You are on page 1of 12

J Am Oil Chem Soc (2021) 98: 305–316

DOI 10.1002/aocs.12468

ORIGINAL ARTICLE

Production of Industrially Useful and Renewable p-Cymene by


Catalytic Dehydration and Isomerization of Perillyl Alcohol
Bryan R. Moser1 · Michael A. Jackson2 · Kenneth M. Doll1

Received: 6 November 2020 / Revised: 3 December 2020 / Accepted: 28 December 2020


Published 2021. This article is a U.S. Government work and is in the public domain in the USA.

Abstract We report the dehydration and isomerization of (5%), which was attributed to the strong Brønsted acidity of
renewable perillyl alcohol to industrially useful p-cymene pTsOH. In addition, camphene (44%), terpinene isomers
in 91% yield utilizing 2.0 mol% para-toluenesulfonic acid (15%), and limonene (14%) were obtained when
(pTsOH) catalyst at 110  C as a 3.0 M solution in toluene. dehydroisomerization was attempted on 3.0 M solutions of
Lower reaction temperatures, catalyst loadings, and/or α- and β-pinene, which was once again attributed to the
starting concentrations resulted in lower yields of p-cymene acidity of the catalyst. Oligomerization was strongly
as well as longer reaction times. Conversion of perillyl favored when dehydroisomerization of dl-limonene, α- and
alcohol to p-cymene yielded atom and carbon economies of β-pinene was attempted neat. In summary, synthesis of
88.1% and 100% as well as an E-factor of 2.7, thereby indi- renewable p-cymene was readily achieved from perillyl
cating that the process was both green and sustainable. A alcohol with catalytic pTsOH but competing side reactions
lower yield of 86% was observed when the reaction was suppressed yield when dehydroisomerization of dl-limo-
performed neat, but a lower E-factor of 0.4 indicated that nene, α- and β-pinene was attempted due to the strong
neat conditions were more desirable from an environmental Brønsted acidity of the catalyst.
perspective. Application of the optimized parameters to
3.0 M solutions of dl-limonene led predominantly to oligo- Keywords p-Cymene  Isomerization  Limonene 
merization (92%) as opposed to dehydroisomerization Perillyl alcohol  Pinene  Renewable  Terpene

J Am Oil Chem Soc (2021) 98: 305–316.


Supporting information Additional supporting information may be
found online in the Supporting Information section at the end of the
article.
Introduction
* Bryan R. Moser
bryan.moser@usda.gov p-Cymene [1; 4-isopropyltoluene; 1-methyl-4-(propan-2-yl)
1
benzene] is a naturally occurring aromatic monoterpenoid that
United States Department of Agriculture, Agricultural Research
Service, National Center for Agricultural Utilization Research, consists of a benzene ring para-substituted with methyl and
Bio-Oils Research Unit, 1815 N. University Street, Peoria, IL iso-propyl groups. Noteworthy natural sources of 1 are essen-
61604, USA tial oils from cumin (Cuminum cyminum L.) and thyme (Thy-
2
United States Department of Agriculture, Agricultural Research mus vulgaris L.) (Asbaghian et al., 2011; Wanner et al., 2010).
Service, National Center for Agricultural Utilization Research, Industrial applications include as a precursor for terephthalic
Renewable Products Technology Research Unit, 1815
N. University Street, Peoria, IL 61604, USA acid, p-cresol, p-menthane, pesticides, fungicides, herbicides,
and pharmaceuticals. Other uses are as a green solvent, heat
Mention of trade names or commercial products in this publication is transfer medium, ligand for catalysts, and as a constituent in
solely for the purpose of providing specific information and does not
imply recommendation or endorsement by the U.S. Department of
fragrances and perfumes (Fig. 1) (Arockiam et al., 2012; Du
Agriculture. USDA is an equal opportunity provider and employer. et al., 2005; Gandeepan et al., 2019; Volanti et al., 2019).

Published online: 19 February 2021


J Am Oil Chem Soc (2021) 98: 305–316
306 J Am Oil Chem Soc

Fig 1 Uses of p-cymene depicting production of (a) cresol and acetone, (b) p-terephthalic acid, and (c) p-menthane as well as direct uses as a
green solvent, heat transfer medium, ligand for well-defined catalysts, and as a component in fragrances and aroma therapies

Due to the industrial utility of 1 combined with its lim- preferred (Perego et al., 1999). As a result, industrial mixtures
ited biological availability from high-value culinary typically contain 65% m-cymene, 30% 1, and 5% of the steri-
sources, petrochemical routes have been developed to aug- cally hindered ortho isomer (Brzozowski et al., 2002).
ment supply. The most prominent is the Friedel-Crafts Adsorptive separation of the mixture utilizing the Cymex
alkylation of toluene with propene or 2-propanol, which is process is thus necessary to isolate 1 on an industrial scale,
historically catalyzed by highly toxic and corrosive Lewis with positional isomer ratio greatly affecting separation cost
and Brønsted acids such as AlCl3, BF3, FeCl3, FeSO4-HCl, and yield (Johnson and Kabza, 1990).
HF, and H2SO4 (Fig. 2) (Barman et al., 2005). However, Solid acid catalysts such as zeolites, mesoporous molec-
under these conditions the meta isomer is thermodynamically ular sieves, and mixed metal oxides are safer, less corro-
sive, and more selective alternatives to the homogenous
Friedel-Crafts approach (Odedairo and Al-Khattaf, 2011;
Perego and Ingallina, 2002; Zilkova et al., 2016). However,
byproducts such as n-propyltoluene (Fig. 2), arising from
transalkylation of 1 with toluene and m- and o-cymene, are
often produced. n-Propyltoluene is not produced during tra-
ditional homogenous Friedel-Crafts alkylation but is prob-
lematic with H-ZSM-5, H Y zeolite, and H-mordenite due
to the effects of zeolitic channel geometry and pore size on
reactivity and selectivity (Perego and Ingallina, 2002;
Wichterlova and Cejka, 1994; Zilkova et al., 2016).
The petrochemical approach suffers from several signifi-
cant drawbacks, such as the utilization of toxic starting
materials, production of less useful ortho- and meta- iso-
mers that require expensive separation, generation of
n-propyltoluene, energy-intensive reaction temperatures in
excess of 300  C, employment of hazardous and corrosive
Fig 2 Petrochemical production of p-cymene from toluene via catalysts, and the inherently poor long-term sustainability
Friedel-Crafts alkylation with propene or 2-propanol. Yield-reducing
transalkylation with toluene to give n-propyltoluene (para isomer and negative environmental impact of a non-renewable
depicted) is also shown approach. Consequently, bio-based routes have been

