General Kinetic Modeling of the Selective ... - ACS Publications

Nov 2, 2015 - MBY over a modified 5 wt % Pd/CaCO3 catalyst at 333 K and. 2.8 bar, proposing rate equations based on a Langmuir−. Hinshelwood-type ra...
0 downloads 0 Views 670KB Size
Article pubs.acs.org/IECR

General Kinetic Modeling of the Selective Hydrogenation of 2‑Methyl-3-butyn-2-ol over a Commercial Palladium-Based Catalyst Sergio Vernuccio* and Philipp Rudolf von Rohr Institute of Process Engineering, ETH Zurich, Sonneggstrasse 3, 8092 Zurich, Switzerland

Jonathan Medlock Research & Development, DSM Nutritional Products, 4002 Basel, Switzerland S Supporting Information *

ABSTRACT: A general kinetic model has been developed to simulate the three-phase solvent-free hydrogenation of 2-methyl-3butyn-2-ol (MBY) over a commercial palladium-based catalyst. A Langmuir−Hinshelwood mechanism with noncompetitive adsorption between hydrogen and the organic species is assumed. Gas−liquid mass-transfer resistance is included in the model. Experiments were carried out in a stirred slurry reactor to estimate the kinetic parameters. The proposed model is able to predict the concentration profiles of the species involved during MBY hydrogenation, at varying temperatures (313−353 K), pressures (3.0−10.0 bar), and catalyst loadings (0.075−0.175 wt %). The model predictions were successfully validated using additional experimental runs conducted under different operating conditions and at a lower initial concentration of MBY.

1. INTRODUCTION The selective semihydrogenation of 2-methyl-3-butyn-2-ol (MBY) is an important reaction in the production of fine chemicals.1 It is a classic example of a catalytic three-phase process following a consecutive reaction network. The selectivity of this reaction system concerns the possibility of hydrogenating MBY to 2-methyl-3-buten-2-ol (MBE) while preventing further hydrogenation to 2-methyl-2-butanol (MBA), as depicted in Scheme 1.

2.8 bar, proposing rate equations based on a Langmuir− Hinshelwood-type rate mechanism with competitive adsorption of the reactants/products on the catalyst surface. The aforementioned model assumes a first-order reaction with respect to hydrogen and is simplified by the use of estimated apparent kinetic parameters, lumping together kinetic and adsorption constants and hydrogen concentrations. The same mechanism was used by Crespo-Quesada et al.3 to describe the hydrogenation of MBY over a Pd/ZnO structured catalyst. In that work, an estimate of the activation energy for the main hydrogenation (of MBY to MBE) was proposed according to a simplification of the reaction rate expression in a power-law rate. A different mechanism was developed by Rebrov et al.7 to describe the kinetics of MBY hydrogenation over Pd-based catalysts at 333 K and 5.0 bar. The proposed reaction rate equations suggest that hydrogen reacts with the liquid phase directly from the gas phase, without adsorption on the catalyst surface, resulting in a typical Eley−Rideal approach. Despite the extensive use of the hydrogenation of MBY, a detailed characterization of the kinetic parameters governing this process has not been proposed in the literature. Thus, all of the preceding results are related to specific temperatures and pressures and have not been demonstrated to predict the evolution of the system under a wide range of operating conditions. The present work aims at developing and validating a general kinetic model to simulate the system behavior at varying temperatures (313−353 K), pressures (3.0−10.0 bar), and

Scheme 1. Reaction Network of MBY Hydrogenation

The partially hydrogenated product MBE has great importance in the industrial synthesis of vitamins A and E and is also a key intermediate in the manufacture of aroma compounds.1,2 The selective hydrogenations of alkynols are usually carried out in the liquid phase in batch processes in the presence of supported Pd-based catalysts.3 A commonly used catalyst for the hydrogenation of alkynes is the Lindlar catalyst, a 5 wt % Pd/CaCO3 powder, modified by the addition of lead to improve alkene selectivity.4,5 Because of its industrial relevance, the hydrogenation of MBY has been investigated by several researchers. 3,6,7 Bruehwiler et al.6 studied the kinetics of the hydrogenation of MBY over a modified 5 wt % Pd/CaCO3 catalyst at 333 K and © 2015 American Chemical Society

Received: Revised: Accepted: Published: 11543

September 14, 2015 October 29, 2015 November 2, 2015 November 2, 2015 DOI: 10.1021/acs.iecr.5b03424 Ind. Eng. Chem. Res. 2015, 54, 11543−11551

Article

Industrial & Engineering Chemistry Research

For each analysis, a sample of approximately 1.5 mL of liquid was withdrawn so that the total uptake for each experiment never exceeded 10% of the initial liquid volume. No significant differences were observed between test experiments conducted by sampling 10% and 5% of the initial liquid quantity. Hence, the liquid-to-catalyst ratio was considered to remain constant during the sampling. The densities of the reacting solutions during the experimental runs were assumed to be constant and equal to that of MBY. The average molecular weight of the dimers formed during the reactions was assumed to be 160 g· mol−1.3 Hydrogenation tests under two different sets of experimental conditions were repeated three times to ensure their reproducibility. The results, in terms of species concentration, were found to be reproducible to within 7%.

catalyst loadings (0.075−0.175 wt %). The experimental runs were designed to investigate the typical operating ranges of industrial reactors. A kinetic model based on noncompetitive adsorption between hydrogen and organics has been successfully applied for the first time to the hydrogenation of MBY. The final purpose was to provide reliable kinetic expressions and a detailed set of adsorption and kinetic parameters (activation energies, enthalpies of adsorption, and pre-exponential factors) for each of the reactions involved in the process. The proposed model, reliably validated, is able to predict the kinetic behavior of the system given its temperature and pressure and the amount of catalyst added.

