You are on page 1of 9

Fuel 260 (2020) 116376

Contents lists available at ScienceDirect

Fuel
journal homepage: www.elsevier.com/locate/fuel

Full Length Article

Thermodynamic analysis, experimental and kinetic modeling of levulinic T


acid esterification with ethanol at supercritical conditions
Vinícius Kothea, Diego Trevisan Melfib, Kallynca Carvalho dos Santosb, Marcos Lúcio Corazzab,

Luiz Pereira Ramosa,
a
Research Center in Applied Chemistry, Department of Chemistry, Federal University of Paraná, CEP 81531-990 Curitiba, PR, Brazil
b
Department of Chemical Engineering, Federal University of Paraná, CEP 81531-990, Curitiba, PR, Brazil

G R A P H I C A L A B S T R A C T

A R T I C LE I N FO A B S T R A C T

Keywords: Ethyl levulinate is an environmentally friendly biomass-derived ester that is an alternative to the classic pet-
Levulinic acid roleum-derived fuel additives. Several studies have been addressed to its chemical production pathways. The
Supercritical esterification supercritical esterification of levulinic acid to ethyl levulinate, however, remains understudied. This work re-
Ethanol ports the effect of process variables and a kinetic study for the esterification of levulinic acid with ethanol under
Ethyl levulinate
sub and supercritical conditions. Experimental data were obtained in a continuous tubular reactor at a fixed
Tubular reactor
pressure of 100 bar. The reaction temperature varied from 220 to 280 °C, and the ethanol to levulinic acid molar
ratios from (2:1) to (9:1). Ethyl levulinate was synthesized with high selectivity under all evaluated reaction
conditions, achieving conversions up to 80% and 93% when ethanol to levulinic acid molar ratios of (2:1) and
(9:1) were used, respectively. A PFR model approach was considered with an elementary reversible self-cata-
lyzed rate law, and the effect of considering the mixture density behavior through the reactor using the PC-SAFT
equation of state was discussed. The proposed kinetic approach was able to correlate the kinetic experimental
data for all experimental conditions used in this study. Furthermore, a thermodynamic analysis was performed to
elucidate trends in reaction performance.

1. Introduction sustainable alternatives for the production of energy, fuels, chemicals,


and materials. Biomass stands out for this purpose due to its large
The gradual decline in easily accessible petroleum reservoirs and availability, versatility, and low cost [1,2]. In this regard, en-
environmental concerns related to its use have raised the need to find vironmentally friendly biomass-derived esters have been identified as


Corresponding author.
E-mail address: luiz.ramos@ufpr.br (L.P. Ramos).

https://doi.org/10.1016/j.fuel.2019.116376
Received 9 July 2019; Received in revised form 3 October 2019; Accepted 7 October 2019
Available online 12 October 2019
0016-2361/ © 2019 Elsevier Ltd. All rights reserved.
V. Kothe, et al. Fuel 260 (2020) 116376