J Am Oil Chem Soc (2021) 98: 305–316


J Am Oil Chem Soc 307

devised that mitigate one or more of these disadvantages. Transformation of monoterpenes to 1 is typically accom-
Such methods primarily focus on the dehydroisomerization plished utilizing solid acid catalysts and is conducted in the gas
of monoterpenes (Fig. 3) such as limonene (2), α-pinene phase at high temperatures (≥300  C) with a continuous flow
(α-3), and β-pinene (β-3), which represent improvements of H2 or N2 gas. Examples are summarized in Table 1 with
over the petrochemical approach regarding renewability, the reaction sequence depicted in Fig. 3. In most cases, the
sustainability, and environmental impact. catalyst requires energy-intensive calcination at >500  C for
Limonene and pinene are architecturally similar to 1 and several hours before sufficient activity is realized. In some
enjoy wide natural availability. Found in citrus rind, (+)-2 (d- examples, clay catalysts require activation with mineral acids
limonene) is the primary constituent of essential oils from such as HCl or H2SO4. In addition, catalysts containing
orange, lemon, mandarin, grapefruit, and lime. The other iso- expensive and/or toxic metals such as palladium or chromium
mer, (−)-2 (l-limonene), is less common, has a turpentine-like are often employed along with pressures exceeding ambient.
odor, and is a component of volatiles emitted by oaks and Such metals are also known to catalyze unproductive side
pines (Ciriminna et al., 2014). A racemic mixture of (+)- and reactions such as disproportionation that yield complex mix-
(−)-2 is referred to as dipentene or dl-limonene and is a con- tures of 1, p-menthane and other compounds (Lesage
stituent of crude sulfate turpentine that is produced during et al., 1996; Tavera Ruiz et al., 2019; Thomas and
processing of wood into paper products. The most naturally Bessiere, 1989). Thus, a new bio-based approach is of interest
abundant monoterpenes are α-3 and β-3 (Fig. 3), which are that avoids expensive metals, corrosive mineral acids,
the primary components of turpentine oil and pine and coni- unproductive side reactions, and elevated temperatures and
fer resins (Winnacker, 2018). pressures while simultaneously achieving a high yield of 1 in
an environmentally sustainable fashion.
The objective of the current investigation was the cata-
lytic production of 1 from perillyl alcohol [4;
(4-isopropenyl-1-cyclohexen-1-yl)methanol] utilizing mild
reaction conditions. Principally found in the essential oils
of lavender, sage, and peppermint, 4 is a monoterpene
metabolite of 2 (Anis et al., 2018; Erland et al., 2016). Con-
version of 2 to 4 is accomplished in vivo by hydroxylation
enzymes that belong to the superfamily of cytochrome
P450 proteins. Biocatalytic in vitro production of 4 from
2 is achieved using cytochrome P450 hydroxylases
expressed in Pseudomonas putida (van Beilen et al., 2005).
The chemical synthesis of 4 from 2 in four steps has also
been reported (Geoghegan and Evans, 2014). Although 4 is
currently less abundant and more expensive than both
1 and 2, we became interested in 4 as a precursor to 1 due
to the serendipitous discovery that 1 can be produced from
4 in the presence of a homogenous Brønsted acid catalyst
while attempting an unrelated transformation. We thus
report herein the optimization of 1 from 4 and attempt to
apply the optimized methodology to 2, α-3, and β-3 to
enhance the supply of renewable and industrially useful 1.

Experimental

Materials

Camphene (95%), 3-carene (90%), o-cymene (98%),


m-cymene (99%), p-cymene (99%), eucalyptol (99%), dl-
limonene (1:1 mixture of (+)/(−)-limonene; 100%),
Fig 3 Chemical pathway to p-cymene (1) from limonene (2), α-pinene
[α-3; (1R)-(+)-α-pinene depicted], and β-pinene [β-3; (1R)-(+)–βpinene mesitylene (98%), S-(−)-perillyl alcohol (96%),
depicted] showing key carbocationic and p-menthadienoic intermediates α-phellandrene (>95%), (+)-α-pinene (98%), (−)-β-pinene

J Am Oil Chem Soc (2021) 98: 305–316


308 J Am Oil Chem Soc

Table 1 Literature survey of catalyst, reaction conditionsa, and p-cymene yield from various terpenes under heterogeneous conditions

Terpene Catalyst Calcb Temp Time Pressure p- Ref


( C; hour) ( C) Cymene (%)

α-Limonene (Ce/Pd)ZSM-5 550; 6 250 contc1 1 bar 80.8 Weyrich and


Holderich, 1997
Dipentene Pd/SiO2 550; 8 300 contc1 1 bar 90.0 Buhl et al., 1998
α/β-Pinene Pd/D-11-10 540; 6 300 contc2 1.01 bar 67.0 Roberge et al., 2001
Dipentene Zn/SBA-15 550; 5 450 contc2 1 atm 85.7 Du et al., 2005
Limonene HCl/bentonite 150; 16 80 15 min 1 atm 15.0 Fernandes et al., 2007
α-Pinene ZnO/Cr2O3 300; 5 350 contc1 1 atm 78.0 Al-Wadaani et al., 2009
α-Limonene Ni, Fe, Mn sepiolite 400; 6 165 20 min 1 atm 100 Martin-Luengo et al., 2010
α-Limonene TiO2 – Anatase 400; 6 300 contc2 1 atm 90 Kamitsou et al., 2014
α-Limonene mesoporous Si Al 300; 5 225 20 min 1 atm 100 Martin-Luengo et al., 2008
oxides
α-Limonene H2SO4/mordenite – 140 7 hours 1 atm 63 Makarouni et al., 2018
α-Limonene Pd PVP nanoparticles – >150 3 hours 2 bar H2 76 Zhao et al., 2008
α-Pinene Pd-Zn/Al-SBA-15 550; 6 300 contc2 1 atm 77 Golets et al., 2013
dl- Pd/Al2O3 – 300 30 s 8.0 MPa 80 Yilmazoglu and
Limonene Akgun, 2018
α-Pinene Au-Ni-TiO2/SBA-15 550; 5 300 3 hours 1 atm 55.9 Ajaikumar et al., 2013
dl- HCl/mordenite – 140 7 hours 1 atm 65 Lycourghiotis et al., 2018
Limonene
a
Catalyst and reaction conditions that provided the highest yield of p-cymene; 1 atm refers to atmospheric pressure.
b
Calcination temperature ( C) and time (hour).
c
cont, continuous process under a constant flow of (c1) N2 or (c2) H2 gas.