2. EXPERIMENTAL SECTION 2.1. Materials. 2-Methyl-3-butyn-2-ol (purum, > 98%) was supplied by DSM Nutritional Products. 2-Methyl-3-buten-2-ol (purum, ≥ 98%) and 2-methyl-2-butanol (purum, ≥ 99%) were purchased from Sigma-Aldrich. Pure hydrogen (99.995%) and nitrogen (99.995%), for inertization purposes, were supplied by Pangas. The hydrogenation experiments were conducted using a commercial Lindlar catalyst (5 wt % Pd/CaCO3 modified with Pb) purchased from Sigma-Aldrich. All of the reagents and the catalyst were used as received. Some of the catalyst characteristics and the physical properties of MBY are listed in the Supporting Information. A fresh sample of catalyst was used for each hydrogenation experiment. 2.2. Experimental Setup and Procedure. All experiments were performed with 200 g of the initial mixture charged in a 400 mL batch autoclave (Premex Reactor AG, Lengnau, Switzerland) made of titanium with an inner diameter of 65 mm and a height of 118 mm. The reactor was equipped with a system of four equidistant baffles with a width of 12 mm and a thickness of 2.5 mm. A four-blade gas-entrainment stirrer with a diameter of dI = 38 mm and a width of wI = 12 mm allowed the gas in the head space to be dispersed into the liquid. The rotational speed of the impeller was controlled by means of an electronic frequency controller set at 20 s−1. The reactor was equipped with an electrical heating jacket. Isothermal conditions during the experiments were ensured by a watercooling control system. Before heating, the headspace of the reactor was flushed three times with nitrogen for safety issues.3 In typical experiments for the kinetic study, the reactor was flushed and then pressurized with pure hydrogen to the required level, once the temperature had stabilized at the desired value. During the reaction, the pressure was maintained constant by supplying hydrogen from an external cylinder (all pressure values mentioned in this work are given as absolute). The reaction time was initialized to zero as soon as the liquid phase with suspended catalyst came into contact with hydrogen and the stirrer was switched on. Hydrogenation experiments were carried out in the temperature range of 333−353 K and the pressure range of 3.0−10.0 bar. Liquid samples were withdrawn at defined intervals of time through a valve connected to the reactor and analyzed using a Bruker GC-450 gas chromatograph with a flame ionization detector (FID). The gas chromatograph was equipped with a VF-Wax ms separation column (25 m × 0.25 mm, coating thickness = 0.25 μm). The injector and flame ionization detector temperatures were 523 and 493 K, respectively. The oven temperature was increased at 20 K·min−1 to 473 K.

3. RESULTS AND DISCUSSION 3.1. Gas−Liquid Mass Transfer. The volumetric gas− liquid (G−L) mass-transfer coefficients kLa for hydrogen in MBY were estimated in the absence of the catalyst by the gas absorption method.8,9 Some of the experiments were conducted under inert atmosphere (N2) in the presence and absence of the catalyst because, as reported in the literature, kLa can be affected by the presence of solid particles.8,10 The values for kLa were assumed not to change with the concentration of MBY during the reaction because of the similar physical properties of the species involved in the system. Table 1 reports the obtained results; it is concluded that low concentrations of catalyst do not affect kLa. On the other hand, Table 1. Gas−Liquid Mass-Transfer Coefficients for Hydrogen and Nitrogen in MBY gas

T (K)

catalyst loading (wt %)

H2 H2 H2 H2 N2 N2 N2 N2

313 333 343 353 353 353 353 353

0 0 0 0 0 0.075 0.125 0.175

kLa (s−1) 0.36 0.43 0.47 0.51 0.32 0.30 0.33 0.34

± ± ± ± ± ± ± ±

0.01 0.02 0.02 0.01 0.02 0.02 0.01 0.02

increasing the temperature considerably increases kLa. At high temperature, the liquid viscosity and surface tension decrease, leading to a decrease in the average bubble size and, accordingly, to an increase in the specific G−L contact area a. Furthermore, the increase in diffusivity at high temperature implies a direct increase of the G−L mass-transfer coefficient k L. It is usually accepted in the literature that the effect of pressure on kLa is small or negligible.11,12 For this reason, kLa was considered in this work to be independent of pressure in the range between 3.0 and 10.0 bar. 3.2. Liquid−Solid Mass Transfer. To estimate the limitations set by external liquid−solid (L−S) mass transfer, the Sherwood−Frössling correlation6,13 was used: Shi = 2 + 0.4Rei1/4Sc1/3