suitable replacements to petroleum-derived chemicals in various areas An HPLC pump (Shimadzu, model LC-20AT) fed the reactor with a
such as pharmaceuticals, transportation fuels, plasticizers, solvents, controlled flow rate of the reaction mixture. The pressure was adjusted
food flavors, coating, and fragrance [3–6]. using a back-pressure regulator (KPB series, Swagelok). The tempera-
Levulinic acid, also known as 4-oxopentanoic acid, is a renewable ture profile and the pressure inside the reactor were monitored re-
carboxylic acid that can be obtained from lignocellulosic materials such spectively by four K-type thermocouples (1 °C uncertainty) and one
as corn starch, sugarcane bagasse, wheat straw, rice husks, paddy straw, pressure transducer (Huba Control, uncertainty within 0.025 MPa)
sorghum grain, water hyacinth, paper mill sludge, tobacco chops, and connected in a field logger register (Novus). After the steady-state con-
olive tree pruning [7,8]. This acid has been selected by the US De- dition was reached, considering at least twice the spatial time at each
partment of Energy (DOE), with ethanol and other compounds, as part different mass flow, samples of the reaction mixture were collected at
of the 15 carbohydrate-derived chemical building blocks for the de- the reactor outlet.
velopment of biorefineries, owing to its potential to serve as a green Quantification analysis of reaction products was carried out in a gas
platform chemical [9]. chromatograph (Shimadzu GC 2010 Plus) equipped with an AOC 20i
Ethyl levulinate, among the levulinate esters, has gained attention autosampler, a capillary Agilent CP-Wax 58 FFAP CB (50 m × 0.25 mm;
because it can be applied for fragrance manufacturing, as neat fuel and 0.20 µm) column and a flame ionization detector (FID). Samples (1 µL)
as additive for diesel and biodiesel [10–13]. Several studies have been were injected at 200 °C in a 1:20 split ratio. The column temperature
performed so far to elucidate and optimize its chemical production program included one isothermal step at 100 °C for 1 min, one heating
pathways either from raw biomass [14–17], carbohydrates [17–19], or ramp at 25 °C min−1 until 175 °C with an isothermal holding time of
directly from levulinic acid [20–25]. 0.5 min, a second heating ramp at 5 °C min−1 until 230 °C and a final
Chemical routes starting from levulinic acid and ethanol are parti- isothermal holding time of 3 min. Quantification was carried out by
cularly attractive since they combine two environmentally friendly external calibration. Calibration curves were built for levulinic acid
compounds of high interest for biorefineries. Levulinic acid esterifica- (99%, Sigma-Aldrich) and ethyl levulinate (99%, Sigma-Aldrich) using
tion to ethyl levulinate exhibits low reaction rates under auto-catalyzed standard solutions covering the concentration range of 100–1000 ppm.
and low-temperature conditions. Russo et al. [21] reached only 25% of
levulinic acid conversion after 1 h at 60 °C using a (5:1) ethanol to le- 2.2. Design of experiments
vulinic acid molar ratio. Thus, this reaction has been carried out with
homogeneous catalysts such as sulfuric acid [20,21], as well as with A 23 factorial design was performed keeping the inlet volumetric
heterogeneous catalysts such as enzymes [22], ion-exchange resins flow at ambient conditions fixed at 2 mL∙min−1, and using temperature,
[21,23], sulfonated hydrothermal carbons [20], transition metal salts ethanol to levulinic acid molar ratio and pressure as factors. The levels
[24], zeolites, and sulfated oxides [25]. were chosen to include ethanol subcritical and supercritical conditions
Although different catalytic systems have led to high levulinic acid according to the following maximum and minimum values: 220 and
conversions, these routes lead to subsequent complications such as 280 °C, 100 and 200 bar, and ethanol to levulinic acid molar ratios of
consumption, deactivation, separation, and recovery of catalysts, and (1:1) and (9:1). The center point of the factorial design (250 °C, 150 bar
low selectivity in some cases. Bankole and Aurand [26] attempted to and molar ratio of 5:1) was performed in three replicates.
bypass these issues by conducting the auto-catalyzed esterification of
levulinic acid at high temperatures, reaching about 60% conversion at 2.3. Kinetic modeling
250 °C using an ethanol to levulinic acid molar ratio of (1:1). This was
an interesting approach because supercritical ethanol has been suc- Kinetic data for the levulinic acid esterification were acquired under
cessfully used for esterification and transesterification in various reac- ethanol subcritical (P = 100 bar; T = 220 °C) and supercritical condi-
tion systems [27–29]. However, Bankole and Aurand [26] did not tions (P = 100 bar; T = 250 and 280 °C) at two different ethanol to
evaluate the effect of high temperatures and different molar ratios. levulinic acid molar ratios (MR = 2:1 and 9:1). At each reaction con-
Also, because the pressure was not measured in their study, it is not dition, the mass flow (ṁ ) was varied from 0.1 to 4.7 g min−1 and each
possible to infer the thermodynamic conditions of the system and if corresponding outlet amount of ethyl levulinate was measured.
ethanol indeed reached its critical point (pressure above 61.48 bar and The kinetic of levulinic acid esterification with ethanol at high
temperature above 240.75 °C). temperatures and high pressures was assumed to follow an elementary
This work provides consistent experimental data and the mathe- and self-catalyzed rate law, as represented in Eq. (1). Thus, Eq. (2) was
matical modeling of a tubular isobaric and isothermal reactor where the used to calculate the corresponding reaction rate (rA ),
levulinic acid esterification under ethanol supercritical conditions took
k1
place. The information and the proposed modeling approach are fun- A + Et + A⟷E + W + A
damental to understand and provide basic parameters for modeling, k2 (1)
designing and optimization of an alternative process regarding the su-
rA = (−k1 CA CEt + k2 CE CW ) CA (2)
percritical esterification of levulinic acid to ethyl levulinate.
where A, Et , E , and W represent levulinic acid, ethanol, ethyl levuli-
2. Materials and methods nate and water, Ci is the concentration of each “i ” compound, and k1 and
k2 are the forward and backward kinetic rate constants, respectively.
2.1. Experimental section This approach is representing the self-catalytic feature of acid ester-
ification under near and supercritical conditions of ethanol [27].
The esterification of levulinic acid (99%, purchased from Sigma- Therefore, the acid appears in both sides of the equations because it is
Aldrich) and ethanol (99.8%, purchased from Vetec®), without pre- considered to act as reagent and catalyst in this reaction system.
treatment of the reagents, was carried out inside a continuous tubular A steady-state plug flow reactor model was assumed to represent the
reactor as described by Santos et al. [28]. Briefly, the reactor consisted reaction system. Therefore, the dispersion effects were neglected and
of a stainless-steel column (316 1/8″ OD × 0.91 cm, Swagelok) with the differential material balance for each i compound was expressed as
22 ± 1 mL capacity (free volume) that was inserted into an electric shown in Eq. (3),
furnace, whose maximum operating temperature was 420 °C (Sanchis,
dFi
Brazil). Further details of this reaction system can be obtained from the = ri
dV (3)
schematic diagram that has been included in Supplementary Material
(Fig. SM-1). where Fi is the mol flow of each i component and V is the reactor

2
V. Kothe, et al. Fuel 260 (2020) 116376

M N
volume.
Since a single reaction is assumed to occur (see Section 3.2), the OF = ∑ ∑ (XjiExp − X Calc
ji )
2

j i (15)
levulinic acid conversion ( X ) is defined as represented in Eq. (4). The
material balance and the concentrations of individual components were where X jiExp
and X Calc
are the experimental and its corresponding
ji
rewritten on the basis of this variable as shown in Eqs. (5)–(9), calculated conversion (Eq. (13)) for each point i of each experimental
condition j , M is the number of points in each isothermal condition and
FA0 − FA
X= N is the number of the experimental conditions evaluated in this work.
FA0 (4)
Root mean square deviation (RMSD ) was used to compute the
correlation between the model predictions and the experimental data as
dX −rA C (k C C − k2 CE CW )
= = A 1 A Et shown in Eq. (16).
dV FA0 FA0 (5)
M N
∑ j ∑i (X jiExp − X Calc
ji )
2
FA C RMSD = 100
CA = = A0 (1 − X ) pmix M·N (16)
V̇ ρ0mix (6)