(99%), α-terpinene (85%), γ-terpinene (97%), δ-terpinene for γ-terpinene (0.780), 11.56 for n-propyltoluene, 12.15 for
(≥90%), and para-toluenesulfonic acid (pTsOH; 98%) were δ-terpinene (0.781), and 18.17 for 4 (0.719).
purchased from Millipore Sigma Corp. (St. Louis, MO,
USA) and n-propyltoluene (≥99%) was obtained from NMR and FTIR Spectroscopy
VWR International, Inc. (Radnor, PA, USA). All other
materials were purchased from Millipore Sigma Corp. and Nuclear magnetic resonance (NMR) spectra were collected at
used as received. 26.9  C on a Bruker Avance-500 spectrometer (Billerica, MA,
USA) operating at 500 (1H) and 125 (13C) MHz using a 5 mm
Gas Chromatography–Mass Spectroscopy broadband observe probe. Samples were dissolved in CDCl3
(Cambridge Isotope Laboratories, Andover, MA, USA) and
Reaction progress was monitored with an Agilent (Santa Clara, chemical shifts (δ) were reported as ppm from tetramethylsilane
CA, USA) 7890A GC equipped with a 5975C quadrupole and multiplicity indicated by s (singlet), bs (broad singlet), d
electron impact (EI) mass detector and a Phenomenex Zebron (doublet), or m (multiplet). FTIR (Fourier transform infrared)
ZB-35HT column (30 m × 0.32 mm × 0.25 μm) with He as spectra were obtained on a Thermo-Nicolet Nexus 470 FTIR
carrier gas at 1.4 mL min−1 with inlet and detector temperatures spectrometer (Madison, WI, USA) with a scanning range of
of 340 and 230  C. The oven held at 40  C for 3 min, ramped 600–4000 cm−1 for 64 scans at a spectral resolution of 4 cm−1.
to 90  C at 5  C min−1, ramped to 190  C at 10  C min−1, held
for 5 min, and ramped to 340  C at 25  C min−1. Aliquots were Synthesis of 1
diluted in heptane to give signals that were identified by com-
parison to reference standards and by mass spectral analysis. To a flask containing 4 (1.0 g, 6.57 mmol) and a magnetic stir
Retention times (min) of reference standards and response fac- bar was added catalytic pTsOH (1.0 or 2.0 mol%) and the
tors relative to p-cymene (indicated in parenthesis) were 6.50 mixture stirred (600 rpm) at 100  C or 110  C for 3–24 hours.
for α-3 (0.782), 7.20 for camphene (0.771), 8.22 for β-3 Reactions were performed solventless (neat) or as 1.0, 2.0, or
(0.789), 8.98 for mesitylene (0.953), 9.30 for α-phellandrene, 3.0 M solutions of 4 in toluene with mesitylene (10 mol%) as
9.65 for α-terpinene (0.782), 10.07 for (+/−)-2 (0.726), 10.48 an internal standard. Aliquots were periodically removed for
for m-cymene, 10.56 for 1 (1.00), 11.11 for o-cymene, 11.25 GC–MS determination of reaction progress. Yield data for

J Am Oil Chem Soc (2021) 98: 305–316


J Am Oil Chem Soc 309

most reactions were obtained using GC–MS taking into con- pTsOH in refluxing toluene to give the corresponding tris-
sideration the internal mesitylene standard and relative unsaturated perillyl decenoate as a monomer for subsequent
response factors. In some cases, isolation of 1 and related acyclic diene metathesis polymerization and thiol-ene pho-
byproducts was achieved by equilibrating the reaction mixture tochemical polycondensation. The esterification procedure
to room temperature, filtering through a short (< 1 in.) pad of had proven simple and effective for several previous con-
silica gel (70–230 mesh, Sigma-Aldrich Corp.) to remove cat- densations involving different alcohol/acid combinations
alyst, and solvent was recovered (if applicable) via reduced (Moser, 2014; Moser et al., 2019, 2020), but in this case
pressure rotary evaporation (50 mbar, 35  C) to give a color- the expected ester was not observed. Instead, a mixture
less liquid: 1H NMR (CDCl3) δ 7.16 (4H, bs, 4 × Ar-H), 2.92 comprised of 4, unreacted 9-decenoic acid, and a curiously
(1H, septet, J = 6.9 Hz, CH3CH(Ar)CH3), 2.36 (3H, s, Ar- high amount of 1 (31%) was noted. With this serendipitous
CH3), 1.28 (6H, d, J = 6.9 Hz, CH3CH(Ar)CH3); 13C NMR discovery in hand, we set out to improve the yield of indus-
(CDCl3) δ 145.9, 135.2, 129.1, 126.3, 33.7, 24.1, 21.0; IR trially useful 1 from 4 utilizing catalytic pTsOH. We chose
(neat) νmax 2958, 2912, 2869, 1515, 1462, 1382, 1362, 1302, pTsOH because it is simple, inexpensive, less corrosive,
1109, 1056, 1020, 813, 720, 540 cm−1; MS (EI) m/z calcd for and less toxic than mineral acids such as sulfuric acid and
C10H14 134.1, found 134.1 [M]+, 119.1 [M – CH3]+, 91.1 [M is less likely to sulfonate our alkenoic substrates and
– Pr]+, 77.1 [M – CH3 – Pr]+. 3-Isopropyl-6-methylcyclohex- products.
2-en-1-one (5) was periodically identified as a minor compo-
nent (3–5%) via GC–MS and confirmed by NMR spectros- Synthesis of p-Cymene from Perillyl Alcohol at 100  C
copy: 1H NMR (CDCl3) δ 5.90 (1H, s, >C CH ),
2.46–2.42 (1H, m, CH(CH3)2), 2.39–2.37 (2H, m, Initial efforts directed at optimization of 1 from 4 were con-
CH2 ), 2.35–2.34 (1H, m, CH2CH(CH3)C( O)CH ), ducted neat at 100  C with catalyst loadings of 1.0 and
2.13–2.09 (1H, m, >CHCH2CH2 ), 1.69–1.67 (1H, m, 2.0 mol% pTsOH. Consumption of 4 was complete by
>CHCH2CH2 ), 1.19 (3H, d, J = 6.9 Hz, CH2CH(CH3)C 2 hours under these conditions. At a 2.0 mol% catalyst,
( O)CH ), 1.15 (3H, d, J = 6.9 Hz, CH(CH3)2), 1.14 (3H, maximum yield of 1 was achieved (86%) within 2 hours,
d, J = 6.9 Hz, CH(CH3)2); 13C NMR (CDCl3) δ 202.8, whereas 6 hours was required to give a lower yield of 84%
170.7, 123.1, 41.1, 35.6, 31.1, 27.3, 20.6, 19.9, 15.2; MS when 1.0 mol% catalyst was utilized (Fig. 4d). Most of the
(EI) m/z calcd for C10H16O 152.1, found 152.1 [M]+, 137.1 remaining material was non-aromatic dimeric species (16%
[M – CH3]+, 110.1 [M – Pr]+, 95.1 [M – CH3 – Pr]+. and 13% for 1.0 and 2.0 mol% pTsOH, respectively). It
was speculated that dilution of 4 in an inert solvent would
favor intramolecular conversion to 1 while suppressing
Green Metrics
intermolecular dimerization. Therefore, a 1.0 M solution of
4 (14.9 wt.%) in toluene was investigated. Toluene was
Renewable carbon content (RCC; %) was calculated as the
selected due to its relatively high boiling point and chemi-
ratio of the number of renewable carbons to total carbons
cally inert nature under these conditions.
in the product. Atom economy (AE; %) was determined by
As seen in Fig. 4a, consumption of 4 and accumulation
dividing the MW of product by the MW of all stoichiomet-
of 1 were slower at 1.0 M relative to neat conditions, espe-
ric products (Trost, 1991). Carbon economy (CE; %) was
cially with 1.0 mol% catalyst. The yield of 1 was only 22%
calculated as the ratio of the mass of carbon in the product
after 24 hours at 1.0 mol% pTsOH with 12% unreacted
by the total mass of carbon in the stoichiometric reactants
4 remaining as well as a considerable amount (53%) of
(Curzons et al., 2001). The environmental factor (E-factor)
dimeric species. Maximum yield (85%) was reached faster
was determined as the ratio of the mass of non-product to
(16 hours) when the catalyst load was doubled to 2.0 mol
the mass of product (Sheldon, 2007).
%. As a result, the starting concentration of 4 was increased
from 1.0 to 2.0 M (29.8 wt.%) to accelerate conversion
while attempting to mitigate dimerization.
Results and Discussion Consumption of 4 and accumulation of 1 at 1.0
(24 hours) and 2.0 mol% pTsOH (6 hours) were faster at
Serendipitous Discovery of p-Cymene from Perillyl 2.0 M than 1.0 M, as indicated in Fig. 4b. Furthermore,
Alcohol dimeric species were limited to less than 8% at maximum
yield of 1 at both catalyst loadings, which was an improve-
As part of our continuing efforts to produce aliphatic poly- ment over neat and 1.0 M conditions. Lastly, 3.0 M con-
esters from renewable starting materials (Moser centrations of 4 (44.7 wt.%) in toluene were studied to
et al., 2019, 2020), we were originally attempting to ester- further reduce the time required to reach maximum yield.
ify 9-decenoic acid with 4 in the presence of catalytic As shown in Fig. 4c, a maximum yield of 89% was attained