(1)

with Shi = 11544

k S, id p, i DH

(2) DOI: 10.1021/acs.iecr.5b03424 Ind. Eng. Chem. Res. 2015, 54, 11543−11551

Article

Industrial & Engineering Chemistry Research NpdI 5nI 3d p, i 4ρL 3

Rei =

Sc =

VLμL 3

3.3. Hydrogen Solubility. The solubility of hydrogen in MBY was experimentally estimated between 313 and 353 K using the physical absorption method introduced in section 3.1. The effect of pressure on solubility is described by Henry’s law as p C H* = H (8) H where Henry’s constant H was assumed to be independent of the pressure.17 The hydrogen partial pressure pH was obtained as the difference between the absolute pressure of the system and the vapor pressure of MBY reported in Table 2. The temperature dependence of the Henry’s constant is shown in Figure 1. The linear trend suggests that the effect of

(3)

μL ρL DH

(4)

The subscript i refers to the catalyst particle diameter, dp,i, according to the cumulative distribution function of particle sizes presented in the Supporting Information. The power number Np was assumed to be 5.5 for a four-blade stirrer with wI/dI ≈ 0.3.14 All properties of the liquid phase were considered to be equal to those of pure MBY. The molecular diffusivity of hydrogen in the liquid, DH, was estimated from the Wilke−Chang correlation for unassociated solvents, which is usually accurate to within 10−15%15 DH = 7.4 × 10−15

T ML μL VH 0.6

(5)

with the hydrogen molar volume VH equal to 14.3 cm3·mol−1.15 Based on the particle size distribution of the catalyst powder, the L−S mass-transfer coefficients were calculated as n

k SaS =

∑ qikS,iaS,i

(6)

i=1

Assuming that the catalyst particles can be approximated as spherical, the specific L−S interface area was estimated as aS, i =

6 mS ρL d p, i mL ρS

Figure 1. Effect of temperature on the Henry’s constant of hydrogen in MBY.

(7)

The calcium carbonate catalyst support is nonporous,16 and no large aggregates of particles were detected by laser diffraction (maximum diameter of 60 μm). For this reason, the catalyst effectiveness factor was considered to be equal to 1, and internal L−S mass transfer neglected. The results in Tables 1 and 2 show that the calculated L−S mass-transfer coefficients kSaS are substantially higher than the corresponding G−L kLa values. For this reason, the L−S mass-transfer resistance was neglected in this work.

temperature on H should be modeled by an Arrhenius-type equation following the approach proposed by Fillion and Morsi. 18 Thus, the hydrogen solubility in MBY was theoretically expressed as pH C H* = ΔE H0 exp − RT (9)

(

Estimated values for the pre-exponential factor and the apparent activation energy of adsorption are H0 = 33.8 bar·L· mol−1 and ΔE = −6.81 kJ·mol−1. These results were obtained by linear regression of the experimental data depicted in Figure 1 (r2 = 99.3%). At 333 K, the Henry’s constant was estimated to be 396 bar· L·mol−1. This value is close to those reported in the literature at the same temperature for similar systems: 357 bar·L·mol−1 for hydrogen in alkynes19 and 440 bar·L·mol−1 for hydrogen in MBY.6 3.4. Kinetic Model. The reaction network proposed for the hydrogenation of MBY is summarized as follows:

Table 2. Overview of the Hydrogenation Experiments on Pure MBY Used for the Calculation of the Kinetic Parameters with the Estimated Liquid−Solid Mass-Transfer Coefficients run

T (K)

p (bar)

catalyst loading (wt %)

kSaS (s−1)

1 2 3 4 5 6 7 8 9 10 11 12 13 14

333 353 333 333 333 313 333 333 353 353 313 353 333 353

9.0 9.0 9.0 9.0 7.0 9.0 10.0 4.0 10.0 3.0 7.0 9.0 6.0 9.0

0.125 0.125 0.175 0.075 0.125 0.125 0.125 0.125 0.125 0.125 0.125 0.175 0.125 0.075

2.67 4.33 3.74 1.60 2.67 1.48 2.67 2.67 4.33 4.33 1.48 6.33 2.67 2.71

)

r1

MBY + H 2 → MBE r2

MBE + H 2 → MBA r3

2MBY + 2H 2 → dimers

Reaction r3 involves the formation of dimeric C10 products, as a result of the reaction between two molecules of MBY adsorbed on the catalyst surface and hydrogen. 11545

DOI: 10.1021/acs.iecr.5b03424 Ind. Eng. Chem. Res. 2015, 54, 11543−11551

Article

Industrial & Engineering Chemistry Research KH

According to various authors, the formation of the alkane takes place not only in a consecutive manner, but also in parallel by direct hydrogenation of MBY to MBA.3,6,7 However, in our experiments, only negligible amounts of MBA were detected before the complete hydrogenation of MBY, resulting in a high selectivity to MBE in the process. Therefore, to simplify the network, we decided to attribute the consumption of MBY to only its partial hydrogenation to MBE and its dimerization. The same approach was followed by Nikoshvili et al.20 Figure 2 shows the effect of the hydrogen partial pressure on the initial hydrogenation rate of MBY in the range of 313−353