FEt C 3. Results and discussion


CEt = = A0 (MR − X ) ρmix
V̇ ρ0mix (7)
3.1. Experimental design as a preliminary assessment
FE C
CE = = A0 Xρmix
V̇ ρ0mix (8) As a preliminary study to investigate the reliability of the proposed
reaction system and to collect some basic information for the devel-
FW C
CW = = A0 Xρmix opment of the kinetic modeling, a 23 factorial design was performed.
V̇ ρ0mix (9) The conditions selected and the resulting levulinic acid conversion are
where CA0 is the levulinic acid inlet concentration, V̇ is the volumetric presented in Table 1, and the statistical analysis are presented in
flow, and ρ0mix is the mixture density in the inlet reactor condition. As Supplementary Material. Temperature and ethanol to levulinic acid
the density of the mixture ( ρmix ) varies through the reactor volume, the molar ratio were identified as the most significant process parameters
treatment given to this variable is further detailed in Section 3.4. Also, influencing levulinic acid conversion to ethyl levulinate, as shown in
the CA0 ratio in Eqs. (6)–(9) is expressed in Eq. (10) in terms of reagent the Pareto plot presented in Supplementary Material (Fig. SM-2). De-
ρ0mix spite slightly significant (p > 0.05) in the statistical analysis of the
molar mass (MMA andMMEt ) and ethanol to levulinic acid molar ratio
reaction performance, pressure was chosen to be fixed at 100 bar for the
(MR ).
kinetic modeling. Besides, the effect of pressure on reaction yield was
CA0 xA0 1 minor compared to other reaction variables.
= =
ρ0mix MM0mix MM0mix (MR + 1)
1 3.2. Selectivity analysis
=
1
MR + 1(MM A +
MR
MR + 1
MMEt (MR + 1) ) (10) In this study, the reaction selectivity was monitored by GC–MS
The solution of the differential Eq. (5) is the conversion profile analysis. This evaluation was needed because levulinic acid has two
along the reactor volume (Vr ). The knowledge of the conversion profile functional groups (ketone and carboxylic acid), and since the reaction
allows estimating the mixture density along the reactor volume. With system is subject to high pressure and temperature, parallel reactions
this, the residence time (tr ) was calculated for each experimental point may take place (i.e., lactonization or cyclization of levulinic acid)
according to Eq. (11), leading to the formation of undesirable products.
The results obtained at every reaction condition of the experimental
V = Vr 1 V = Vr ρmix design of Table 1 suggested the absence of byproducts. The GC–MS run
tr = ∫V =0 V̇
dV = ∫V =0 ṁ
dV
(11) of one representative reaction aliquot is given in Supplementary
where ṁ states for the mass flow. Eq. (12) was obtained by the Material (Fig. SM-3), in which no residual levulinic acid is observed.
differentiation of Eq. (11), Hence, the supercritical esterification of levulinic acid with ethanol was
highly selective for ethyl levulinate, allowing the assumption of a single
V̇ dtr = dV (12) reaction taking place at the experimental conditions used in this study
(Section 2.2).
then a change of variable in Eq. (5), with Eq. (12), led to Eq. (13),
which provides the conversion profile along the reaction residence
Table 1
time.
Experimental conditions used in the 23 factorial design developed for the le-
dX −rA·V̇ vulinic acid esterification with ethanol and the resulting levulinic acid con-
= = (k1 CA CEt − k2 CE CW )(1 − X ) version.
dtr FA0 (13)
Temperature (°C) Ethanol to levulinic Pressure Levulinic acid
An Arrhenius-type Eq. (14) was applied to express the kinetic con- acid molar ratio (bar) conversion (%)
stants for Eqs. (5) and (13).
220 1:1 100 49.4
Ea
⎛− i ⎞ 220 1:1 200 53.1
ki = 10 k 0, i e⎝ RT ⎠ i = 1, 2 (14) 220 9:1 100 41.2
220 9:1 200 43.5
Therefore, the parameters to be adjusted were the pre-exponential 250 5:1 150 72.6
(or frequency) factors (k 0,1 and k 0,2 ) and the activation energies (Ea1and 250 5:1 150 71.1
Ea2 ). 250 5:1 150 71.5
The stochastic Particle Swarm Optimization (PSO) algorithm [30] 280 1:1 100 61.7
280 1:1 200 63.1
was applied to minimize the objective function (15) by pre-estimating
280 9:1 100 78.0
the kinetic parameters, which were optimized afterwards using the 280 9:1 200 83.4
“fminsearch” subroutine from Matlab®,

3
V. Kothe, et al. Fuel 260 (2020) 116376

analysis of the simulated P-T diagrams revealed also that the behavior
at low and high molar ratios (with one critical line and without azeo-
trope) were similar, suggesting that the reaction systems involving
ethanol, levulinic acid, ethyl levulinate and water follows a Class I –
Type I phase behavior according to the Konynenburg and Scott classi-
fication [33]. This was so because such compounds are highly asso-
ciative to one another [34].
As expected, the simulated diagrams of Fig. 1 are predicting that the
impact reaction conversion on the thermodynamic behavior is more
pronounced at lower ethanol to levulinic acid molar ratios. Hence, with
an increase in molar ratio, the system composition becomes less sen-
sible to levulinic acid conversion due to the presence of ethanol in
excess. It can also be noticed that the bubble points are more sensible to
the reaction extension than the dew points and both always require
higher pressure to be achieved at the same temperature in relation to an
increase in reaction conversion. As the esterification took place, the
critical point decreased probably due to the consumption of levulinic
acid, which is the compound with the highest critical temperature and
pressure in the system. Thus, as the reaction proceeds, the system be-
comes more soluble on the account of ethyl ester to ethanol and water
to ethanol interactions. Santos et al. [27] came to similar observations
while modeling the esterification of fatty acids using the same ther-
modynamic approach.