J Am Oil Chem Soc (2021) 98: 305–316


310 J Am Oil Chem Soc

Fig 4 Conversion of perillyl alcohol (PA) to p-cymene (CY) at 100  C catalyzed by 1.0 and 2.0 mol% para-toluenesulfonic acid in the presence
of (a) 1.0 M toluene, (b) 2.0 M toluene, (c) 3.0 M toluene, and d) no solvent (neat). Molarities were calculated with respect to PA

after 24 and 4 hours at 1.0 and 2.0 mol% catalyst, respec- concentration of 4. Under these conditions, 4 was consumed
tively. Consumption of 4 was complete after 6 and 2 hours after 2 hours, the maximum yield of 1 was achieved in
at 1.0 and 2.0 mol% pTsOH, respectively, with dimer pro- 4 hours, and accumulation of dimers was minimized. Investi-
duction limited to less than 7% at maximum yield of 1. The gation of a higher temperature (110  C) was of interest to
early accumulation of dimers followed by their sharp both shorten the reaction time and further improve the
decline (see Fig. 4b, c and d), except for 1.0 mol% pTsOH yield of 1.
at 1.0 M (Fig. 4a), led us to believe that they served as
intermediates during conversion of 4 to 1. The incongruous Synthesis of p-Cymene from Perillyl Alcohol at 110  C
results obtained at 1.0 M with 1.0 mol% pTsOH were
attributed to a reduced reaction rate caused by lower reac- Based on the promising results obtained at 100  C, reac-
tant and catalyst concentrations relative to other conditions. tions at 110  C were conducted for 3 hours with a starting
Dimer production is discussed in more detail in the “Pro- concentration of 3.0 M utilizing 1.0 and 2.0 mol% pTsOH
posed Mechanism”. (Fig. 5). The yield of 1 approached 80% with 1.0 mol%
In summary, the optimum reaction parameters at 100  C pTsOH at 3 hours (Fig. 5a), which was a significant
were a catalyst loading of 2.0 mol% and a 3.0 M improvement over otherwise similar conditions at 100  C

J Am Oil Chem Soc (2021) 98: 305–316


J Am Oil Chem Soc 311

that required 16 hours to achieve comparable results (see mostly of oligomeric species (92%). Oligomers continued
Fig. 4c). The yield of 1 increased to 91% at 3 hours when to accumulate at the expense of 1 as the reaction prog-
the catalyst load was doubled to 2.0 mol% (Fig. 5b). Accu- ressed, reaching a maximum of 99% at 3 hours.
mulation of 1 increased sharply for the first 1.5 hours at cat- Dehydroisomerization of (+/−)-2 was then attempted at
alyst loadings of 1.0 (Fig. 5a) and 2.0 (Fig. 5b) mol % but 80  C in 3.0 M cyclohexane utilizing 2.0 mol% pTsOH in
slowed thereafter. The slower accumulation of 1 after an effort to suppress oligomerization. However, these con-
1.5 hours was accompanied by a slow decrease in dimeric ditions resulted in accumulation of even more oligomers
species. The final percentage of dimers at 3 hours with (99%). Previous studies that utilized heterogeneous cata-
2.0 mol% pTsOH was 4%, whereas 18% dimers remained lysts such as zeolites and acid-activated clays also encoun-
when 1.0 mol% catalyst was utilized. Lastly, a small tered significant oligomerization, which was attributed to
amount of ketone 5 gradually accumulated to 3% with both regions of strong Brønsted acidity on the catalytic surface
1.0 and 2.0 mol% pTsOH, which was not observed when (Du et al., 2005; Fernandes et al., 2007; Golets et al., 2015;
reactions were conducted at 100  C. The nature of 5 is dis- Linnekoski et al., 2014; Lycourghiotis et al., 2018;
cussed more extensively in the “Proposed Mechanism”. In Makarouni et al., 2018). Thus, we speculated that the
summary, the highest yield of 1 (91%) was achieved with a strong Brønsted acidity of pTsOH (pKa −1.3; Berkowitz
starting concentration of 4 at 3.0 M in toluene in 3 hours at and Grunwald, 1961) favored oligomerization over
110  C with 2.0 mol% pTsOH. Lower catalyst loadings dehydroisomerization, thereby resulting in production of
(1.0 mol%), concentrations (1.0 and 2.0 M) and/or tempera- oligomers as opposed to the production of 1. Most of the
tures (100  C) gave lower yields of 1, even when longer oligomers were dimeric (75%), with the remainder con-
reaction times (up to 24 hours) were investigated. sisting of trimers (20%) and tetramers (5%), as deter-
mined by GC–MS analysis. It is likely that such lower MW
Attempted Synthesis of p-Cymene from Limonene and oligomers arose via a cationic mechanism, given the acidic
Pinene reaction conditions employed herein. Cationic homo-
polymerization of 2 to high MW is normally performed
The optimum reaction conditions for production of 1 from with Friedel-Crafts or Ziegler-Natta catalysts, and copoly-
4 (110  C, 2.0 mol% pTsOH, 3.0 M in toluene) were merization with alkenoic comonomers proceeds by free
applied to the synthesis of 1 from (+/−)-2 (Fig. 6a), α-3 radical mechanisms (Zhang and Dube, 2014).
(Fig. 6b) and β-3. As seen in Fig. 6a, synthesis of 1 from When the operational parameters of 110  C, 2.0 mol%
(+/−)-2 did not proceed as readily as that for 1 from 4. pTsOH, and 3.0 M in toluene were applied to α-3 (Fig. 6b),
Although (+/−)-2 was consumed within 30 min, the maxi- only 3% of 1 was obtained after 6 hours. A significant
mum yield of 1 was only 5%, with the remainder consisting amount of camphene (44%), terpinenes (15%), 2 (14%),

