Sei sulla pagina 1di 6

Catalysis Today 263 (2016) 1621

Contents lists available at ScienceDirect

Catalysis Today
journal homepage: www.elsevier.com/locate/cattod

Reactivity of levulinic acid during aqueous, acid-catalyzed HMF


hydration
Shruti Karwa, Varun M. Gajiwala, Jacob Heltzel, Sushil K.R. Patil, Carl R.F. Lund
Department of Chemical and Biological Engineering University at Buffalo, SUNY Buffalo, NY 14260, United States

a r t i c l e

i n f o

Article history:
Received 2 April 2015
Received in revised form 9 June 2015
Accepted 13 June 2015
Available online 3 August 2015
Keywords:
Levulinic acid dehydration
HMF hydration
Humins formation
Angelica lactones

a b s t r a c t
In 0.1 M sulfuric acid at 125 C, levulinic acid did not form humins, and if HMF was present, levulinic acid
was not incorporated in the humins that formed from it. Levulinic acid was converted to unidentied
products believed to include angelica lactones, at a rate more than 500 times smaller than the rate
of HMF conversion. Quantum chemical calculations indicate that at reaction conditions, levulinic acid
is protonated to form protonated dihydro-5-hydroxy-5-methyl-2(3H)-furanone which is very stable.
Very small amounts of levulinic acid and 5-hydroxy--valerolactone may be present in equilibrium with
this cation. Angelica lactones can then be formed from 5-hydroxy--valerolactone in an acidic aqueous
environment.
2015 Elsevier B.V. All rights reserved.

1. Introduction
Glucose, 5-hydroxymethylfurfural (HMF) and levulinic acid
are sequentially formed during aqueous, mineral acid-catalyzed
hydrolysis of cellulose. Similarly, the acid-catalyzed conversion of
glucose, fructose or HMF leads to the formation of levulinic acid.
Apparently levulinic acid is stable at reaction conditions. Stoichiometry suggests it is produced in a 1:1 ratio with formic acid, but
experimentally observed ratios of levulinic to formic acid that are
greater than 1:1 [1], equal to 1:1 [24] and less than 1:1 [5,6] have
been reported. In addition, Girisuta et al. [2] have shown that neither levulinic acid nor formic acid decomposes at 141 C in 1 M
H2 SO4 .
Concurrent with the generation of levulinic acid, the acidcatalyzed conversion of cellulose, cellobiose, glucose, fructose or
HMF also produces undesired solids known as humins [2,4,7]. Horvat et al. [8] suggested that 2,5-dioxo-6-hydroxyhexanal (DHH)
is an intermediate formed from HMF that leads to formation of
humins. DHH is presumed to be highly reactive because it is present
in very low concentrations that have precluded its isolation from
the reaction mixture [8,9]. It was later suggested that humins form
via the aldol addition and condensation of DHH with aldehydes and
ketones that are available in the reacting solution [3,10]. In support of this proposed pathway for humin formation, it has been
shown that when HMF is used as the reactant, its furan ring is

Corresponding author. Tel.: +1 716 645 1180; fax: +1 716 645 3822.
E-mail address: lund@buffalo.edu (C.R.F. Lund).
http://dx.doi.org/10.1016/j.cattod.2015.06.020
0920-5861/ 2015 Elsevier B.V. All rights reserved.

incorporated into the humins; and if benzaldehyde is added to the


reacting solution, its benzene ring is incorporated in the humins
[3].
If DHH is highly reactive toward aldehydes and ketones, it is
somewhat surprising that the ketone group of levulinic acid does
not react, leading to the incorporation of levulinic acid in the
humins and reduction of the levulinic to formic acid ratio below
its stoichiometric value of 1. Levulinic acid is reported to form
enols, 5-hydroxy-gamma-valerolactone, and other related unsaturated lactones [1114]. The present investigation was undertaken
to better understand the relative stability of levulinic acid during
acid-catalyzed hydrolysis of carbohydrates such as glucose.
2. Experimental and computational methods
The selectivity of acid-catalyzed HMF conversion was studied in
a manner described previously [3]. Briey, thick walled glass vials
were charged with a solution containing 0.1 M H2 SO4 and known
concentrations of HMF, levulinic acid and formic acid. The vials
were sealed using septa underneath screw-on plastic caps. Sets of
ve vials were rapidly heated to the desired reaction temperature
at which time one vial was removed and quenched. The remaining
vials were maintained at the desired reaction temperature until,
one by one, they were removed and quenched after progressively
longer reaction times. The liquid in the quenched vials was analyzed using an Agilent 1200 Series HPLC with a Bio-rad Aminex
HPX-87H column and a refractive index detector. The HPLC mobile
phase was 5 mM sulfuric acid. Solutions of the pure components
at known concentrations were prepared and analyzed to generate

