You are on page 1of 11

Article

Cite This: Ind. Eng. Chem. Res. XXXX, XXX, XXX−XXX pubs.acs.org/IECR

Selective Oxidation of n‑Butane over Vanadium Phosphate Based


Catalysts: Reaction Network and Kinetic Analysis
Christian Schulz,†,⊥ Felix Pohl,† Matthias Driess,†,‡ Robert Glaum,§ Frank Rosowski,†,∥
and Benjamin Frank*,†

Technische Universität Berlin, BasCat − UniCat BASF JointLab, Hardenbergstraße 36, 10623 Berlin, Germany

Department of Chemistry, Metalorganics and Inorganic Materials, Technische Universität Berlin, Straße des 17. Juni 135, Sekr. C2,
10623 Berlin, Germany
§
Institut für Anorganische Chemie, Rheinische Friedrich-Wilhelms-Universität Bonn, Gerhard-Domagk-Straße 1, 53121 Bonn,
Downloaded via UNIV OF LOUISIANA AT LAFAYETTE on September 23, 2018 at 06:05:35 (UTC).
See https://pubs.acs.org/sharingguidelines for options on how to legitimately share published articles.

Germany

BASF SE, Process Research and Chemical Engineering, Carl-Bosch-Straße 38, 67056 Ludwigshafen, Germany
*
S Supporting Information

ABSTRACT: The reaction network of n-butane selective oxidation was


comparatively analyzed over a novel αII-V0.8W0.2OPO4 orthophosphate and a
reference vanadyl pyrophosphate (VPP) based catalyst via parameter field
studies (T = 360−420 °C, x(n-butane) = 0.5−2.0%, x(O2) = 10−20%) as
well as cofeed and pulse experiments with presumed reaction intermediates.
For VPP the selectivity to the target product maleic anhydride (MAN) is
particularly sensitive to the reaction temperature, the n-butane concentration,
and the amount of cofed H2O, whereas the selectivity to MAN over the W-
containing catalyst is almost independent of all reaction parameters. The
roles of 1-butene, acetylene, furan, acetaldehyde, water, and 2,5-dihydrofuran
are discussed. Pulsing of possible C4 intermediates indicates that their
desorption from the catalyst surface is detrimental to MAN selectivity.
Ethylene and acetylene may be formed from MAN. The consecutive reaction
of acetylene with H2O can lead to acetic acid, whereas all other byproducts are predominantly formed directly from n-butane.
The pronounced stability of both samples was confirmed by repeated catalyst performance test under reference conditions,
XRD, and SEM. Mass and heat transport limitations were experimentally excluded. A formal kinetic model including n-butane,
MAN, CO, CO2, acetic acid, and acrylic acid was developed, in which the acids were found to be less relevant on the
V0.8W0.2OPO4 catalyst. However, the main reaction pathways were found to be similar over both catalysts, which differ mainly in
product selectivities.

■ INTRODUCTION
Since the discovery of vanadyl pyrophosphate (VO)2P2O7
however, on the cost of complex engineering and attrition or
friction issues to be solved.21 Recent theoretical studies by
Goddard et al. introduce a novel reaction model, i.e., a
(VPP) based catalysts in the 1970s,1 major efforts were put
reduction-coupled oxo activation (ROA) mechanism respon-
into the understanding of the reaction mechanism of the
sible for the catalytic selective activation and functionalization
selective oxidation of n-butane to maleic anhydride (MAN),
of n-butane to MAN over a P-based active site.22−24
which is one of the few industrial examples of oxidative
Several kinetic studies under steady state conditions have
activation of alkanes with molecular oxygen.2−13 A significant been performed to derive kinetic expressions of the reaction
improvement of the yield could be realized by advanced sequence.25 All of them, however, focus on the main reaction
synthesis protocols leading to high specific surface areas and products MAN and COx. In a highly cited report Buchanan
defect densities, respectively, as well as by the addition of and Sundaresan26 published kinetic data for VPP with different
promotors.10,14−19 The selectivity of MAN, however, appears V/P ratios (1.0 and 1.1), however, fitted only for MAN, CO,
to be limited to 70% at an industrially relevant conversion of and CO2, although byproducts like acetic acid (AceA), acrylic
80−85%, and no further breakthrough could be obtained
within the past 30 years.20 The origin of this observation and
Special Issue: Reinhard Schomacker Festschrift
especially the nature of the active sites as well as the reaction
mechanism of this highly complex reaction are still under Received: September 5, 2018
debate. A Mars−van Krevelen mechanism as proven by the Revised: September 7, 2018
participation of bulk oxygen in the reaction led to the concept Accepted: September 8, 2018
of a riser reactor, which can further increase the yield of MAN,