Fig 5 Conversion of perillyl alcohol (PA) to p-cymene (CY) in 3.0 M toluene at 110  C catalyzed by (a) 1.0 and (b) 2.0 mol% para-
toluenesulfonic acid

J Am Oil Chem Soc (2021) 98: 305–316


312 J Am Oil Chem Soc

Fig 6 Attempted dehydroisomerization of (a) dl-limonene (LIM) and (b) α-pinene (API) in 3.0 M toluene at 110  C catalyzed by 2.0 mol%
pTsOH

and a small percentage of oligomers (4%) were the primary via a cationic mechanism, given the acidic reaction condi-
products. The presence of 2 and its terpinene derivatives tions utilized herein.
indicated incomplete conversion of α-3 to 1, as they repre- In summary, synthesis of 1 from (+/−)-2, α-3, and β-3
sent intermediates in the mechanistic pathway of α-3 to was hindered by the strong Brønsted acidity of pTsOH. At
1 (Fig. 3). Camphene, however, was produced by a substrate concentration of 3.0 M in toluene, oligomeriza-
rearrangement of a key cationic intermediate resulting from tion was favored over dehydroisomerization when conver-
protonation of α-3, and its accumulation indicated that it sion of (+/−)-2 was attempted whereas camphene was
was preferred over 1 (and 2) under these conditions. Prior produced with α-3 and β-3. Under neat conditions, oligo-
reports noted that the ratio of camphene to 2 upon acid- merization of 2, α-3, and β-3 was strongly favored. Fur-
catalyzed isomerization of α-3 was dependent on acidity, ther studies are necessary to elucidate conditions
with stronger acids favoring camphene production favorable for production of 1 in high yield from (+/−)-2,
(Monteiro and Veloso, 2004; Volzone et al., 2001). Other α-3, and β-3. It is likely that catalysts with lower Brøn-
conditions, such as reaction temperature, the nature of the sted acidity are needed to mitigate competing oligomeri-
acid (Brønsted vs. Lewis), and pore geometry (in the case zation and rearrangement side reactions.
of zeolitic catalysts) reportedly impacted the ratio of cam-
phene to 2. Thus, we speculate that the strong Brønsted Proposed Mechanism
acidity (pKa −1.3) of pTsOH favored the production of
camphene over 2 and its terpinene derivatives, thereby Ideal dehydroisomerization of 2 and 3 proceeds through
suppressing the yield of 1. numerous terpinene (α-, β-, γ-, and δ-) and phellandrene
Application of the same reaction conditions to β-3 gave (α- and β-) intermediates, ultimately resulting in the libera-
faster consumption of starting material and greater accumu- tion of equimolar H2 gas (Fig. 3) (Al-Wadaani et al., 2009;
lation of oligomers relative to α-3 (results not shown). Such Golets et al., 2013; Roberge et al., 2001; Thomas and
a result was attributed to the higher reactivity of β-3 due to Bessiere, 1989). In contrast, the proposed mechanism for
its more reactive terminal exocyclic double bond vs. the conversion of 4 to 1 entails initial dehydration, followed by
more highly substituted endocyclic double bond of α-3. a sequence of p-menthatriene intermediates (I–III) before
When conducted neat but with otherwise similar condi- producing 1 (Fig. 7). Consequently, equimolar H2O as
tions, both α-3 and β-3 were converted entirely to oligo- opposed to H2 is liberated and terpinenes and phellandrenes
mers with MW > trimers. Oligomerization of α-3 and β-3 do not serve as intermediates during the mechanistic
during dehydroisomerization was generally ignored in pre- sequence. Such mechanistic aspects were confirmed by GC–
vious reports, but was noted in Nie et al., 2014 and MS, as terpinenes and phellandrenes were not observed dur-
Roberge et al., 2001, who attributed it to strong Brønsted ing production of 1 from 4 but compounds with molecular
acidity. Similar to (+/−)-2, α-3, and β-3 likely oligomerized weights (MW) corresponding to p-menthatrienes (I–III;

J Am Oil Chem Soc (2021) 98: 305–316


J Am Oil Chem Soc 313

Fig 7 Proposed mechanism for conversion of perillyl alcohol (4) to p-cymene (1) with liberation of water and gradual accumulation of dimer VII