S. Karwa et al. / Catalysis Today 263 (2016) 1621

concentration calibration curves for HMF, levulinic acid and formic


acid.
The GAMESS computational chemistry program [15] was used
to determine the free energies of reactants, products and suspected
intermediates involved in the reactions of levulinic acid. A typical
set of calculations began with geometry optimization via molecular
mechanics using the MMFF94 force eld parameters incorporated
in the Avogadro molecular visualization program [16]. For acyclic
molecules, several (100) rotational conformations were randomly
sampled and the one with the lowest energy was then used as the
initial geometry for quantum chemical calculations.
The quantum chemical calculations were performed at the
G3(MP2, CCSD(T)) level of theory that is implemented in GAMESS
[1719]. In these calculations, the geometry is optimized, nuclear
forces and vibrational frequencies are computed, and the energy is
computed, correcting for most computational shortcomings. Based
upon calculations on the 299 molecules in the G2/97 test set, this
level of theory is expected to be accurate to within ca. 5.61 kJ mol1
(1.34 kcal mol1 ). This was considered to be sufciently accurate for
present purposes, particularly since a full conformational analysis
was not practical. The results of the G3(MP2, CCSD(T)) computations were used to determine the gas phase Gibbs free energy of
each species considered at 298 K. Henceforth these will simply be
referred to as G3 calculations with the understanding that the MP2,
CCSD(T) variant of G3 was used.
The free energy of solvation in water was also determined for
each species using its equilibrium gas phase geometry as determined in the G3 calculations. In a few cases, the geometry was
re-optimized using the SMD model and the corresponding free
energy of solvation was calculated. The differences were not signicant, and consequently the equilibrium gas phase geometry
was used in all calculations reported here. The SMD polarized continuum model that is incorporated within GAMESS was used for
these calculations [20]. The solvation free energy was computed at
the same level of theory as the G3 geometry optimization, namely
MP2(full)/6-31G(d). A 1 atm. standard state for free energies of solvation was used in the present study.

3. Results and discussion


3.1. Conversion and selectivity
It is quite possible that reactant (i.e. glucose vs. fructose vs. HMF,
etc.) and reaction conditions both affect the levulinic acid to formic
acid ratio during acid-catalyzed hydrolysis. Here, the reaction of an
aqueous solution of 0.1 M HMF and 0.1 M H2 SO4 at 125 C will be
examined. Three experiments were performed; one using only the
0.1 M H2 SO4 and 0.1 M HMF as reactants, one where 0.1 M levulinic
acid was added and one where 0.05 M formic acid was added. In
each experiment the conversion of the HMF and the ratio of the
produced levulinic acid to the produced formic acid were measured.
Fig. 1 shows that to within experimental uncertainty levulinic acid
and formic acid were produced in the expected stoichiometric ratio
of one to one over the full range of HMF conversion. This was also
true for the two experiments where either excess levulinic acid or
excess formic acid was present. In all experiments, the selectivity
for levulinic and formic acids was between 80% and 85%.
The more important result in Fig. 1 is that the levulinic to formic
acid production ratio is constant and equal to one in all runs. Small
amounts of humins do form during these experiments. We have
previously shown that small and nearly equal amounts of both
levulinic acid and formic acid are retained on the humins surface
and can be removed by repeated hot water washing. The retained
amounts have a negligible effect on the apparent conversion, and

17

Fig. 1. Acid-catalyzed, 125 C HMF conversion (diamonds) and levulinic-formic acid


production ratio (triangles) using 0.1 M H2 SO4 and 0.1 M HMF alone (unlled), with
0.1 M added levulinic acid (gray) and with 0.05 M added formic acid (lled).