3.4. Mixture density evaluation

As described previously, the mixture density varies through the


reactor volume because, at fixed pressure-temperature conditions, it is a
function of composition and/or reaction conversion. Hence, the Aspen
Plus V8.4 (30.0.033) sensitivity analysis was performed to evaluate
changes in mixture density with changes in reaction conversion (Fig. 2)
using the PC-SAFT EoS to deal with non-idealities of this complex
mixture. For this, the pure component parameters are given in Table 2
and all binary interaction parameters were set equal to zero (kij =0).
Fig. 2 reveals that, at all evaluated conditions, the reaction causes
the system to expand. Also, the density of the reactant mixture was
Fig. 1. Pressure-temperature diagrams simulated on Aspen Plus V8.4
more sensible to conversion at high temperatures, reaching 15% and
(30.0.033) for the synthesis of ethyl levulinate using ethanol to levulinic acid
12% of deviation at 280 °C with ethanol to levulinic acid molar ratios of
molar ratios of (a) (2:1) and (b) (9:1) at different simulated reaction conver-
sions. (9:1) and (2:1), respectively. Thus, despite the assumption that in-
compressible fluid fits well under subcritical conditions (for the eval-
uated conditions implying a maximum of 6% deviation at 220 °C and an
3.3. Phase behavior analysis of the reactant mixture
ethanol to levulinic acid molar ratio of 2:1), rigorous modeling of su-
percritical reactors should consider density variations or at least per-
Depending on its temperature, pressure and composition, systems
form a similar thermodynamic analysis.
containing supercritical fluids may be homogeneous or non-homo-
A 4ª order polynomial equation (such as
geneous [31]. Due to the absence of mass transfer limitations, single
ρmix = a·X 4 + b·X 3 + c·X 2 + d·X + e ) was used to adjust the simulated
phase systems are desirable to maximize the reaction rate. Thus, to
mixture density data presented in Fig. 2 for each temperature and molar
verify if the kinetic study is restricted to the homogeneous region, a
ratio evaluated in this study. This approach was used to include in the
phase behavior analysis was performed by simulating pressure-tem-
kinetic modeling the density dependence as a function of composition
perature diagrams at different reaction conversions (X = 0, 20, 40, 60,
at constant pressure and temperature, thus avoiding an extra compu-
80, and 100%). Fig. 1(a) and (b) depict the results of the simulations for
tational effort in adding a subroutine for the PC-SAFT EoS.
both molar ratios that were used in this study (2:1 and 9:1, respec-
tively). For this thermodynamic analysis, PC-SAFT EoS was used to
3.5. Kinetic study
predict the phase behavior for the levulinic acid esterification with
ethanol, where all binary interaction parameters were set equal to zero
Levulinic acid esterification under sub and supercritical conditions
(kij = 0) following the same approach used in previous studies
of ethanol were acquired in a tubular reactor. A steady-state PFR model
[27,28,32]. All simulations were performed using the Aspen Plus V8.4
approach was considered and an elementary reversible self-catalyzed
(30.0.033), with the PC-SAFT pure compound parameters described in
rate law was assumed.
Table 2.
As discussed in Section 3.4, the mixture density varies through the
Fig. 1 demonstrates that the defined experimental conditions are
reactor volume. Earlier studies have neglected such variations in similar
placed in a homogeneous region, however, some of them are close to
systems by taking the assumption of incompressible flow [27,28]. This
the vapor-liquid region predicted using the PC-SAFT equation, parti-
approach is simpler to implement than the detailed kinetic modeling of
cularly at 280 °C for an ethanol to levulinic acid molar ratio of (9:1).
Section 2.2 and was followed for the acquired data to provide a dis-
Since the two-phase region was estimated and therefore subjected to
cussion about the effects of such theoretical assumption.
prediction errors, there is no certainty in stating that all reactions ex-
The adjusted kinetic parameters for the esterification of levulinic
periments were performed under homogeneous conditions. The
acid, assuming and neglecting mixture density variations, are presented

4
V. Kothe, et al. Fuel 260 (2020) 116376

Table 2
Critical properties and PC-SAFT parameters of pure components.
Parameter ASPEN name Levulinic acid1 [40] Ethyl levulinate2 [40] Ethanol [41] Water [41]

mi PCSFTM 2.0311 6.4558 2.3827 1.0656


εi/ κ (K ) PCSFTU 266.4953 226.9371 198.24 3.0007
σi (Ȧ ) PCSFTV 4.1241 3.0938 3.1771 366.51
κ Ai Bi PCSFAU 0.0171 – 0.032384 0.034868
ε Ai Bi·κ −1 (K ) PCSFAV 4578.3655 0 2653.4 2500.7
Assoc. Aceptor – 1 2 1 1
Assoc. Donor – 1 0 1 1
P SAT AAD% – 1.07% 0.13% 0.99% 1.88%
ν[a]or ρ[b] AAD% – 0.06%[b] 1.18%[b] 0.79%[a] 6.83%[a]