H 2 + 2□ ↔ 2H□ Kφ

MBY + MBY○ ↔ (MBY)2 ○ K1′

MBY○ + H□ ↔ R○ + □ k1′

R○ + H□ → MBE○ + □ KE

MBE + ○ ↔ MBE○

K. These values were obtained from experiments carried out with pure MBY and the same catalyst loading under different operating conditions. The initial rate of MBY hydrogenation appears to vary linearly with hydrogen partial pressure for all three investigated temperatures. This analysis indicates that the rate of reaction is first-order with respect to hydrogen under the investigated range of conditions. One of the relevant catalytic mechanisms for the hydrogenation of MBY over Pd was proposed by Crespo-Quesada et al.,3 assuming competitive adsorption of the reactants on one type of active site. However, in this work, we employed the simplest Langmuir−Hinshelwood model for a catalyzed reaction between large and small molecules, namely, noncompetitive adsorption of hydrogen and organics. This hypothesis, usually accepted in the literature for the hydrogenation of both alkynes21 and alkenes,22 is justified considering that the studied solvent-free system was highly concentrated in unsaturated organic species and the pressure range was limited to 3.0−10.0 bar. For these reasons, a high catalyst surface coverage is expected for MBY, whereas the corresponding surface coverage for hydrogen is lower and assumed not to change dramatically over the pressure range. Dissociative adsorption of hydrogen was assumed to occur on sites represented by open squares (□), different from those involved in the adsorption of organic species, represented by open circles (○). The elementary steps proposed to describe MBY hydrogenation to MBE are listed below. A similar mechanism was assumed by Alves et al.23 for the liquid-phase hydrogenation of 1-butyne over a Pd-based catalyst. KY

(s3) (s4) (s5) (s6)

Steps s1, s2, and s6 describe the absorption/desorption of MBY, hydrogen, and the product MBE, respectively, on the catalyst surface. All of the adsorption and desorption steps were assumed to be fast enough to reach quasi-equilibrium. The complex (MBY)2 originates in step s3 when two molecules of MBY adsorb on one active site. The addition of four hydrogen atoms to this complex (steps not shown) leads to the formation of a dimer. It is accepted that the formation of MBE involves two consecutive elementary steps by the dissociative adsorption of hydrogen.3 A first hydrogen atom is added to the adsorbed molecule of MBY in step s4. The semihydrogenated radical R reacts with an additional hydrogen atom in step s5 to produce an adsorbed molecule of MBE. The same adsorption/desorption and reaction steps (not shown) were assumed to describe the overhydrogenation of MBE to MBA. The observed dependency of the MBY hydrogenation rate on the hydrogen pressure (Figure 2) suggests that the first hydrogen addition (step s4) is quasi-equilibrated and the second (step s5) is nonreversible. The same approach was followed for the production of dimers, considering the first three hydrogen-atom additions reversible. During the hydrogenation of carbon−carbon multiple bonds, the last hydrogen addition is usually considered nonreversible because the resulting molecule desorbs from the surface of the catalyst.24 Under this assumption, the rate equation for the hydrogenation reaction of MBY to MBE can be written as

Figure 2. Effect of hydrogen pressure on the initial rate of MBY hydrogenation. Conditions: 0.125 wt % catalyst.

MBY + ○ ↔ MBY○

(s2)

r1 = k1′ϑR ϑ′H

(10)

The fractional surface coverages ϑR and ϑ′H are obtained from the combination of the two mass balances on the different active sites (eqs 11 and 12) with the expressions of the equilibrium constants of the reversible reaction steps. The fraction of active sites covered with R was neglected in the mass balance ϑY + ϑE + ϑA + ϑD + ϑY2 + ϑ0 = 1

(11)

ϑ′H + ϑ′0 = 1

(12)

The following expression for the reaction rate of MBY hydrogenation was finally deduced r1 = k1 ×

KYC Y 1 + KYC Y(1 + KφC Y ) + KEC E + KACA + KDC D KHC H 1 + KHC H

(13)

where k1 = k′1K′1 is the apparent kinetic constant for the hydrogenation reaction of MBY to MBE. More details about the derivation of the kinetic equations are presented in the Supporting Information.

(s1) 11546

DOI: 10.1021/acs.iecr.5b03424 Ind. Eng. Chem. Res. 2015, 54, 11543−11551

Article

Industrial & Engineering Chemistry Research

The objective function F to be minimized, obtained as the sum of the squares of the errors between the experimental ( f) and calculated (c) concentration data, was expressed as

The concentrations of dimers detected during all of the experiments were very low compared to the concentrations of the other species involved in the system. Furthermore, the adsorption constant of MBA is small, because of the weak adsorption of the single bond on the catalyst surface. For these reasons, the terms KDCD and KACA were neglected in eq 13. The weak hydrogen adsorption on the Pd surface25 and the low hydrogen concentrations allow the hydrogen inhibition contribution to be neglected as well. The hydrogen inhibition term would become important in the case of high pressures leading to higher values of hydrogen solubility, according to eq 9. Under these assumptions, eq 13 can be simplified to r1 = k1