more importantly, since the retained amounts of the two acids are
nearly the same, the levulinic to formic acid ratio is not affected.
Noting that the expected stoichiometry would lead to a levulinic
to formic acid ratio of 1, the results in Fig. 1 show that their net rates
of production are equal. Nonetheless, Fig. 1 does not unequivocally
demonstrate that levulinic acid and formic acid are unreactive. It
is conceivable that they do react, either with each other or individually at equal rates. To probe this possibility the reaction of
aqueous solutions of 0.1 M H2 SO4 containing (a) 0.1 M levulinic
acid, (b) 0.1 M formic acid and (c) both 0.1 M levulinic acid and 0.1 M
formic acid was examined at 125 C. Fig. 2 shows that in all cases, the
concentrations of levulinic acid and formic acid were invariant, to
within experimental error, over periods of 180275 min. This time
scale corresponds to that required for the complete conversion of
HMF and is consistent with past studies nding levulinic acid to
be relatively unreactive at HMF processing conditions [24]. We
also were interested in determining whether levulinic acid alone
might react in 0.1 M sulfuric acid over a longer time scale. Table 1
shows that when an aqueous solution of sulfuric (0.1 M) and levulinic (0.05 M) acids was allowed to react at 125 C over a period
of 72 h, the levulinic acid concentration decreased by 7% (A replicate experiment using 0.1 M levulinic acid and 0.1 M sulfuric acid
at 125 C and a 100 h reaction time resulted in a 5.4% decrease in
the levulinic acid concentration.). It should be noted that at comparable conditions, HMF is completely converted after ca. 4 h. If the

Fig. 2. Fractional concentration changes of formic and levulinic acid reacting individually and simultaneously in 0.1 M H2 SO4 at 125 C.

18

S. Karwa et al. / Catalysis Today 263 (2016) 1621

Table 1
Relative levulinic acid concentration during processing at a 0.1 M concentration in
0.1 M H2 SO4 at 125 C.
Time (min)

LA(t)/LA(0)

0
30
60
90
120
180
1440
4320

1.00
1.00
1.01
1.03
0.99
0.99
0.96
0.93

kinetics are assumed to be rst order, the rate coefcient for the
conversion of HMF is 550 times greater than the rate coefcient for
the conversion of levulinic acid.
Indeed, the decrease in the levulinic acid concentration is small
enough that one might question whether it is meaningful relative
to experimental measurement uncertainties. We believe that it is
meaningful because four new, unidentied peaks started to appear
in the HPLC chromatograms after 24 h, and all four of those peaks
increased after 72 h of processing. They were also observed in the
100 h replicate experiment. The largest of the four unidentied
peaks had an area equal to 5.1% of the area of the levulinic acid peak
after 72 h; the other three had areas equal to 3.1%, 1.8% and 1.3% of
the area of the levulinic acid peak. None of these HPLC peaks corresponds to -angelica lactone; our HPLC column would not separate
-angelica lactone from levulinic acid. We have not experimentally established the identity of the products, but the computational
results to follow suggest they may include other angelica lactones
or dimers thereof.
To summarize the experimental observations, levulinic acid is
quite stable over the time scale for the acid-catalyzed hydration
of HMF. More specically, levulinic acid is not incorporated into
humins that form during HMF processing. Additionally, when levulinic acid alone reacts in an aqueous acid environment, it does not
form humins. Over a much longer time scale, levulinic acid appears
to be slowly converted into four unidentied products. The rate of
this conversion is less than 0.2% of the rate of HMF hydration.
3.2. Protonation of levulinic acid
There is good evidence that aldol addition/condensation is a
signicant process in the formation of humins [3,10]. While levulinic acid has a keto group, the experimental results indicate
it is not incorporated in humins nor does it form them, itself.
Quantum chemical calculations were undertaken to explain the
observed stability of levulinic acid. The pKa of levulinic acid is
4.59, so in an aqueous solution with 0.1 M H2 SO4 , levulinic acid
is not expected to dissociate. In fact, at such a low pH, either of
the oxo groups of levulinic acid can be protonated. G3 calculations
show that protonation of the ketone group results in the formation
of protonated dihydro-5-hydroxy-5-methyl-2(3H)-furanone (I-1),
Fig. 3. Non-cyclic cation conformations are also possible but they
are metastable compared to the cyclic conformation, reaction (1).
The free energy of formation of non-cyclic cations is of the order
of G298K,aq = 6.7 kJ mol1 , based on one such structure that was
optimized, while formation of the cyclic cation is much more favorable: G298K,aq = 29.9 kJ mol1 .