1
Values in brackets provide the reference for the experimental dada;
2
Induced association modeling approach.

in Table 3. It is worth emphasizing that these parameters were obtained


from global estimation, considering all evaluated isotherms for both
ethanol to levulinic acid molar ratios simultaneously. To assure the
accuracy of the reported parameters, a sensitivity analysis was per-
formed by perturbating each parameter at a time (by 5–15%) and re-
optimizing the others. A plot of the minimum obtained RMSD (%) for
each perturbed parameter is presented in Supplementary Material (Fig.
SM-4).
Both methodologies displayed low RMSD values and fitted well the
experimental data. The slightly lower RMSD acquired neglecting the
mixture density variations through the reactor volume suggests that the
system thermodynamic behavior might not have been well predicted by
PC-SAFT EoS or that some neglected aspects such as pressure loss,
temperature variations and dispersion effects may be even more sig-
nificant than density variations. These aspects will be evaluated for
rigorous modeling in future work.
Fig. 3(a) and (b) show, for each isothermal condition at ethanol to
levulinic acid molar ratios of 2:1 and 9:1, respectively, the experimental
levulinic acid conversion and the corresponding calculated profile with
the adjusted parameters assuming compressible flow as described in
Section 2.2.
The error bars at the central isotherm were established from du-
plicated experimental runs and from the previous experience of the
group with the employed experimental apparatus. These errors were
classified as type B according to the “Simple Guide for Evaluating and
Expressing the Uncertainty of NIST Measurement Results” [35].
Results obtained in this work shows that the proposed reaction
approach led to higher reaction rates for the levulinic acid conversion
than other approaches presented in the literature. For example, at
280 °C with an ethanol to levulinic acid molar ratio of (9:1), 85% of
levulinic acid conversion was experimentally acquired after a relatively
short residence time of less than 25 min (Fig. 3a). Table 4 summarizes
some reaction conditions that already have been reported for levulinic
acid esterification, as well as the needed residence time to reach similar
reaction conversions.
Fig. 3 also depicts that increases in the feed (ethanol and levulinic
Fig. 2. PC-SAFT EoS simulation on Aspen Plus V8.4 (30.0.033) of the mixture
acid at different molar ratios) and temperature generally led to higher
density at 100 bar in relation to levulinic acid conversion using ethanol to le-
vulinic acid molar ratios of (a) (2:1) and (b) (9:1).
conversions and initial reaction rates, which can be better visualized by
the kinetic model adjusted to the set of experimental data obtained in
this study. Higher initial rates were obtained for reactions performed at
lower molar ratios of 2:1 (CA0 = 1.05 ± 0.15 mmol.L−1) (Fig. 3a)
compared those of higher molar rations of 9:1

Table 3
Adjusted kinetic parameters for levulinic acid esterification with ethanol at supercritical conditions.
Compressible Flow 10 k 0,1 (L∙mol−1)2min−1 10 k 0,2 (L∙mol−1)2min−1 Ea1∙R−1 (K−1) Ea2∙R−1 (K−1) RMSD (%)

Neglected 62866.89 4935.06 8061.45 7230.15 4.82


Assumed 155665.37 8081.87 8494.82 7466.63 5.12

5
V. Kothe, et al. Fuel 260 (2020) 116376

Table 5
More accurate values for gaseous enthalpy and free Gibbs energy of formation
of levulinic acid, ethyl levulinate, ethanol and water.
Levulinic acid Ethyl levulinate Ethanol Water

−491695* −443200 −167285 −228510


ΔG fo ( )
kJ
kmol
Reference [37] [This work] [38] [38]
−613100 −620065 −234531 −241818
ΔH fo ( )
kJ
kmol
Reference [37] [39] [38] [38]

* Calculated directly from the reported formation enthalpy and entropy


using the Gibbs free energy definition.

molar ratio than the other chemical pathways mentioned in Table 4.


However, it presents significant advantages for high levulinic acid
conversions such as lower residence times and no need of adding an
exogenous catalyst. Also, this work serves as a reference material for the
development of scaled-up reactor projects and optimization within the
applied range of temperature and ethanol to levulinic acid molar ratios.

3.6. Chemical equilibrium analysis

In order to perform the chemical equilibrium analysis for the le-


vulinic acid esterification with ethanol, the equilibrium constant (Keq)
was obtained by Eqs. (17) and (18),
ΔGR = ΔHR − T ΔSR = −RT ln(KEQ ) (17)

k ΔHR ΔSR
ln(KEQ ) = ln ⎛ 1 ⎞ = −
⎜ ⎟ +
k
⎝ ⎠
2 RT R (18)
where changes in the reaction enthalpy and entropy were obtained
by fitting Keq versus T−1. Keq values were obtained by the forward and
backward reaction rate constants (k1 and k2), which were assessed from
the adjusted pre-exponential factors and the activation energies pre-
sented in Table 3.
Fig. 4 shows a plot for the Eq. (18) where the obtained parameters
Fig. 3. Experimental and calculated conversion profiles for levulinic acid es-
were ΔHR = -8548.37 J∙mol−1 and ΔSR = 24.59 J.mol−1∙K−1. Two
terification under sub and supercritical conditions of ethanol, using ethanol to
different approaches were considered to calculate the levulinic acid
levulinic acid molar ratios of (a) (2:1) and (b) (9:1). The initial concentration of
levulinic acid at each reaction condition was calculated according to Eq. (10). equilibrium conversion under the applied experimental conditions: the
kinetic and the thermodynamic approaches. The kinetic approach uses
a simulation of the proposed modeling with adjusted parameters to
(CA0 = 3.45 ± 0.23 mmol.L−1) (Fig. 3a). Also, at lower molar ratios identify the levulinic acid conversion that corresponds to the infinity
the reaction reached equilibrium conversion values lower than those residence time (equilibrium conversion) at each reaction condition. By
obtained at 9:1. However, conversions were statistically identical at contrast, the thermodynamic approach consisted in simulations of the
250 °C and 280 °C after ~ 70 min residence time for both molar ratios. Aspen Plus V8.4 (30.0.033) “Equilibrium reactor” block, with the PC-
Therefore, from a process engineering point of view, reactions at SAFT EoS parameters presented in Table 2 and additional pure com-
250 °C, even at lower molar ratios of 2:1, might be more feasible than pound parameters retrieved from Aspen Plus Databank at each reaction
reactions at 280 °C because lower temperatures and ethanol to levulinic condition. This block calculates the standard enthalpy, entropy and
acid molar ratios demand lower energy consumption in the reactor Gibbs free energy of formation of pure compounds and correct them to
setup and in the downstream ethanol recovery. In addition, as pre- the reaction conditions. Therefore, it estimates the equilibrium constant
viously pointed out in Section 3.3, the most severe reaction condition as well as the equilibrium conversion without any kinetic information.
(molar ratio of 9:1 at 280 °C) approaches the simulated vapor-liquid Considering the parameters retrieved from Aspen Plus Databank,
envelop (Fig. 1). With this, mass transfer limitations may have been the two approaches conducted to different results, with the kinetic
present if the reaction was not completely homogeneous. approach always leading to lower equilibrium conversions (Table 6). As
In general, the supercritical esterification of levulinic acid with discussed by Voll et al. [36], a small variation in parameters such as the
ethanol required higher temperature and ethanol to levulinic acid standard Gibbs free energy of formation and the standard formation