KYC YKHC H 1 + KYC Y(1 + KφC Y ) + KEC E

n3

F=

n2

n1

∑ ∑ ∑ (fi ,j ,k k=1 j=1 i=1

− ci , j , k)2 (24)

in which n1 is the number of experimental points recorded during each run, n2 is the number of detected species, and n3 is the number of the experiments involved in the optimization. The analysis was conducted using 528 data points to estimate 14 parameters. In Figures 3 and 4, the experimental (points) and calculated (lines) concentrations are shown for two runs carried out at the

(14)

For the reaction rates of MBE to MBA and of MBY to dimers, the following equations were similarly obtained r2 = k 2

r3 = k 3

KEC EKHC H 1 + KYC Y(1 + KφC Y ) + KEC E

(15)

KYKφC Y 2KH 2C H 2 1 + KYC Y(1 + KφC Y ) + KEC E

(16)

According to the reaction network, we derived the following set of ordinary differential equations to express the material balances of the species involved in the process n dC Y = Pd ( −r1 − 2r3) dt VL

(17)

n dC E = Pd (r1 − r2) dt VL

(18)

n dCA = Pd r2 dt VL

(19)

n dC D = Pd r3 dt VL

(20)

n dC H = kLa(C H* − C H) − Pd (r1 + r2 + 2r3) dt VL

(21)

Figure 3. Experimental points and calculated kinetic curves. Conditions: 0.175 wt % catalyst, 333 K, 9.0 bar, pure MBY as the initial solution.

The kinetic constants ki were assumed to obey an Arrhenius temperature dependence ⎛ E ⎞ ki = k 0i exp⎜ − i ⎟ ⎝ RT ⎠

(22)

Figure 4. Experimental points and calculated kinetic curves. Conditions: 0.125 wt % catalyst, 353 K, 9.0 bar, pure MBY as the initial solution.

Accordingly, the adsorption equilibrium constants Ki were expressed as ⎛ ΔH ad ⎞ i ⎟⎟ K i = K 0i exp⎜⎜ − ⎝ RT ⎠

same pressure (9.0 bar), two different temperatures (333 and 353 K), and two different catalyst loadings (0.175 and 0.125 wt %). The proposed model appears to have a good capability to simulate the behavior of the system. The total percentage standard deviation for both experiments in Figures 3 and 4 was 3.9%. Similar results in terms of standard deviations were obtained for the data collected under the other studied operating conditions (data not shown). Figures 5−7 focus on the influences of pressure, catalyst loading, and temperature on MBY hydrogenation rate.

(23)

The proposed mathematical model (eqs 9 and 14−23) was solved according to proper initial conditions. The experimental results for the 14 runs listed in Table 2 were simultaneously used in a single optimization procedure to estimate the equilibrium and the kinetic constants (pre-exponential factors, activation energies, and enthalpies of adsorption) governing the process. 11547

DOI: 10.1021/acs.iecr.5b03424 Ind. Eng. Chem. Res. 2015, 54, 11543−11551

Article

Industrial & Engineering Chemistry Research

Figure 7. Influence of temperature on MBY hydrogenation: experimental points and calculated kinetic curves. Conditions: 0.125 wt % catalyst, 9.0 bar, pure MBY as the initial solution.

mechanism. The negative order of reaction with respect to MBY suggests that the active sites of the catalyst remain occupied with MBY even at very low concentrations. For this reason, the adsorption of MBE and the consequent overhydrogenation to MBA are inhibited. The estimated parameters, resulting from the optimization procedure, are reported in Table 3 with their 95% confidence intervals. The uncertainties associated with the estimates, remarkably low, show that the parameters were calculated with a satisfactory precision. At 333 K, the adsorption equilibrium constant of MBY is found to be 0.80 ± 0.08 L·mol−1. This finding is in line with the value of 1.03 L·mol−1 reported by Bruehwiler et al.6 For the range of investigated temperatures, KY appears to be significantly higher than the adsorption equilibrium constant of MBE, KE (KY/KE is between 70 and 100), as reported in the literature for Pd-based catalysts.26 Meanwhile, the apparent kinetic constant of MBE hydrogenation, k2, is higher than the corresponding constant of MBY hydrogenation k1 (k2/k1 is between 1.2 and 2.0). These results agree with the thermodynamics of alkyne hydrogenation. As long as the alkyne is still present in the reacting system, it occupies the catalyst active sites, displacing the alkene. The high selectivity to the olefinic product is due to the stronger adsorption of alkyne compared to alkene and not to the higher kinetic constant.27 The value obtained for the activation energy of MBY hydrogenation, E1, agrees satisfactorily with that reported by Crespo-Quesada et al.3 of 25.2 kJ·mol−1 for Pd/ZnO structured catalyst. Furthermore, the ratio E2/E1 is about 1.4, as proposed in the literature for 1-hexyne hydrogenation over Pd-based catalyst.28 The enthalpies of adsorption of MBY and MBE range from −20 to −30 kJ·mol−1, in fair agreement with the data available in the literature.3,22,28 To validate the model and the estimated parameters, the simulation work was extended to additional experiments not included in the group used for the parametric optimization. It is worth noting that, during this validation procedure, the calculated values of the parameters were used only to predict the experimental results without any additional adjustment. Comparisons between the experimental and calculated concentrations are shown in Figures 8 and 9. Figure 8 shows the results obtained during an additional run conducted under a new combination of operating conditions chosen within the investigated ranges (343 K, 7.0 bar, and catalyst loading of 0.150 wt %). The total percentage standard deviation was 5.5%. In Figure 9 are depicted the results obtained with an initial solution of MBY and MBA (approximately 80 wt % MBY). Similar results were also obtained for an initial solution of MBY and MBE at the same concentration. The good agreement between the experimental and the calculated concentrations (total percentage standard deviation of 2.3%) shows that the proposed model can successfully simulate the behavior of the system even at lower initial concentrations of MBY.