Fig. 3. Optimized geometry of the cyclic cation (I-1) formed by protonation of the
ketone group of levulinic acid.

starting geometries always transformed to the cyclic conformation during G3 geometry optimization. Non-cyclic conformations,
if they exist, are also expected to be metastable relative to the
cyclic conformation. Reaction (2) was found to be nearly free energy
neutral, G298K,aq = 1.0 kJ mol1 .

(2)
The quantum chemical calculations indicate that under the reaction conditions studied, levulinic acid is present predominantly in
the form of the cyclic cation, I-1, resulting from protonation of the
ketone group followed by ring closure. As such, this cation, shown
in Fig. 3, is the rst intermediate formed in the conversion of levulinic acid. Quantum chemical calculations were used to study the
reactivity of I-1, and the results are discussed in Sections 3.33.5.
3.3. Isomerization of levulinic acid
Five isomers of levulinic acid, not counting levulinic acid itself,
can be formed upon de-protonation of I-1. The rst is 5-hydroxy-valerolactone, also known as pseudo levulinic acid, which forms
according to reaction (3). Fig. 4(a and b) shows that the free energy
for de-protonation of I-1 to form of 5-hydroxy--valerolactone by
reaction (3) is nearly the same as that for de-protonation to reform levulinic acid by the reverse of reaction (1): 32.0 kJ mol1 vs.
29.9 kJ mol1 .

(3)
Hydrogen transfer and de-protonation of I-1 will produce
5-hydroxy-pent-3-enoic acid (the 3-4 enol) and 5-hydroxy-pent5-enoic acid (the 4-5 enol). Timokhin et al. [11] noted that in D2 O

(1)
The protonation of the oxo group of the acid, reaction (2) leads to
protonated 2,2-dihydroxy-3-hydro-5-methyl furan (I-2). We were
not able to optimize a non-cyclic conformation of I-2. Non-cyclic

S. Karwa et al. / Catalysis Today 263 (2016) 1621

alone there is no exchange at C3 or C4 indicating that the enol


concentration is negligible. However, in acidic D2 O there is rapid
exchange at C3 and C5, leading Timokhin et al to argue that the
enol concentrations are much larger in acidic media. However, the
calculated free energies of formation of these species shown in
Fig. 4(c and d) indicate that even in acid media, their formation
is not favorable.
I-1 can also de-protonate to form dihydroxy furans. The
most likely pathway for this transformation would involve

19

work [3,10] has shown that the 2-3 enol of DHH is the most
likely of the four possible enols. The free energy change for that
addition, reaction (4) is G298K,aq = 69.5 kJ mol1 . On top of that,
the tautomerization of DHH to the 2-3 enol is slightly unfavorable, G298K,aq = 7.5 kJ mol1 . Similarly, aldol self-addition of I-1 to
another levulinic acid molecule would require that second levulinic
acid molecule to isomerize to an enol. Fig. 4 shows that enol formation is not expected; the free energy change for the formation of the
more favorable enol, 5-hydroxy-pent-3-enoic acid, is 78.0 kJ mol1 .