(Fig. 5) were detected. Conversion of II to III and III to could then proceed through intermediates I–III shown in
1 involved a series of sigmatropic 1,3-hydride shifts, which Fig. 7 before producing 1. However, decomposition of VI
was similar mechanistically to the interconversion of ter- to IV and 4 was not possible because cleavage of a carbon–
pinenes and phellandrenes depicted in Fig. 3. The different carbon double bond would have been necessary. Instead,
mechanisms involved in dehydroisomerization of (+/−)-2, protonation and subsequent elimination of the hydroxyl
α-3, and β-3 (Fig. 3) vs. 1 (Fig. 7) as well as the higher reac- moiety of VI as water yielded relatively unreactive penta-
tivity of (+/−)-2, α-3, and β-3 toward pTsOH relative to 4, unsaturated dimer VII, which had an MW of 268.2. There-
we propose were the primary reasons behind the excellent fore, the dimer remaining at the conclusion of the reaction
yield of 1 from 4 but poor yield from (+/−)-2, α-3, and β-3. was VII (or its isomers), which had slowly accumulated
As noted in Figs. 4 and 5, dimer production reached a from VI, which was the minor dimer formed by condensa-
maximum before declining as 1 accumulated. Mass spectral tion of the alkenoic moiety of 4 with IV.
results indicated that 93% of dimers near the beginning of A minor amount of 3-isopropyl-6-methylcyclohex-2-en-
the reaction had an MW of 286.2 while the remainder had 1-one (5) was produced during the synthesis of 1 from 4, as
MW = 268.2. By the conclusion of the reaction, the dimer shown in Fig. 5. The structure of 5 was indicated by mass
MW was almost exclusively 268.2. We speculate that spectral analysis and confirmed by NMR spectroscopy.
dimers were produced via nucleophilic attack of 4 on sec- Experiments conducted under an inert atmosphere (Argon)
ondary carbocation IV, as depicted in Fig. 7. Attack of the utilizing carefully dried solvent and glassware provided 5 in
nucleophilic oxygen of 4 on IV produced dimer V whereas the same yield (3%) as reactions conducted less meticulously,
attack of the pi electrons of the isopropenyl moiety gave thereby suggesting that the oxygen from the ketone moiety of
dimer VI, both of which had an MW of 286.2. The nucleo- 5 originated from 4 as opposed to ambient H2O or O2.
philicity of the hydroxyl moiety was significantly greater Lastly, as expected, o- and m-cymene were not observed
than the alkene, thereby resulting in a higher amount of during these reactions, nor was n-propyltoluene. Such
V produced vs. VI (13.2:1). Subsequent acid-catalyzed pro- results indicated that unwanted isomerization and trans-
tonation of V resulted in decomposition to 4 and IV, which alkylation reactions either did not occur or did so in

J Am Oil Chem Soc (2021) 98: 305–316


314 J Am Oil Chem Soc

imperceptible amounts. Avoidance of these byproducts rep- Conclusions


resented an advantage over the traditional Friedel-Crafts
petrochemical approach as well as approaches using zeo- The highest yield of 1 (91%) was obtained when starting
litic heterogeneous catalysis. with 4 as a 3.0 M solution in toluene in 3 hours at 110  C
with 2.0 mol% pTsOH. Lower catalyst loadings, concentra-
tions, and/or temperatures gave lower yields of 1, even
Green Metrics when longer reaction times of up to 24 hours were
attempted. From a green chemistry perspective, neat condi-
Conversion of 4 to 1 was an inherently green process since tions had a lower environmental impact (E-factor of 0.4)
it adhered to numerous principles of green chemistry than otherwise similar conditions at 3.0 M (2.7) despite a
(Anastas and Eghbali, 2010; Sheldon, 2012). Several met- lower yield of 1 (86%) when solvent was not used. Mecha-
rics measure sustainability and “greenness,” among those nistically, the reaction entailed dehydration followed by a
are RCC, CE, and AE (Constable et al., 2002; Curzons series of p-menthatriene intermediates (I–III) and dimers
et al., 2001; Sheldon, 2018; Trost, 1991). The RCC and CE (V–VII) before giving 1. Such mechanistic aspects were
of 1 were 100% since all carbons from 4 were incorporated confirmed by GC–MS, as compounds with MW
into 1. Because 4 was dehydrated to form 1, the AE was corresponding to p-menthatrienes and dimers were detected
88.1% due to loss of H2O. For comparison, the RCC, CE, during production of 1.
and AE calculated for dehydroisomerization of (+/−)-2, α-3, Synthesis of 1 from (+/−)-2, α-3, and β-3 utilizing simi-
and β-3 to 1 were 100, 100, and 98.5%, respectively. The lar operational parameters was hindered (yield <5%) by the
higher AE for (+/−)-2, α-3, and β-3 (98.5%) vs. 4 (88.1%) strong Brønsted acidity (pKa −1.3) of the pTsOH catalyst.
was attributed to loss of H2 as opposed to H2O. At 3.0 M, oligmerization was favored over
The E-factor is defined as the ratio of everything but the dehydroisomerization when conversion of (+/−)-2 was
desired product (waste) to the mass of product. It differs attempted whereas camphene, limonene, and terpinenes
significantly from the other green metrics because it takes were primarily produced with α-3 and β-3. Under neat con-
chemical yield into account and includes all reagents, sol- ditions oligomerization was strongly favored in all cases.
vents, and processing aids. Lower E-factors are indicative Given the acidic reaction conditions, it is likely that the
of less waste and lower environmental impact. Thus, the oligomers formed via a cationic mechanism. Thus, the
ideal E factor is zero (Sheldon, 2007). E-factors for produc- greater reactivity of (+/−)-2, α-3, and β-3 toward pTsOH
tion of 1 from (+/−)-2, α-3, β-3, and 4 were calculated relative to 4 were the primary reasons behind the excellent
from the highest yielding reactions that employed a starting yield of 1 from 4 but poor yield from (+/−)-2, α-3, and β-3.
substrate concentration of 3.0 M in toluene. Primarily
because H2 or H2O were eliminated and toluene was uti- Acknowledgments This research was supported by the
lized as solvent, the lowest (best) theoretical E-factors for U.S. Department of Agriculture, Agricultural Research Service. Ms
Benetria N. Banks and Ms Judy Blackburn are acknowledged for
production of 1 in quantitative yield from (+/−)-2, α-3, excellent technical assistance and Dr Karl E. Vermillion is acknowl-
β-3, and 4 were 2.3, 2.3, 2.3, and 2.4, respectively. For edged for acquisition of NMR spectra.
neat conditions, the theoretical optimum E-factors would
be much lower at 0.1, 0.1, 0.1, and 0.3, respectively, Conflict of Interest The authors declare that they have no conflict
of interest.
thereby closely approaching the ideal E-factor of 0. The
experimental E-factors obtained from production of 1
from (+/−)-2 (64.5), α-3 (93.3), and β-3 (98.7) were con-
siderably higher than the corresponding value from 4 (2.7) References
due to the significantly higher yield with 4 than from the
Ajaikumar, S., Golets, M., Larsson, W., Shchukarev, A., Kordas, K.,
other monoterpenes. The E-factor determined for produc- Leino, A. R., & Mikkola, J. P. (2013) Effective dispersion of Au
tion of 1 from 4 utilizing 3.0 M solvent (2.7) was close to and Au-M (M = Co, Ni, Cu and Zn) bimetallic nanoparticles
the theoretical optimum value (2.4) as a result of the high over TiO2 grafted SBA-15: Their catalytic activity on
observed yield of 91%. The E-factor utilizing neat condi- dehydroisomerization of α-pinene. Microporous and Mesoporous
Materials, 173:99–111.
tions (0.4) was also close to the corresponding theoretical Al-Wadaani, F., Kozhevnikova, E. F., & Kozhevnikov, I. V. (2009)
optimum value of 0.3 under neat conditions due to a yield Zn(II)-Cr(III) mixed oxide as efficient bifunctional catalyst for
of 86%. Thus, from a green chemistry perspective, as indi- dehydroisomerization of α-pinene to p-cymene. Applied Catalysis
cated by E-factor, neat conditions had a lower environ- A: General, 363:153–156.
Anastas, P., & Eghbali, N. (2010) Green chemistry: Principles and
mental impact than otherwise similar processing
practice. Chemical Society Reviews, 39:301–312.
conditions at 3.0 M despite the lower yield when solvent Anis, E., Zafeer, M. F., Firdaus, F., Islam, S. N., Fatima, M., &
was not used. Mobarak Hossain, M. (2018) Evaluation of phytochemical potential