These figures show that the proposed model is able to accurately describe the experimental trends observed when varying the operating conditions in the investigated ranges. It is interesting to note that the depicted results exhibit a slight acceleration of MBY hydrogenation during the reaction. This observation confirms the negative order of hydrogenation with respect to MBY obtained including step s3 in the kinetic

4. CONCLUSIONS A general kinetic model was developed to predict the hydrogenation reaction of MBY over a commercial Pd-based catalyst. Volumetric G−L mass-transfer coefficients k La were estimated in the absence of the catalyst, for different temperatures by gas absorption experiments. The L−S mass-

Figure 5. Influence of pressure on MBY hydrogenation rate: experimental points and calculated kinetic curves. Conditions: 0.125 wt % catalyst, 333 K, pure MBY as the initial solution.

Figure 6. Influence of catalyst loading on MBY hydrogenation: experimental points and calculated kinetic curves. Conditions: 333 K, 9.0 bar, pure MBY as the initial solution.

11548

DOI: 10.1021/acs.iecr.5b03424 Ind. Eng. Chem. Res. 2015, 54, 11543−11551

Article

Industrial & Engineering Chemistry Research Table 3. Estimated Kinetic Parameters with 95% Confidence Intervals parameter

estimate

parameter

estimate

k01 (mol·molPd−1·min−1) k02 (mol·molPd−1·min−1) k03 (mol·molPd−1·min−1) K0Y (L·mol−1) K0E (L·mol−1) K0H (L·mol−1) K0φ (L·mol−1)

(6.18 ± 0.3) × 109 (5.28 ± 0.2) × 1011 (1.72 ± 0.1) × 1010 (3.87 ± 0.4) × 10−5 (6.44 ± 0.3) × 10−6 (1.83 ± 0.1) × 10−2 (3.8 ± 0.2) × 10−2

E1 (kJ·mol−1) E2 (kJ·mol−1) E3 (kJ·mol−1) −1 −ΔHad Y (kJ·mol ) ad −ΔHE (kJ·mol−1) −1 −ΔHad H (kJ·mol ) −1 −ΔHad (kJ·mol ) φ

29.2 ± 1.4 40.2 ± 2.0 29.4 ± 1.5 27.5 ± 1.4 20.2 ± 1.0 4.7 ± 0.2 7.3 ± 0.4

describe the behavior of the reacting system was observed for the full range of experimental conditions. The model was validated using the best estimates of the parameters, without any further adjustment, to simulate the concentration profiles of the species during additional experiments. The validation procedure was also extended to experimental runs conducted at lower initial concentrations of MBY.



ASSOCIATED CONTENT

S Supporting Information *

The Supporting Information is available free of charge on the ACS Publications website at DOI: 10.1021/acs.iecr.5b03424. Physical properties of catalyst and MBY; mathematical derivation of the equations used for kinetic modeling (PDF)

Figure 8. Comparison between experimental points and calculated kinetic curves. Conditions: 0.150 wt %, 343 K, 7.0 bar, pure MBY as the initial solution.



AUTHOR INFORMATION

Corresponding Author

*E-mail: [email protected]. Tel.: +44 44 632 24 99. Fax: +44 44 632 13 25. Notes

The authors declare no competing financial interest.

■ ■

ACKNOWLEDGMENTS We gratefully acknowledge financial support from the Swiss Commission for Technology and Innovation (CTI). NOMENCLATURE

Roman Symbols

a = specific gas−liquid contact area, m−1 aS = specific liquid−solid contact area, m−1 CH* = hydrogen solubility in the liquid at equilibrium, mol· L−1 Ci = concentration of species i, mol·L−1 DH = molecular diffusivity of hydrogen in the liquid, m2·s−1 dI = stirrer diameter, m dp = catalyst particle diameter, m Ei = activation energy of reaction i, kJ·mol−1 F = objective function H = Henry’s constant, bar·L·mol−1 H0 = pre-exponential factor in the expression for the Henry’s constant, bar·L·mol−1 ki = kinetic constant of reaction i, mol·molPd·min−1 Ki = equilibrium adsorption constant of species i, L·mol−1 kL = gas−liquid mass-transfer coefficient, m·s−1 kLa = volumetric gas−liquid mass-transfer coefficient, s−1 kS = liquid−solid mass-transfer coefficient, m·s−1 kSaS = volumetric liquid−solid mass-transfer coefficient, s−1 k0i = pre-exponential factor of reaction i, mol·molPd·min−1