(4)
de-protonation to reform levulinic acid followed formation of I-2
by reaction (2) and another de-protonation leading to the products.
Fig. 4(e and f) shows that the formation of the dihydroxy furans is
even less favorable than enol formation.
Fig. 4 provides strong evidence that at reaction conditions,
almost all of the original levulinic acid will be present as I-1, perhaps with very small equilibrium amounts of levulinic acid and
5-hydroxy--valerolactone and negligibly small concentrations of
the enols and dihydroxy furans. This is consistent with the HPLC
analyses where only one elution peak is observed for aqueous acidic
solutions of levulinic acid. The mobile phase in the HPLC is 5 mM
H2 SO4 (pH 2), so within the HPLC, I-1 is essentially the only form of
levulinic acid present, and hence a single elution peak is observed.
3.4. Aldol addition of levulinic acid
I-1 could add to an aldehyde or ketone such as DHH or another
levulinic acid molecule. Reaction with DHH would lead to the
incorporation of levulinic acid in HMF-derived humins while aldol
self-addition would be expected to lead to the formation of humins
or other products from levulinic acid. The experimental results
in Section 3.1 indicate that neither of these reactions occurs to
a measurable extent. Addition of I-1 to DHH in an acidic environment requires that DHH rst be converted to an enol. Prior

Fig. 4. Aqueous Gibbs free energy for de-protonation of I-1 to form (a) levulinic acid,
(b) 5-hydroxy--valerolactone, (c) 5-hydroxy-pent-3-enoic acid (the 3-4 enol), (d)
5-hydroxy-pent-5-enoic acid (the 4-5 enol), (e) 2,2-dihydroxy-3-hydro-5-methyl
furan and (f) 2,2-dihydroxy-3,4-hydro-5-methylene furan.

These results show that aldol addition of I-1 to either DHH or


levulinic acid is not expected to occur. That is, participation of levulinic acid in aldol addition/condensation is not favorable. This is
consistent with the experimental observations that levulinic acid
alone does not form humins, and that during humin formation from
HMF, levulinic acid is not incorporated into the humins. Calculations at the same level of theory resulted in an energetic pathway
for conversion of HMF to levulinic acid that was consistent with
prior studies [10]. In those calculations a range of free energies were
found for aldol condensation of DHH with various ketones; some
were favorable and some were not. The result for levulinic acid was
intermediate in the resulting range of free energies. Therefore, the
present results can be accepted with some condence, though additional calculations on other experimentally studied systems would
be useful in calibrating the results. Clearly, I-1 is very stable in an
aqueous acidic environment as evidenced by the fact that reaction
only begins to become apparent after ca. 70 h, and the reaction that
does occur does not produce humins.
3.5. Dehydration of levulinic acid
Levulinic acid can be reversibly dehydrated to 3-penten-4-olide
(-angelica lactone) with or without an acid catalyst [12,21,22].
Acid catalysts or thermal treatment permit isomerization of
-angelica lactone, creating a mixture additionally containing
2-penten-4-olide (-angelica lactone) and 4-penten-4-olide (angelica lactone) [13,23]. Mascal et al. note that -angelica lactone
can also dimerize [22]. The reaction pathway has been suggested to
involve either enols (5-hydroxy-pent-3-enoic and 5-hydroxy-pent5-enoic acid) [11,21] or 5-hydroxy--valerolactone [12]. Guntrum
et al. [14] report that when -angelica lactone is treated with an
acid, levulinic acid is produced.
The present computational results, Fig. 4, show that 5-hydroxy-valerolactone is much more likely to be an intermediate in the
dehydration of levulinic acid to angelica lactones because its free
energy of formation from I-1 is less than half that of the more likely
of the two enols. In addition, it is easy to visualize a pathway starting
from 5-hydroxy--valerolactone, as shown in Fig. 5. As the gure
shows, protonation of the hydroxyl group and elimination of water
yields a protonated 5-methyl-2(3H)-furanone (I-3) with a Gibbs
free energy increase of only 4.5 kJ mol1 . Simple de-protonation
of I-3 leads to either - or -angelica lactone. The formation of
either of these angelica lactones involves a total increase of less than
10 kJ mol1 relative to 5-hydroxy--valerolactone. Isomerization of
-angelica lactone to produce -angelica lactone is uphill by less
than 4 kJ mol1 . The free energy of formation of all three angelica

20

S. Karwa et al. / Catalysis Today 263 (2016) 1621

that the angelica lactones are produced via the formation of enols,
as suggested in some studies.
Acknowledgements
The authors, particularly C.L., would like to express their deep
appreciation to Professor Eli Ruckenstein for his many insightful
comments on this work and on numerous previous occasions. The
use of the computational resources in the Center for Computational
Research at the University at Buffalo is also gratefully acknowledged.
Appendix A. Supplementary data
Supplementary data associated with this article can be found,
in the online version, at http://dx.doi.org/10.1016/j.cattod.2015.06.
020
Fig. 5. Aqueous Gibbs free energies, relative to I-1, for the formation of angelica
lactones from 5-hydroxy--valerolactone.