J Am Oil Chem Soc (2021) 98: 305–316


J Am Oil Chem Soc 315

of perillyl alcohol in an in vitro Parkinson’s disease model. Drug Lesage, P., Candy, J. P., Hirigoyen, C., Humblot, F., & Basset, J. M.
Development Research, 79:218–224. (1996) Selective dehydrogenation of dipentene (R-(+)-limonene)
Arockiam, P. B., Bruneau, C., & Dixneuf, P. H. (2012) Ruthenium into paracymene on silica supported palladium assisted by α-olefins
(II)-catalyzed C-H bond activation and functionalization. Chemical as hydrogen acceptor. Journal of Molecular Catalysis A: Chemical,
Reviews, 112:5879–5918. 112:431–435.
Asbaghian, S., Shafaghat, A., Zarea, K., Kasimov, F., & Salimi, F. Linnekoski, J. A., Asikainen, M., Heikkinen, H., Kaila, R. K.,
(2011) Comparison of volatile constituents, and antioxidant and Rasanen, J., Laitinen, A., & Harlin, A. (2014) Production of
antibacterial activities of the essential oils of Thymus caucasicus, p-cymene from crude sulphate terpentine with commercial zeolite
T. kotschyanus, T. vulgaris. Natural Product Communications, 6: catalyst using a continuous flow fixed bed reactor. Organic Process
137–140. Research & Development, 18:1468–1475.
Barman, S., Maity, S. K., & Pradham, N. C. (2005) Alkylation of tol- Lycourghiotis, S., Makarouni, D., Kordouli, E., Bourikas, K.,
uene with isopropyl alcohol catalyzed by Ce-exchanged NaX zeo- Kordulis, C., & Dourtoglou, V. (2018) Activation of natural mor-
lite. Chemical Engineering Journal, 114:39–45. denite by various acids: Characterization and evaluation in the
Berkowitz, B. J., & Grunwald, E. (1961) Molar refraction as an index transformation of limonene into p-cymene. Molecular Catalysis,
of proton transfer: An estimate of the acid strength of 450:95–103.
p-toluenesulfonic acid. Journal of the American Chemical Society, Makarouni, D., Lycourghiotis, S., Kordouli, E., Bourikas, K.,
83:2956–2956. Kordulis, C., & Dourtoglou, V. (2018) Transformation of limonene
Brzozowski, R., Dobrowolski, J. C., & Jamroz, M. H. (2002) Theoret- into p-cymene over acid activated natural mordenite utilizing atmo-
ical estimation of isomeric composition of cymenes in equilibrium spheric oxygen as a green oxidant: A novel mechanism. Applied
mixture. Catalysis Communications, 3:141–144. Catalysis B: Environmental, 224:740–750.
Buhl, D., Weyrich, P. A., Sachtler, W. M. H., & Holderich, W. F. Martin-Luengo, M. A., Yates, M., Martinez Domingo, M. J.,
(1998) Support effects in the Pd catalyzed dehydrogenation of ter- Casal, B., Iglesias, M., Esteban, M., & Ruiz-Hitzky, E. (2008) Syn-
pene mixtures to p-cymene. Applied Catalysis A: General, thesis of p-cymene from limonene, a renewable feedstock. Applied
171:1–11. Catalysis B: Environmental, 81:218–224.
Ciriminna, R., Lomeli-Rodriguez, M., Cara, P. D., Lopez- Martin-Luengo, M. A., Yates, M., Saez Rojo, E., Huerta Arribas, D.,
Sanchez, J. A., & Pagliaro, M. (2014) Limonene: A versatile chem- Aguilar, D., & Ruiz Hitzky, E. (2010) Sustainable p-cymene and
ical of the bioeconomy. Chemical Communications, 50: hydrogen from limonene. Applied Catalysis A: General, 387:
15288–15296. 141–146.
Constable, D. J. C., Curzons, A. D., & Cunningham, V. L. (2002) Monteiro, J. L. F., & Veloso, C. O. (2004) Catalytic conversion of ter-
Metrics to ‘green’ chemistry – Which are the best? Green Chemis- penes into fine chemicals. Topics in Catalysis, 27:169–180.
try, 4:521–527. Moser, B. R. (2014) Preparation and evaluation of multifunctional
Curzons, A. D., Constable, D. J. C., Mortimer, D. N., & branched diesters as fuel property enhancers for biodiesel and
Cunningham, V. L. (2001) So you think your process is green, how petroleum diesel fuels. Energy and Fuels, 28:3262–3270.
do you know? – Using principles of sustainability to determine Moser, B. R., Doll, K. M., & Peterson, S. C. (2019) Renewable
what is green – A corporate perspective. Green Chemistry, 3:1–6. poly(thioether-ester)s from fatty acid derivatives via thiol-ene
Du, J., Xu, H., Shen, J., Huang, J., Shen, W., & Zhao, D. (2005) Cata- photopolymerization. Journal of the American Oil Chemists’ Soci-
lytic dehydrogenation and cracking of industrial dipentene over ety, 96:825–837.
M/SBA-15 (M = Al, Zn) catalysts. Applied Catalysis A: General, Moser, B. R., Vermillion, K. E., Banks, B. N., & Doll, K. M. (2020)
296:186–193. Renewable aliphatic polyesters from fatty dienes by acyclic diene
Erland, L. A. E., Bitcon, C. R., Lemke, A. D., & Mahmoud, S. S. metathesis polymerization. Journal of the American Oil Chemists’
(2016) Antifungal screening of lavender essential oil constituents Society, 97:517–530.
on three post-harvest fungal pathogens. Natural Product Communi- Nie, G., Zou, J. J., Feng, R., Zhang, X., & Wang, L. (2014)
cations, 11:523–527. HPW/MCM-41 catalyzed isomerization and dimerization of pure
Fernandes, C., Catrinescu, C., Castilho, P., Russo, P. A., pinene and crude terpentine. Catalysis Today, 234:271–277.
Carrott, M. R., & Breen, C. (2007) Catalytic conversion of limo- Odedairo, T., & Al-Khattaf, S. (2011) Alkylation and transalylation of
nene over acid activated Serra de Dentro (SD) bentonite. Applied alkylbenzenes in cymene production over zeolite catalysts. Chemi-
Catalysis A: General, 318:108–120. cal Engineering Journal, 167:240–254.
Gandeepan, P., Kaplaneris, N., Santoro, S., Vaccaro, L., & Perego, C., Amarilli, A., Carati, A., Flego, C., Pazzuconi, G.,
Ackermann, L. (2019) Biomass-derived solvents for sustainable Rizzo, C., & Bellussi, G. (1999) Mesoporous silica-aluminas as cat-
transition metal-catalyzed C-H activation. ACS Sustainable Chem- alysts for the alkylation of aromatic hydrocarbons with olefins.
istry & Engineering, 7:8023–8040. Microporous and Mesoporous Materials, 27:345–354.
Geoghegan, K., & Evans, P. (2014) Synthesis of (+)-perillyl alcohol Perego, C., & Ingallina, P. (2002) Recent advances in the industrial
from (+)-limonene. Tetrahedron Letters, 55:1431–1433. alkylation of aromatics: New catalysts and new processes. Catalysis
Golets, M., Ajaikumar, S., & Mikkola, J. P. (2015) Catalytic Today, 73:3–22.
upgrading of extractives to chemicals: Monoterpenes to Roberge, D. M., Buhl, D., Niederer, J. P. M., & Holderich, W. F.
“EXCALS”. Chemical Reviews, 115:3141–3169. (2001) Catalytic aspects in the transformation of pinenes to
Golets, M., Ajaikumar, S., Mohln, M., Warna, J., Rakesh, S., & p-cymene. Applied Catalysis A: General, 215:111–124.
Mikkola, J. P. (2013) Continuous production of the renewable Sheldon, R. A. (2007) The E factor: Fifteen years on. Green Chemis-
p-cymene from α-pinene. Journal of Catalysis, 307:305–315. try, 9:1273–1283.
Johnson, J. A., & Kabza, R. G. (1990) SORBEX. Industrial-scale Sheldon, R. A. (2012) Fundamentals of green chemistry: Efficiency in
adsorptive separation. American Institute of Chemical Engineers reaction design. Chemical Society Reviews, 41:1437–1451.
Symposium Series, 118:35–54. Sheldon, R. A. (2018) Metrics of green chemistry and sustainability:
Kamitsou, M., Panagiotou, G. D., Triantafyllidis, K. S., Bourikas, K., Past, present, and future. ACS Sustainable Chemistry & Engineer-
Lycourghiotis, A., & Kordulis, C. (2014) Transformation of ing, 6:32–48.
α-limonene into p-cymene over oxide catalysts: A green chemistry Tavera Ruiz, C. P., Gauthier-Maradei, P., Capron, M., Pirez, C.,
approach. Applied Catalysis A: General, 474:224–229. Gardoll, O., Katryniok, B., & Dumeignil, F. (2019) Transformation