Figure 9. Comparison between experimental points and calculated kinetic curves. Conditions: 0.125 wt %, 353 K, 7.0 bar, initial solution containing 80 wt % MBY and 20 wt % MBA.

transfer coefficients kSaS were theoretically calculated with the Sherwood−Frö ssling correlation and were found to be significantly higher than the corresponding kLa values. For this reason, the L−S mass-transfer resistance was neglected. The hydrogen solubility in MBY was expressed as a function of temperature and hydrogen pressure. The intrinsic kinetics of the reactions involved in the network were based on a Langmuir−Hinshelwood mechanism with noncompetitive adsorption of hydrogen and organics on the catalyst surface. The data for the 14 runs conducted on pure MBY under different experimental conditions (temperatures of 313−353 K, pressures of 3.0−10.0 bar, and catalyst loadings of 0.075−0.175 wt %) were used to estimate the kinetic and adsorption parameters. A good capability of the model to 11549

DOI: 10.1021/acs.iecr.5b03424 Ind. Eng. Chem. Res. 2015, 54, 11543−11551

Article

Industrial & Engineering Chemistry Research

and kinetic studies with in situ hydrogen monitoring. Ind. Eng. Chem. Res. 2008, 47, 6862−6869. (7) Rebrov, V. E.; Klinger, A. E.; Berenguer-Murcia, A.; Sulman, M. E.; Schouten, C. J. Selective hydrogenation of 2-methyl-3-butyn-2-ol in a wall-coated capillary microreactor with Pd25Zn75/TiO2 Catalyst. Org. Process Res. Dev. 2009, 13, 991−998. (8) Dietrich, E.; Mathieu, C.; Delmas, H.; Jenck, J. Raney-nickel catalyzed hydrogenations: gas-liquid mass transfer in gas-induced stirred slurry reactors. Chem. Eng. Sci. 1992, 47, 3597−3604. (9) Deimling, A.; Karandikar, B. M.; Shah, Y. T.; Carr, N. L. Solubility and mass transfer of CO and H2 in Fischer−Tropsch liquids and slurries. Chem. Eng. J. 1984, 29, 127−140. (10) Inga, J.; Morsi, B. I. Effect of catalyst loading on gas/liquid mass transfer in a slurry reactor: a statistical experimental approach. Can. J. Chem. Eng. 1997, 75, 872−881. (11) Oyevaar, M.; Zijl, A.; Westerterp, R. Interfacial areas and gas hold-ups at elevated pressures in mechanically agitated gas-liquid reactors. Chem. Eng. Technol. 1988, 11, 1−10. (12) Gérard, C.; Roizard, C.; Midoux, N.; Laurent, A.; Charpentier, J. C. Influence de la pression sur l’absorption d’un gaz dans un liquid suivie de réaction chimique. Récents Prog. Génie Procédés 3 1989, 8, 26−31. (13) Sano, Y.; Yamaguchi, N.; Adachi, T. Mass transfer coefficients for suspended particles in agitated vessels and bubble columns. J. Chem. Eng. Jpn. 1974, 7, 255−261. (14) Bates, R. L.; Fondy, P. L.; Corpstein, R. R. An examination of some geometric parameters of impeller power. Ind. Eng. Chem. Process Des. Dev. 1963, 2, 310−314. (15) Wilke, C. R.; Chang, P. Correlation of diffusion coefficients in dilute solutions. AIChE J. 1955, 1, 264. (16) Vilé, G.; Almora-Barrios, N.; Mitchell, S.; Lopez, N.; PerezRamirez, J. From the Lindlar catalyst to supported ligand-modified palladium nanoparticles: selectivity patterns and accessibility constraints in the continuous flow three-phase hydrogenation of acetylenic compounds. Chem. - Eur. J. 2014, 20, 5926−5937. (17) Meille, V.; de Bellefon, C.; Schweich, D. Kinetics of αMethylstyrene Hydrogenation on Pd/Al2O3. Ind. Eng. Chem. Res. 2002, 41, 1711. (18) Fillion, B.; Morsi, B. I. Gas-liquid mass transfer and hydrodynamic parameters in a soybean oil hydrogenation process under industrial conditions. Ind. Eng. Chem. Res. 2000, 39, 2157−2168. (19) Meyberg, M.; Roessler, F. In situ measurement of steady-state hydrogen concentrations during a hydrogenation reaction in a gasinducing stirred slurry reactor. Ind. Eng. Chem. Res. 2005, 44, 9705− 9711. (20) Nikoshvili, L. Zh.; Makarova, A. S.; Lyubimova, N. A.; Bykov, A. V.; Sidorov, A. I.; Tyamina, I. Yu.; Matveeva, V. G.; Sulman, E. M. Kinetic study of selective hydrogenation of 2-methyl-3-butyn-2-ol over Pd-containing hypercrosslinked polystyrene. Catal. Today 2015, 256, 231−240. (21) Alves, J. A.; Bressa, S. P.; Martinez, O. M.; Barreto, G. F. Kinetic study of the liquid-phase selective hydrogenation of 1-butyne in presence of 1-butene over a commercial palladium-based catalyst. Chem. Eng. Res. Des. 2011, 89, 384−397. (22) Bressa, S. P.; Martinez, O. M.; Barreto, G. F. Kinetic study of the hydrogenation and hydroisomerization of the n-butenes on a commercial palladium/alumina catalyst. Ind. Eng. Chem. Res. 2003, 42, 2081−2092. (23) Alves, J. A.; Bressa, S. P.; Martinez, O. M.; Barreto, G. F. Kinetic study of the liquid-phase hydrogenation of 1-butyne over a commercial palladium/alumina catalyst. Chem. Eng. J. 2007, 125, 131−138. (24) Smith, G. V.; Notheisz, F. Heterogeneous Catalysis in Organic Chemistry; Academic Press: San Diego, CA, 1999. (25) Crespo-Quesada, M.; Dykeman, R. R.; Laurenczy, G.; Dyson, P. J.; Kiwi-Minsker, L. Supported nitrogen-modified Pd nanoparticles for the selcetive hydrogenation of 1-hexyne. J. Catal. 2011, 279, 66−74. (26) Joannet, E.; Horny, C.; Kiwi-Minsker, L.; Renken, A. Palladium supported on filamentous active carbon as effective catalyst for liquid-