lactones is more than 30 kJ mol1 less than that for formation of


either enol.
The reaction of levulinic acid in sulfuric acid solution yielded
very small amounts of four unidentied products after 72 h. If
-angelica lactone was produced, we would not have been able
to detect it because its HPLC retention time is nearly the same
as levulinic acid. While none of the four detected peaks corresponds to -angelica lactone, the results shown in Fig. 5 suggest
that those products conceivably could be other angelica lactones
and/or dimers thereof [22]. Fig. 5 indicates that -angelica lactone
would be present in the largest concentration at equilibrium, and
-angelica lactone would be present in the smallest concentration. Experimental studies typically nd -angelica lactone in the
greatest concentration [22,23], but such systems may not be fully
equilibrated. At the same time, the free energy differences shown in
Fig. 5 are quite close and subject to uncertainty, as well. Additional
studies, both experimental and computational, will be required to
resolve the issue of the relative amounts of the individual angelica
lactones. Nonetheless, the results in Fig. 5 clearly suggest that I-3 is
more plausible as an intermediate for the formation of the angelica
lactones than are the enols.

4. Conclusions
At temperatures of 125 C in 0.1 M sulfuric acid, levulinic acid
does not form humins. If HMF is simultaneously present, levulinic
acid is not incorporated into the humins that form from HMF. Calculated free energies indicate that at these conditions, almost all
of the levulinic acid is protonated and exists in the form of protonated dihydro-5-hydroxy-5-methyl-2(3H)-furanone, shown in
Fig. 3. This cyclic cation is very stable; in particular de-protonation
to form either enols or dihydroxy furans is highly unfavorable. Two
de-protonation products, the original levulinic acid and 5-hydroxy-valerolactone are nearly equally probable and may be present
in small amounts in equilibrium with the protonated dihydro-5hydroxy-5-methyl-2(3H)-furanone.
The experiments also show that levulinic acid does react very
slowly at 125 C in 0.1 M sulfuric acid producing small amounts of
four unidentied products. Calculated free energies suggest that
if small amounts of 5-hydroxy--valerolactone are present, they
could easily be converted to angelica lactones. It is far less likely