© XXXX American Chemical Society A DOI: 10.1021/acs.iecr.8b04328


Ind. Eng. Chem. Res. XXXX, XXX, XXX−XXX
Industrial & Engineering Chemistry Research Article

acid (AcrA), and ethylene were also detected. Noteworthy, a Herein we present a detailed parameters field study of the
kinetic inhibition of the redox process by MAN and H2O was selective oxidation of n-butane over V0.8W0.2OPO4 (in the
found. In a recent study Mestl et al. distinguish between the following denoted as VWPO) and a reference VPP catalyst,
intrinsic kinetics and P and/or H2O induced dynamics of the including the derivation of a formal kinetic model including
activity.27 In their study experimental data were measured in a MAN, CO, CO2, H2O, acetic acid, and acrylic acid. The loss of
bench-scale fixed bed reactor using an industrial vanadium phosphorus of VPP, which is discussed as an important issue in
phosphorus oxide (VPO) catalyst. However, the data quality long-term application, is not considered in this study.
possibly suffered from mass and heat transport limitation, and However, the decrease in selectivity of MAN was found to
the formation of acrylic and acetic acid was neglected. be negligible within a total time on stream (TOS) of ≤50 d.
Regarding the reaction network over VPP the main products An optimized parallel testing unit facilitates the acquisition of
are MAN, CO, and CO2. Byproducts like acetic and acrylic dense and consistent catalytic performance data, which allows
acids are formed in selectivity ranges below 3%.25,26 a detailed analysis of aging issues. The general catalyst
Furthermore, a variety of alkenes and oxygenates are found. performance in terms of activity and selectivity to MAN is
The sum of all byproducts (w/o acids) is typically below 1% similar to previously described industrial VPO.27 However, we
selectivity. Since the CO/CO2 ratio in the product gas is believe that the quality of the manuscript benefits from the
almost constant, i.e., VPP is not active for CO oxidation, most detailed comparison with previously published data. The
kinetic expressions are based on a triangular reaction scheme ability to reproduce state-of-the-art knowledge from literature
including the main products (Scheme 1). will enhance the credibility of novel aspects.

