl = 1:3 v/v ; 0.5 mL) and HF (3 mL) for 1 h. After cooling to RT, de-ionized water (10 mL) and H3BO3 (2.8 g) were added, and the mix-ture was further diluted to a final volume of 100 mL. The C and Ncontents were determined according to the Dumas method byusing a Vario Max CN Analyzer.

Typical reaction procedure

Catalytic reactions were performed in glass batch reactors (11 mL)and were stirred magnetically at RT. Unless stated otherwise,[Ni0.64Al0.36(OH)2](NO3)0.04(CO3)0.14(WO4)0.02 with a tungstate loading of0.18 mmol g�1 was used as the catalyst. In a standard reaction,amino acid (0.5 mmol), bromide salt (0.5 mmol), catalyst (0.1 g),and solvent (4 mL) were added to the reactor. H2O2 (5.0 mmol;35 wt % in water) was diluted in the solvent (total volume: 4 mL)and added continuously by using a B jBraun Perfusor Space pumpat a rate of 0.33 mol h�1. After this addition step, the heterogene-ous catalyst was removed by centrifugation. Either GC or 1H NMRspectroscopy was applied for analysis, which depended on the ex-pected products.

Product analysis and identification

Reaction mixtures were analyzed by 1H NMR spectroscopy to deter-mine the conversion and selectivity in the oxidative decarboxyla-tion reaction. 1H NMR spectra were recorded by using a BrukerAvance 400 MHz spectrometer equipped with a BBI 5 mm probe.As the reaction was performed in water for many of the aminoacids except for 2 a, the NMR sample was prepared by diluting theproduct mixture (300 mL) in D2O (300 mL) in an NMR tube. Other-wise the reaction was performed in CD3CN or CD3OD, and the reac-tion mixture was analyzed without further modification. The broadsignal caused by the presence of water, as a result of using aque-ous H2O2 as the terminal oxidant for oxidative decarboxylation,was suppressed by the application of an adapted pulse program:p1 8 ms; pl1 �1 db; pl9 50 db; o1P on the resonance signal ofwater, determined from the previous 1H NMR spectroscopy mea-surement: ds 2; ns 32; d1 5 s; aq 2.55 s; sw 16. In the case of GC,the samples were analyzed by using a Shimadzu 2010 or 2014 GCequipped with a CP-Sil 8 CB and CP-Chirasil-Dex CB capillarycolumn, respectively, and a flame ionization detector (FID). Thestandard temperature program allowed the baseline separation ofall compounds; the temperature is increased from 50 to 300 8C at10 8C min�1. To verify the absence of any racemization in the prod-uct of isoleucine, chromatographic analysis was performed byusing a CP-Chirasil-Dex CB capillary column. A slightly adaptedtemperature program was applied, with a start at 50 8C for 20 minand heating to 200 8C at 10 8C min�1. For quantitative analysis, ben-zonitrile (0.5 mmol, 2 m in methanol) was added as a standardwhen H2O2 addition was finished; the GC peak areas were correct-

Scheme 2. Oxidative decarboxylation of 2 a and 3 c on a larger scale (S : se-lectivity). Reaction conditions: amino acid (10 mmol), NH4Br (2.5 mmol),[Ni,Al]-LDH-WO4 (0.5 g), solvent (20 mL), room temperature. H2O2 (100 mmol)diluted in solvent (total volume: 20 mL), addition rate: 6.75 mmol h�1.

ChemSusChem 2015, 8, 345 – 352 www.chemsuschem.org � 2015 Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim351

Full Papers

Page 8: Bio-Based Nitriles from the Heterogeneously Catalyzed Oxidative … · 2017-08-17 · cilitates both bromide oxidation and amino acid oxi-dative decarboxylation. By analogy with what

ed and validated by determining the appropriate response factors.Products were identified by GC–MS by using an Agilent 6890 GC,equipped with a HP-5 ms column, coupled to a 5973 MSD massspectrometer. 1H and 13C NMR and MS data are given in the Sup-porting Information.