References
[1] F.S. Asghari, H. Yoshida, Kinetics of the decomposition of fructose catalyzed by
hydrochloric acid in subcritical water: formation of 5-hydroxymethylfurfural,
levulinic, and formic acids, Ind. Eng. Chem. Res. 46 (2007)
77037710.
[2] B. Girisuta, L.P.B.M. Janssen, H.J. Heeres, A kinetic study on the decomposition
of 5-hydroxymethylfurfural into levulinic acid, Green Chem. 8 (2006)
701709.
[3] S.K.R. Patil, C.R.F. Lund, Formation and growth of humins via aldol addition
and condensation during acid-catalyzed conversion of
5-hydroxymethylfurfural, Energy Fuels 25 (2011) 47454755.
[4] B. Girisuta, L.P.B.M. Janssen, H.J. Heeres, Kinetic study on the acid-catalyzed
hydrolysis of cellulose to levulinic acid, Ind. Eng. Chem. Res. 46 (2007)
16961708.
[5] R. Weingarten, J. Cho, R. Xing, W.C. Conner Jr., G.W. Huber, Kinetics and
reaction engineering of levulinic acid production from aqueous glucose
solutions, Chemsuschem 5 (2012) 12801290.
[6] R. Weingarten, W.C. Conner Jr., G.W. Huber, Production of levulinic acid from
cellulose by hydrothermal decomposition combined with aqueous phase
dehydration with a solid acid catalyst, Energy Environ. Sci. 5 (2012)
75597574.
[7] B. Girisuta, L. Janssen, H. Heeres, A kinetic study on the conversion of glucose
to levulinic acid, Chem. Eng. Res. Des. 84 (2006) 339349.
[8] J. Horvat, B. Klaic, B. Metelko, V. Sunjic, Mechanism of levulinic acid
formation, Tetrahedron Lett. 26 (1985) 21112114.
[9] J. Horvat, B. Klaic, B. Metelko, V. Sunjic, Mechanism of levulinic acid formation
in acid catalyzed hydrolysis of 2-(hydroxymethyl)furan and
5-(hydroxymethyl)-2-furancarboxaldehyde, Croatica Chem. Acta 59 (1986)
429438.
[10] S.K.R. Patil, J. Heltzel, C.R.F. Lund, Comparison of structural features of humins
formed catalytically from glucose, fructose, and
5-hydroxymethylfurfuraldehyde, Energy Fuels 26 (2012) 52815293.
[11] B.V. Timokhin, V.A. Baransky, G.D. Eliseeva, Levulinic acid in organic
synthesis, Russ. Chem. Rev. 68 (1999) 7384.
[12] R.H. Leonard, Conversion of levulinic acid into -angelica lactone, U.S.
2809203, assigned to Heyden Newport Chemical Corp., 1957.
[13] A.B. Hornfeldt, Tautomeric properties and some reactions of the
2-hydroxythiophene and the 2-hydroxyfuran systems, Sv. Kem. Tidskr. 80
(1968) 343356.
[14] E. Guntrum, W. Kuhn, W. Spoenlein, V. Jaeger, Synthesis of 2-penten-4-olides,
3-penten-4-olides, and 4,4-dialkyl-1,3-cyclopentanediones by acid- and
base-induced isomerization of 4-penten-4-olides
(-methylene--butyrolactones), Synthesis (1986) 921925.
[15] M.W. Schmidt, K.K. Baldridge, J.A. Boatz, S.T. Elbert, M.S. Gordon, J.J. Jensen, S.
Koseki, N. Matsunaga, K.A. Nguyen, S. Su, T.L. Windus, M. Dupuis, J.A.
Montgomery, General atomic and molecular electronic structure system, J.
Comp. Chem. 14 (1993) 1347.
[16] M.D. Hanwell, D.E. Curtiss, D.C. Lonie, T. Vandermeersch, E. Zurek, G.R.
Hutchison, Avogadro: an advanced semantic chemical editor, visualization
and analysis platform, J. Cheminform. (2012).
[17] L.A. Curtiss, K. Raghavachari, P.C. Redfern, A.G. Baboul, J.A. Pople, Gaussian-3
theory using coupled cluster energies, Chem. Phys. Lett. 314 (1999) 101107.
[18] L.A. Curtiss, K. Raghavachari, P.C. Redfern, V. Rassolov, J.A. Pople, Gaussian-3
(G3) theory for molecules containing rst and second-row atoms, J. Chem.
Phys. 109 (1998) 77647776.
[19] L.A. Curtiss, P.C. Redfern, K. Raghavachari, V. Rassolov, J.A. Pople, Gaussian-3
theory using reduced Moller-Plesset order, J. Chem. Phys. 110 (1999)
47034709.
[20] A.V. Marenich, C.J. Cramer, D.G. Truhlar, Universal solvation model based on
solute electron density and on a continuum model of the solvent dened by

S. Karwa et al. / Catalysis Today 263 (2016) 1621


the bulk dielectric constant and atomic surface tensions, J. Phys. Chem. B 113
(2009) 63786396.
[21] D.P. Langlois, H.W. Wolff, Pseudo esters of levulinic acid, J. Am. Chem. Soc. 70
(1948) 26242626.
[22] M. Mascal, S. Dutta, I. Gandarias, Hydrodeoxygenation of the angelica lactone
dimer, a cellulose-based feedstock: simple, high-yield synthesis of branched

C7 -C10 gasoline-like hydrocarbons, Angew. Chem. Int. Ed. Engl. 53 (2014)


18541857.
[23] M. Zviely, R. Giger, E. Abushkara, A. Kern, H. Sommer, H.-J. Bertram, G.E.
Krammer, C.O. Schmidt, W. Stumpe, P. Werkhoff, Solving quality
problems in several avor aroma chemicals, Perfum. Flavor 27 (30) (2002)
3238.

21

Potrebbero piacerti anche