K0i = pre-exponential factor of the equilibrium adsorption constant of species i, L·mol−1 mL = liquid mass, g ML = molecular weight of the liquid, g·mol−1 mS = mass of catalyst particles, g nI = stirrer speed, s−1 Np = power number nPd = number of moles of palladium, mol p = pressure, bar p0 = vapor pressure, bar pH = hydrogen partial pressure, bar qi = volumetric fraction of catalyst particles with diameter dp,i R = universal gas constant, kJ·mol−1·K−1 Re = Reynolds number ri = reaction rate of reaction i, mol·L−1·min−1 Sc = Schmidt number Sh = Sherwood number t = time, min T = temperature, K VH = hydrogen molar volume at the normal boiling point, cm3·mol−1 VL = liquid volume, L or m3 wI = stirrer blade width, m Greek Symbols

ΔE = apparent absorption activation energy, kJ·mol ΔHad i = enthalpy of adsorption of species i, kJ·mol μL = liquid viscosity, Pa·s ρL = liquid density, kg·m−3 ρS = catalyst particle density, kg·m−3 ϑi = fractional surface coverage of species i ϑH′ = fractional surface coverage of hydrogen ϑ0 = fractional surface coverage of vacant sites (type ○) ϑ′0 = fractional surface coverage of vacant sites (type □) Abbreviations

D = dimers G = gas H = hydrogen L = liquid MBA, A = 2-methyl-2-butanol MBE, E = 2-methyl-3-buten-2-ol MBY, Y = 2-methyl-3-butyn-2-ol (MBY)2, Y2 = complex originating in step s3 R = radical originating in step s4 S = solid



REFERENCES

(1) Bonrath, W.; Medlock, J.; Schütz, J.; Wüstenberg, B.; Netscher, T. Hydrogenation in the Vitamins and Fine Chemicals IndustryAn Overview. In Hydrogenation; Karamé, I., Ed.; Intech Open: Rijeka, Croatia, 2012. (2) 3-Buten-2-ol, 2-methyl-, CAS No. 115-18-4. In Chemicals Screening Information Data Set (SIDS) for High Volume Chemicals; UNEP Publications: Ispra, Italy, 1995. (3) Crespo-Quesada, M.; Grasemann, M.; Semagina, N.; Renken, A.; Kiwi-Minsker, L. Kinetics of the solvent-free hydrogenation of 2methyl-3-butyn-2-ol over a structured Pd-based catalyst. Catal. Today 2009, 147, 247−254. (4) Lindlar, H. Ein neuer Katalysator für selektive Hydrierungen. Helv. Chim. Acta 1952, 35, 446. (5) Lindlar, H.; Dubuis, R. Palladium catalyst for partial reduction of acetylenes. In Organic Syntheses; Wiley: New York, 1973; Vol. 5, p 880. (6) Bruehwiler, A.; Semagina, N.; Grasemann, M.; Renken, A.; KiwiMinsker, L.; Saaler, A.; Lehmann, H.; Bonrath, W.; Roessler, F. Threephase catalytic hydrogenation of a functionalized alkyne: Mass transfer 11550

DOI: 10.1021/acs.iecr.5b03424 Ind. Eng. Chem. Res. 2015, 54, 11543−11551

Article

Industrial & Engineering Chemistry Research phase hydrogenation of 2-butyne-1,4-diol to 2-butene-1,4-diol. Chem. Eng. Sci. 2002, 57, 3453−3460. (27) Molnar, A.; Sarkany, A.; Varga, M. Hydrogenation of carbon− carbon multiple bonds: chemo-, regio- and stereo-selectivity. J. Mol. Catal. A: Chem. 2001, 173, 185−221. (28) Semagina, N.; Renken, A.; Kiwi-Minsker, L. Monodispersed Pdnanoparticles on carbon fiber fabrics as structured catalyst for selective hydrogenation. Chem. Eng. Sci. 2007, 62, 5344−5348.

11551

DOI: 10.1021/acs.iecr.5b03424 Ind. Eng. Chem. Res. 2015, 54, 11543−11551