J Am Oil Chem Soc (2021) 98: 305–316


316 J Am Oil Chem Soc

of DL limonene into aromatic compounds using supported hetero- Weyrich, P. A., & Holderich, W. F. (1997) Dehydrogenation of
polyacid catalysts. Catalysis Letters, 149:328–337. α-limonene over Ce promoted, zeolite supported Pd catalysts.
Thomas, A. F., & Bessiere, Y. (1989) Limonene. Natural Product Applied Catalysis A: General, 158:145–162.
Reports, 6:291–309. Wichterlova, B., & Cejka, J. (1994) Mechanism of n-propyltoluene
Trost, B. M. (1991) The atom economy: A search for synthetic effi- formation in C3 alkylation of toluene: The effect of zeolite struc-
ciency. Science, 254:1471–1477. tural type. Journal of Catalysis, 146:523–529.
van Beilen, J. B., Holtackers, R., Luscher, D., Bauer, U., Winnacker, M. (2018) Pinenes: Abundant and renewable building
Witholt, B., & Duetz, W. A. (2005) Biocatalytic production of blocks for a variety of sustainable polymers. Angewandte Chemie
perillyl alcohol from limonene by using a novel Mycobacterium International Edition, 57:14362–14371.
sp. cytochrome P450 alkane hydroxylase expressed in Pseudomo-
Yilmazoglu, E., & Akgun, M. (2018) p-cymene production from
nas putida. Applied and Environmental Microbiology, 71:
orange peel oil using some metal catalyst in supercritical fluids.
1737–1744.
Journal of Supercritical Fluids, 131:37–46.
Volanti, M., Cespi, D., Passarini, F., Neri, E., Cavani, F.,
Mizsey, P., & Fozer, D. (2019) Terephthalic acid from renewable Zhang, Y., & Dube, M. A. (2014) Copolymerization of n-butyl methac-
resources: Early-stage sustainability analysis of a bio-PET precur- rylate and D-limonene. Macromolecular Reaction Engineering, 8:
sor. Green Chemistry, 21:885–896. 805–812.
Volzone, C., Masini, O., Comelli, N. A., Grzona, L. M., Zhao, C., Gan, W., Fan, X., Cai, Z., Dyson, P. J., & Kou, Y. (2008)
Ponzi, E. N., & Ponzi, M. I. I. (2001) Production of camphene and Aqueous-phase biphasic dehydroaromatization of bio-derived limo-
limonene from pinene over acid di- and tri-octahedral smectite nene into p-cymene by soluble Pd nanocluster catalysts. Journal of
clays. Applied Catalysis A: General, 214:213–218. Catalysis, 254:244–250.
Wanner, J., Bail, S., Jirovetz, L., Buchbauer, G., Schmidt, E., Zilkova, N., Eliasova, P., Al-Khattaf, S., Morris, R. E., Mazur, M., &
Gochevd, V., … Stoyanova, A. (2010) Chemical composition and Cejka, J. (2016) The effect of UTL layer connectivity in isoreticular
antimicrobial activity of cumin oil (Cumin cyminum, Apiaceae). zeolites on the catalytic performance in toluene alkylation. Cataly-
Natural Product Communications, 5:1355–1358. sis Today, 277:55–60.

J Am Oil Chem Soc (2021) 98: 305–316

You might also like