Protein hydrolysis and amino acid analysis

A wheat gluten sample (1.50 g) that contained 1.16 g dry matterprotein was suspended in b-alanine solution (49.5 mL, 9.0 mm, asinternal standard) and HBr solution (100.5 mL, 48 wt %) to resultingin a 6 m HBr solution. b-Alanine was used as the internal standardbecause it remains intact during the oxidative decarboxylation. Thesample was incubated under N2 to prevent amino acid oxidationby flushing the headspace for 60 s with N2 before hydrolysis(110 8C, 24 h). Afterwards, the reaction mixture was evaporated todryness at 110 8C, suspended in water (7.5 mL), and filtered. Afterdilution (4000-fold) and filtration (Millex-GP, 0.22 mm, polyethersul-fone, Millipore), the amino acid levels were determined by high-performance anion-exchange chromatography with integratedpulsed amperometric detection (HPAEC–IPAD).[28]

Acknowledgements

We are grateful to IWT Flanders (L.C.), FWO Flanders (I.R. , I.S. ,and T.D.B.), KU Leuven (R.M., project IKP/10/005 ; B.L. , IOF-ZKC6712), BELSPO (IAP-PAI P7/05), and the Flemish government(Methusalem grant CASAS). We thank Karel Duerinckx and NicoDe Roo for assistance with NMR and XPS measurements.

Keywords: amino acids · biomass · heterogeneous catalysis ·oxidation · tungsten

[1] a) A. Corma, S. Iborra, A. Velty, Chem. Rev. 2007, 107, 2411 – 2502;b) R. A. Sheldon, Green Chem. 2014, 16, 950 – 963; c) C. O. Tuck, E. P�rez,I. T. Horv�th, R. A. Sheldon, M. Poliakoff, Science 2012, 337, 695 – 699.

[2] a) M. Breuer, K. Ditrich, T. Habicher, B. Hauer, M. Keßeler, R. St�rmer, T.Zelinski, Angew. Chem. Int. Ed. 2004, 43, 788 – 824; Angew. Chem. 2004,116, 806 – 843; b) A. L. Demain, Ind. Biotechnol. 2007, 3, 269 – 283.

[3] T. M. Lammens, M. C. R. Franssen, E. L. Scott, J. P. M. Sanders, Biomass Bi-oenergy 2012, 44, 168 – 181.

[4] A. Van Der Borght, H. Goesaert, W. S. Veraverbeke, J. A. Delcour, J. CerealSci. 2005, 41, 221 – 237.

[5] C. Wood, K. A. Rosentrater, K. Muthukumarappan, Ind. Crops Prod. 2014,56, 118 – 127.

[6] P. G. Dalev, Bioresour. Technol. 1994, 48, 265 – 267.[7] C. Mart�n, A. Moure, G. Mart�n, E. Carrillo, H. Dom�nguez, J. C. Paraj�,

Biomass Bioenergy 2010, 34, 533 – 538.

[8] E. Scott, F. Peter, J. Sanders, Appl. Microbiol. Biotechnol. 2007, 75, 751 –762.

[9] G. W. Huber, S. Iborra, A. Corma, Chem. Rev. 2006, 106, 4044 – 4098.[10] a) Y. Teng, E. L. Scott, A. N. T. van Zeeland, J. P. M. Sanders, Green Chem.

2011, 13, 624 – 630; b) O. M. Kattan Readi, E. Rolevink, K. Nijmeijer, J.Chem. Technol. Biotechnol. 2014, 89, 425 – 435.

[11] a) G. W. Stevenson, J. M. Luck, J. Biol. Chem. 1961, 236, 715 – 717; b) G.Laval, B. T. Golding, Synlett 2003, 542 – 546.

[12] a) H. D. Dakin, Biochem. J. 1916, 10, 319 – 323; b) H. D. Dakin, Biochem. J.1917, 11, 79 – 95.

[13] a) G. A. Hiegel, J. C. Lewis, J. W. Bae, Synth. Commun. 2004, 34, 3449 –3453; b) L. De Luca, G. Giacomelli, Synlett 2004, 2180 – 2184.