Table 4
Reaction conditions reported to reach about 85% levulinic acid conversion.
Esterification Experimental conditions Residence time (min) Conversion (%)

−3
Homogeneously catalyzed by sulfuric acid [21] CCAT = 80 mol·m ; T = 70 °C; MR = 5 250 about 85
Homogeneously catalyzed by Novozym 435 [22] CCAT = 10 mg·mL−1; T = 37.5 °C; MR = 4 187.5 84.7
Heterogeneously catalyzed by FE2(SO4)3 [24] CCAT = 108 mol·m−3; T = 50 °C; MR = 3 50 about 84
Uncatalyzed esterification [This work] T = 280 °C; MR = 9 25 85.1

CCAT and MR refer to the catalyst concentration and ethanol to levulinic acid molar ratio, respectively.

6
V. Kothe, et al. Fuel 260 (2020) 116376

performed a new simulation considering the reaction condition de-


scribed by Bankole and Aurand [26] (uncatalyzed esterification, 1:1
ethanol to levulinic acid molar ratio and 250 °C), which is similar to the
reaction conditions evaluated in this study. Since these authors did not
measure the pressure on their reaction system, it was assumed to be
35 bar (slightly above predicted bubble pressure on the equilibrium
condition), which is coherent with their described experimental appa-
ratus. Under such experimental conditions (250 °C, 35 bar, 1:1 molar
ratio), the thermodynamic approach for levulinic acid equilibrium
conversion predicted 62.72%. This value agrees well with the results of
Bankole and Aurand [26], which corresponded to a 60.8% levulinic
acid conversion after 180 min of reaction time.

4. Conclusion

In this work, the levulinic acid esterification under sub and super-
Fig. 4. Linearized plotting of the equilibrium constant natural logarithm as a critical ethanol conditions was investigated in a continuous tubular
function of the inverse of the temperature for levulinic acid esterification under reactor. A PFR model approach with an elementary reversible auto-
ethanol sub and supercritical conditions. catalytic rate law was proposed. The mixture density profile through
the reactor volume was estimated using the PC-SAFT equation of state.
enthalpy of the acid and the ester can lead to huge variations in the The results found suggested that as the reaction proceeds the reactional
esterification equilibrium conversion. To illustrate this effect a Tornado mix expands, at all evaluated conditions. The proposed kinetic model
plot is presented in Fig. 5 for the experimental conditions (a) and (d) as showed good agreement with the experimentally measured values of
specified in Table 5, considering a deviation of 1.50% for each reaction levulinic acid conversion (RMSD = 5.12%). The reaction was highly
parameter. selective for ethyl levulinate regardless of the applied experimental
Fig. 5 shows that small variations in the parameters of formation conditions, and levulinic acid conversions of up to 80% were reached
significantly affect conversion. Therefore, due to its high sensibility to after a fair residence time (around 15 min at 280 °C using an ethanol to
several parameters, the thermodynamic approach should not be trusted levulinic acid molar ratio of 9:1), showing that the supercritical ester-
to predict equilibrium conversion unless the fundamental thermo- ification of levulinic acid may be a promising chemical pathway to
dynamic parameters are known with great confidence for all involved produce ethyl levulinate.
compounds. In any other situation, the kinetic approach is found to be
more reliable. Declaration of Competing Interest
The standard formation enthalpy and entropy of gaseous levulinic
acid were reported by Reichert et al. [37] with good precision, the The authors declare that they have no known competing financial
standard formation Gibbs free energy and enthalpy are well known for interests or personal relationships that could have appeared to influ-
water and ethanol [38] and there are some studies about the enthalpy ence the work reported in this paper.
of formation of ethyl levulinate due to its use as fuel [39]. However, to
the best of our knowledge, there is still a lack of information about the Acknowledgments
Gibbs free energy of formation of ethyl levulinate.
The Gibbs free energy of formation of ethyl levulinate was fitted in The authors want to thank CNPq (Brazil, grants 406737/2013-4,
order to provide small deviation between the equilibrium conversion 305393/2016-2 and 309506/2017-4), COPEL (Companhia Paranaense
acquired through the kinetic and the thermodynamic approaches. These de Energia, grant PD 2866-0470/2017) and Fundação Araucária (PI 07/
new simulations were performed using Aspen Plus V8.4 (30.0.033) 2018 Horizon 2020, grant 004/2019) for financial support and scho-
“Equilibrium reactor” block with the PC-SAFT EoS parameters pre- larships. This work was also financed in part by the Coordenação de
sented in Table 2, the more accurate values for enthalpy and free Gibbs Aperfeiçoamento de Pessoal de Nível Superior – Brazil (CAPES) –
energy of formation (Table 5) and additional pure compound para- Finance Code 001.
meters retrieved from Aspen Plus Databank, at each evaluated reaction
condition. The results are summarized in Table 6. Appendix A. Supplementary data
Lastly, in order to validate the estimated Gibbs free energy of ethyl
levulinate and the thermodynamic approach as a whole, it was Supplementary data to this article can be found online at https://

Table 6
Equilibrium conversions obtained from the proposed kinetic modeling and Aspen Plus V8.4 (30.0.033) “Equilibrium reactor” block, with PC-SAFT EoS parameters
presented in Tables 2, 5 and additional pure compound parameters retrieved from the Aspen Plus Databank at each reaction condition.
Reaction condition Ethanol to levulinic acid molar ratio Temperature (°C) Equilibrium conversion (%)

Kinetic Thermodynamic* Thermodynamic**

a 2:1 220 78.69 97.01 78.74


b 250 80.14 97.19 80.40
c 280 81.39 97.46 82.41
d 9:1 220 95.29 99.29 93.36
e 250 95.77 99.38 94.52
f 280 96.17 99.54 96.09

* Simulation performed with the more accurate enthalpy and Gibbs free energy of formation.
** Simulation performed with enthalpy and Gibbs free energy of formation retrieved from Aspen Plus Databank.