Scheme 1. Simplified Reaction Network over VPP for n-


Butane Oxidation
■ EXPERIMENTAL SECTION
Single-phase microcrystalline VWPO was prepared by solution
combustion synthesis (SCS) using (NH4)6W12O39·x H2O,
NH4VO3, and (NH4)2HPO4 as the precursors. Briefly, the
components were mixed together with glycine as the fuel and
HNO3 as the oxidizer, ignited in a muffle furnace at 500 °C,
and calcined at 700 °C for 4 d in air. Details of the preparation
and characterization of VWPO are described elsewhere.29,31
The low concentration or even absence of (possible) The VPP catalyst was prepared according to patent literature
reaction intermediates hinders the disclosure of the reaction via an organic route.19 Briefly, V2O5 was dissolved in H3PO4
sequence via kinetic analyses. A deuteration study by Pepera et and reduced with isopropanol. The VPP precursor was
al. clearly shows that the C−H activation in a methyl group is obtained by filtration, dried, and preactivated at 300 °C in
the rate-determining step of n-butane oxidation.9 The most air. The pelletized powder was used in kinetic studies after
accepted reaction pathway for MAN formation follows activation and equilibration under reference reaction con-
dehydrogenation of n-butane over 1-butene to butadiene. ditions. Different particle size fractions in the range of 50−
Oxygen insertion and ring closure leads to a 2,5-dihydrofuran 1400 μm were separated using analytical sieves made of
intermediate, which is converted to MAN (Scheme 2). stainless steel.
However, although all possible C4 intermediates can be BET measurements were carried out in a volumetric N2
converted to MAN,12 it remains unclear if the selective physisorption setup (Autosorb-6-B, Quantachrome) at the
conversion of n-butane to MAN necessarily requires temperature of liquid N2. The sample was degassed in a
desorption and readsorption steps or if the entire 14 electron dynamic vacuum at 423 K for 2 h prior to physisorption. Full
transfer is realized on a single active site. adsorption−desorption isotherms were recorded. The specific
To overcome the limitations of VPP based catalysts, which surface area according to the BET method was calculated from
could not be substantially improved in the past decades,2,6,20 a the linear range of the adsorption isotherm (p/p0 = 0.05−0.3).
systematic search for other active and selective phases was The catalyst was characterized before and after catalytic
performed. In a recent study29 we identified the W-substituted testing.
αII-VOPO4 as one of the very few alternatives to produce XRD measurements were performed in Bragg−Brentano
MAN from n-butane at all. Although the selectivity of 30% is reflection geometry on a theta/theta diffractometer (D8
relatively low as compared to the industrial standard, this Advance, Bruker AXS) equipped with a secondary graphite
performance belongs to one of the best ever found for VPP- monochromator (Cu Kα1+2 radiation) and scintillation
free systems. VV has been identified as an important detector. Sample powder was filled into the recess of a cup-
constituent of the selective catalyst.4,30 The substitution of shaped sample holder, the surface of the powder bed being
20% VV by WVI introduces an equilmolar fraction of VIV31 to flushed with the sample holder edge. Data analysis was
approach the average near-surface oxidation state of VPP, performed using the software package Topas (v 2.1, Bruker
which was found to be 4.1−4.3 under reaction conditions.11 AXS). Crystallite size values are based on the double-Voigt
Therefore, the catalyst can be seen as an interesting approach and reported as LVol‑IB values (volume weighted
complement to VPP. mean column length based on integral breadth) without

Scheme 2. Postulated Pathway of the Selective Oxidation of n-Butane to MAN via Gas-Phase and Surface Intermediates7,28

B DOI: 10.1021/acs.iecr.8b04328
Ind. Eng. Chem. Res. XXXX, XXX, XXX−XXX
Industrial & Engineering Chemistry Research Article