[14] J. Le Ntre, E. L. Scott, M. C. R. Franssen, J. P. M. Sanders, Green Chem.2011, 13, 807 – 809.

[15] a) T. Shono, Y. Matsumara, K. Inoue, J. Am. Chem. Soc. 1984, 106, 6075 –6076; b) J. J. Dai, Y. B. Huang, C. Fang, Q. X. Guo, Y. Fu, ChemSusChem2012, 5, 617 – 620; c) R. Matthessen, L. Claes, J. Fransaer, K. Binnemans,D. E. De Vos, Eur. J. Org. Chem. 2014, 6649 – 6652.

[16] a) M. Nieder, L. Hager, Arch. Biochem. Biophys. 1985, 240, 121 – 127; b) A.But, J. Le Ntre, E. L. Scott, R. Wever, J. P. M. Sanders, ChemSusChem2012, 5, 1199 – 1202.

[17] a) B. F. Sels, D. E. De Vos, M. Buntinx, P. A. Jacobs, J. Catal. 2003, 216,288 – 297; b) B. Sels, D. De Vos, M. Buntinx, F. Pierard, A. Kirsch-De Mes-maeker, P. Jacobs, Nature 1999, 400, 855 – 857; c) B. F. Sels, D. E. De Vos,P. A. Jacobs, J. Am. Chem. Soc. 2001, 123, 8350 – 8359.

[18] a) P. Levecque, H. Poelman, P. Jacobs, D. De Vos, B. Sels, Phys. Chem.Chem. Phys. 2009, 11, 2964 – 2975; b) M. Jitianu, D. C. Gunness, D. E.Aboagye, M. Zaharescu, A. Jitianu, Mater. Res. Bull. 2013, 48, 1864 –1873.

[19] S. Aisawa, S. Takahashi, W. Ogasawara, Y. Umetsu, E. Narita, J. Solid StateChem. 2001, 162, 52 – 62.

[20] A. H. Friedman, S. Morgulis, J. Am. Chem. Soc. 1936, 58, 909 – 913.[21] W. E. Pereira, Y. Hoyano, R. E. Summons, V. A. Bacon, A. M. Duffield, Bio-

chim. Biophys. Acta Gen. Subj. 1973, 313, 170 – 180.[22] D. E. De Vos, B. F. Sels, P. A. Jacobs, Cattech 2002, 6, 14 – 29.[23] T. M. Lammens, J. Le Ntre, M. C. R. Franssen, E. L. Scott, J. P. M. Sanders,

ChemSusChem 2011, 4, 785 – 791.[24] B. W. Hoffer, J. A. Moulijn, Appl. Catal. A 2009, 352, 193 – 201.[25] B. M. Choudary, B. Bharathi, C. Venkat Reddy, M. Lakshmi Kantam, J.

Chem. Soc. Perkin Trans. 1 2002, 2069 – 2074.[26] a) Y. Yokoyama, T. Yamaguchi, M. Sato, E. Kobayashi, Y. Murakami, H.

Okuno, Chem. Pharm. Bull. 2006, 54, 1715 – 1719; b) A. Podgorsek, M.Zupan, J. Iskra, Angew. Chem. Int. Ed. 2009, 48, 8424 – 8450; Angew.Chem. 2009, 121, 8576 – 8603.

[27] a) A. Butler, J. V. Walker, Chem. Rev. 1993, 93, 1937 – 1944; b) F. H. Vaillan-court, E. Yeh, D. A. Vosburg, S. Garneau-Tsodikova, C. T. Walsh, Chem.Rev. 2006, 106, 3364 – 3378.

[28] I. Rombouts, L. Lamberts, I. Celus, B. Lagrain, K. Brijs, J. A. Delcour, J.Chromatogr. A 2009, 1216, 5557 – 5562.

Received: August 6, 2014

Revised: October 21, 2014

Published online on December 3, 2014

ChemSusChem 2015, 8, 345 – 352 www.chemsuschem.org � 2015 Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim352

Full Papers