7
V. Kothe, et al. Fuel 260 (2020) 116376

Fig. 5. Tornado plot sensitive analysis for the levulinic acid conversion response under 1.50% independent perturbations on standard formation enthalpy and
entropy of levulinic acid and ethyl levulinate for experimental conditions (a) and (d) according to Table 5.

doi.org/10.1016/j.fuel.2019.116376. org/10.1039/c3gc41492b.
[13] Leal Silva JF, Grekin R, Mariano AP, Maciel Filho R. Making levulinic acid and ethyl
levulinate economically viable: a worldwide technoeconomic and environmental
References assessment of possible routes. Energy Technol 2018;6:613–39. https://doi.org/10.
1002/ente.201700594.
[1] Carole TM, Pellegrino J, Paster MD. Opportunities in the industrial biobased pro- [14] Olson ES, Kjelden RK, Schlag AJ, Sharma RK. Levulinate esters from biomass
ducts industry. Appl Biochem Biotechnol 2004;113:871–85. wastes. Chem Mater Renew Resour 2001;784:51–63. https://doi.org/10.1021/bk-
[2] Luo Y, Li Z, Li X, Liu X, Fan J, Clark JH, et al. The production of furfural directly 2001-0784.ch005.
from hemicellulose in lignocellulosic biomass: a review. Catal Today [15] Le Van Mao R, Zhao Q, Dima G, Petraccone D. New process for the acid-catalyzed
2019;319:14–24. https://doi.org/10.1016/j.cattod.2018.06.042. conversion of cellulosic biomass (AC3B) into alkyl levulinates and other esters using
[3] Datta R, Henry M. Lactic acid: recent advances in products, processes and tech- a unique one-pot system of reaction and product extraction. Catal Lett
nologies – a review. J Chem Technol Biotechnol 2006;81:1119–29. https://doi.org/ 2011;141:271–6. https://doi.org/10.1007/s10562-010-0493-y.
10.1002/jctb. [16] Chang C, Xu G, Jiang X. Production of ethyl levulinate by direct conversion of
[4] Bart HJ, Reidetschläger J, Schatka K, Lehmann A. Kinetics of esterification of le- wheat straw in ethanol media. Bioresour Technol 2012;121:93–9. https://doi.org/
vulinic acid with n-butanol by homogeneous catalysis. Ind Eng Chem Res 10.1016/j.biortech.2012.06.105.
1994;33:21–5. https://doi.org/10.1021/ie00025a004. [17] Mascal M, Nikitin EB. Comment on processes for the direct conversion of cellulose
[5] Kang S, Fu J, Zhang G. From lignocellulosic biomass to levulinic acid: a review on or cellulosic biomass into levulinate esters. ChemSusChem 2010;3:1349–51.
acid-catalyzed hydrolysis. Renew Sustain Energy Rev 2018;94:340–62. https://doi. https://doi.org/10.1002/cssc.201000326.
org/10.1016/j.rser.2018.06.016. [18] Garves K. Acid catalyzed degradation of cellulose in alcohols. J Wood Chem
[6] Eş I, Mousavi Khaneghah A, Barba FJ, Saraiva JA, Sant’Ana AS, Hashemi SMB. Technol 1988;8:121–34. https://doi.org/10.1080/02773818808070674.
Recent advancements in lactic acid production – a review. Food Res Int [19] Zhu W, Chang C, Ma C, Du F. Kinetics of glucose ethanolysis catalyzed by extremely
2018;107:763–70. https://doi.org/10.1016/j.foodres.2018.01.001. low sulfuric acid in ethanol medium. Chinese J Chem Eng 2014;22:238–42. https://
[7] Galletti AMR, Antonetti C, De Luise V, Licursi D, o Di Nasso NN. Levulinic acid doi.org/10.1016/S1004-9541(14)60049-5.
production from waste biomass. BioResources 2012;7:1824–35. https://doi.org/10. [20] Pileidis FD, Tabassum M, Coutts S, Titirici M. Esterification of levulinic acid into
1007/s40664-017-0218-9. ethyl levulinate catalysed by sulfonated hydrothermal carbons. Chinese J Catal
[8] Morone A, Apte M, Pandey RA. Levulinic acid production from renewable waste 2014;35:929–36. https://doi.org/10.1016/S1872-2067(14)60125-X.
resources: bottlenecks, potential remedies, advancements and applications. Renew [21] Russo V, Hrobar V, Mäki-Arvela P, Eränen K, Sandelin F, Di Serio M, et al. Kinetics
Sustain Energy Rev 2015;51:548–65. doi: 10.1016Zj/rser.2015.06.032. and modelling of levulinic acid esterification in batch and continuous reactors. Top
[9] Bozell JJ, Petersen GR. Technology development for the production of biobased Catal 2018;61:1856–65. https://doi.org/10.1007/s11244-018-0998-y.
products from biorefinery carbohydrates – the US Department of Energy’s “top 10” [22] Lee A, Chaibakhsh N, Rahman MBA, Basri M, Tejo BA. Optimized enzymatic
revisited. Green Chem 2010;12:539–54. https://doi.org/10.1039/b922014c. synthesis of levulinate ester in solvent-free system. Ind Crops Prod 2010;32:246–51.
[10] Hayes DJ. An examination of biorefining processes, catalysts and challenges. Catal https://doi.org/10.1016/j.indcrop.2010.04.022.
Today 2009;145:138–51. https://doi.org/10.1016/j.cattod.2008.04.017. [23] Ramli NAS, Zaharudin NH, Amin NAS. Esterification of renewable levulic acid to
[11] Joshi H, Moser BR, Toler J, Smith WF, Walker T. Ethyl levulinate: a potential bio- levulinate esters using amberlyst-15 as a solid acid. Catalyst 2017;79:137–42.
based diluent for biodiesel which improves cold flow properties. Biomass Bioenergy [24] Martins FP, Rodrigues FA, Silva MJ. Fe2(SO4)3-catalyzed levulinic acid esterifica-
2011;35:3262–6. https://doi.org/10.1016/j.biombioe.2011.04.020. tion: production of fuel bioadditives. Energies 2018:11. https://doi.org/10.3390/
[12] Climent MJ, Corma A, Iborra S. Conversion of biomass platform molecules into fuel en11051263.
additives and liquid hydrocarbon fuels. Green Chem 2014;16:516–47. https://doi. [25] Fernandes DR, Rocha AS, Mai EF, Mota CJA, Teixeira da Silva V. Levulinic acid

8
V. Kothe, et al. Fuel 260 (2020) 116376

esterification with ethanol to ethyl levulinate production over solid acid catalysts. Waals mixtures. Philos Trans R Soc A Math Phys Eng Sci 1980;298:495–540.
Appl Catal A Gen 2012;425–426:199–204. https://doi.org/10.1016/j.apcata.2012. https://doi.org/10.1098/rsta.1980.0266.
03.020. [34] Chapman WG, Gubbins KE, Jackson G, Radosd M. New reference equation of state
[26] Bankole KS, Aurand GA. Kinetic and thermodynamic parameters for uncatalyzed for associating liquids. Ind Eng Chem Res 1990;29:1709–21.
esterification of carboxylic acid. Res J Appl Sci Eng Technol 2014;7:4671–84. [35] Taylor BNN, Kuyatt CEE. NIST Guidelines for Evaluating and Expressing the
https://doi.org/10.19026/rjaset.7.850. Uncertainty of NIST Measurement Results Cover. NIST; 1994.
[27] dos Santos PRS, Voll FAP, Ramos LP, Corazza ML. Esterification of fatty acids with [36] Voll FAP, da Silva C, Rossi CCRS, Guirardello R, de Castilhos F, Oliveira JV, et al.
supercritical ethanol in a continuous tubular reactor. J Supercrit Fluids Thermodynamic analysis of fatty acid esterification for fatty acid alkyl esters pro-
2017;126:25–36. https://doi.org/10.1016/j.supflu.2017.03.002. duction. Biomass Bioenergy 2011;35:781–8. https://doi.org/10.1016/j.biombioe.
[28] dos Santos KC, Hamerski F, Pedersen Voll FA, Corazza ML. Experimental and kinetic 2010.10.035.
modeling of acid oil (trans)esterification in supercritical ethanol. Fuel [37] Reichert D, Montoya A, Liang X, Bockhorn H, Haynes BS. Conformational and
2018;224:489–98. https://doi.org/10.1016/j.fuel.2018.03.102. thermodynamic properties of gaseous levulinic acid. J Phys Chem A
[29] de Jesus AA, de Santana Souza DF, de Oliveira JA, de Deus MS, da Silva MG, 2010;114:12323–9. https://doi.org/10.1021/jp107560u.
Franceschi E, et al. Mathematical modeling and experimental esterification at su- [38] NIST. NIST Standard Reference Database Number 69. NIST Chem Webb 2016. doi:
percritical conditions for biodiesel production in a tubular reactor. Energy Convers 10.18434/T4D303.
Manag 2018;171:1697–703. https://doi.org/10.1016/j.enconman.2018.06.108. [39] Ghosh MK, Howard MS, Dooley S. Accurate and standard thermochemistry for
[30] Ferrari JC, Nagatani G, Corazza FC, Oliveira JV, Corazza ML. Application of sto- oxygenated hydrocarbons: a case study of ethyl levulinate. Proc Combust Inst
chastic algorithms for parameter estimation in the liquid-liquid phase equilibrium 2019;37:337–46. https://doi.org/10.1016/j.proci.2018.07.028.
modeling. Fluid Phase Equilib 2009;280:110–9. https://doi.org/10.1016/j.fluid. [40] Altuntepe E, Emel’yanenko VN, Forster-Rotgers M, Sadowski G, Verevkin SP, Held
2009.03.015. C. Thermodynamics of enzyme-catalyzed esterifications: II. Levulinic acid ester-
[31] Subramanlamr B, McHugh MA. Reactions in supercritical fluids-a review. Ind Eng ification with short-chain alcohols. Appl Microbiol Biotechnol 2017;101:7509–21.
Chem Process Des Dev 1986;25:1–12. https://doi.org/10.1021/i200032a001. https://doi.org/10.1007/s00253-017-8481-4.
[32] Carvalho dos Santos K, Pedersen Voll FA, Corazza ML. Thermodynamic analysis of [41] Gross J, Sadowski G. Application of the perturbed-chain SAFT equation of state to
biodiesel production systems at supercritical conditions. Fluid Phase Equilib associating systems. Ind Eng Chem Res 2002;41:5510–5. https://doi.org/10.1021/
2019;484:106–13. https://doi.org/10.1016/j.fluid.2018.11.029. ie010954d.
[33] Konynenburg PHV, Scott RL. Critical lines and phase equilibria in binary Van der

You might also like