further assumptions about crystallite shape or size distribu- Using the parallel testing setup, the variation of space−time
tion.32 velocities was realized by loading the channels with different
SEM measurements were carried out using a Hitachi S-4800 amounts of VPP catalyst (Table S2). According to the results
microscope operating at 2 kV. The relative V, P, and O of the mass and heat transfer analyses, a particle size fraction of
contents were determined by means of energy dispersive X-ray 100−200 μm was chosen. Steatite was used to fix the undiluted
spectroscopy (EDX) using an EDAX Genesis spectrometer bed of catalyst particles in the isothermal zone of the reactors.
attached to the microscope, operated at 15 kV. To continuously control the actual feed concentrations, a blank
Catalytic tests were performed in a fully automated, reactor filled with inert steatite completed the set of channels
commercial 8-channel parallel testing setup (hte GmbH). to be tested. The standard gas flow of 33.3 mL min−1 per
The stainless steel reactors (ID 10 mm) are equipped with reactor, which was kept constant throughout the study, results
thermowells (OD 1/8′′). An additional reactor is filled with in GHSVs in between 500 and 16.000 h−1 to bring up both
inert material (calcined corundum) to control the inlet differential and integral rate data at almost each point of the
concentrations. All effluent gas streams are diluted with hot parameters field study.
inert gas to avoid product condensation. Permanent gases were The sequence of catalytic testing comprises a 3 × 3 matrix of
detected via TCD whereas organic compounds were quantified n-butane (2, 1, and 0.5%) and O2 (20, 15, and 10%) feed
over a FID. A detailed description of the setup can be found concentrations with excess of O2 in each point. Previous
elsewhere.29 experiments showed that a full conversion of O2 can lead to
The conversion (X) of n-butane (eqs 1 and 3) and the successive formation of coke, which irreversibly deactivates the
selectivities (S) of the respective products (eqs 2 and 4) are catalyst. Partial pressures of n-butane beyond 2% are not
calculated from the concentrations measured with reference to accessible without entering the explosive range. The total gas
the internal standard Ar. The latter are averaged over 5 runs pressure was kept constant at 1 bar(a) throughout the entire
measured at each temperature step, respectively. The run. Matching industrially relevant feed conditions, H2O was
selectivities of the (by)products (eq 2) were normalized by typically cofed at a concentration of 3%. The reaction
the number of carbon atoms (NC), respectively. Possible temperatures (420, 400, 380, and 360 °C) were investigated
byproducts include butenes, acrylic acid, propionic acid, acetic in decreasing order. Each point was measured 5 times to check
acid, acrolein, acetaldehyde, propane, propene, acetylene, the time scale of equilibration required for stable catalytic
ethylene, ethane, carbon monoxide, and carbon dioxide. performance. To monitor catalyst aging effects on activity and
There are no unidentified peaks in the gas chromatograms. product selectivity, a reference point was frequently measured
At low levels of n-butane conversion, the product-based (420 °C, 2% n-butane, 20% O2). A cofeed of a gaseous P
analysis of selectivities and conversions (eqs 3 and 4) is source, e.g., triethyl phosphate (TEP), to compensate the loss
preferred. of volatile phosphorus was avoided in this study, as this
c cAr,0 compound significantly affects the catalyst performance by
X = 1 − n‐butane × structural and/or electronic modification of active sites.
cn‐butane,0 cAr (1) Pulsing of reaction products and possible intermediates was
NC,i performed in a self-constructed setup described elsewhere.33
ci × 4 The catalyst was kept inside a single tube reactor in steady
Si =
cn‐butane,0 ×
cAr
− cn‐butane state at 2000 h−1 and 420 °C in a feed comprising 2% n-
cAr,0 (2) butane, 20% O2, and 3% H2O in inert gas (He, Kr) for an
cn‐butane initial equilibration period of 1 week. Pulsing of reactive
X=1− NC,i compounds (carried by an Ar stream) was realized by a pulse
cn‐butane + ∑ ci × ( 4 ) (3) valve in the H2O/He line of the setup, i.e., the concentrations
of n-butane and O2 remained unchanged during the transient
ci × NC,i experiment. The amount of the reactive compound pulsed over
Si = the catalyst was adjusted to its carbon number in a way that 1
∑ (c j × NC,j) (4)
substrate carbon atom comes upon approximately 20 surface
with i,j = reaction products. vanadium atoms to avoid severe and irreversible damage to the
To exclude disturbing macrokinetic factors, and thus ensure catalyst. This estimation is rather a rule of thumb and based on
a high quality of kinetic data, a pretest for mass and heat the specific surface area of the catalyst and the ideal crystal
transport limitations was performed. To address the influence structure of VPP,34 although it is known that the surface layer
of mass transport (pore diffusion), four different particle size of VPP under reaction conditions is rather amorphous35 and
fractions from 50 to 100 μm up to 1−1.4 mm were tested. In dynamically responds to changes in the reaction condi-
addition, different degrees of dilution of the VPP catalyst with tions.11,36 Repeated performance tests (GC) before and after
quartz in the same particle size are expected to reveal the each pulse series (3 per compound) proved the pronounced
influence of heat transport, which possibly lead to the stability of the system. The pulse itself was monitored by MS
formation of a hot spot in the catalyst bed. A detailed with a full scan (m/z 1−100) and a time-resolution of 2 s. The
overview over this pretest is shown in Table S1. The same components pulsed over the catalyst were furan, 2,5-
mass of active material (VPP) was filled into each reactor (0.72 dihydrofuran, 1-butene, acetaldehyde, acrolein, ethylene,
g). The equally distributed gas flow is adjusted to provide a acetylene, acetic acid, and acrylic acid.
space-time velocity of 2775 mL g−1 h−1 over each reactor (33.3 Cofeed experiments were performed in the same setup
mL min−1). The feed contains 2% n-butane, 20% O2, 3% H2O, where the transient experiments were performed. Catalyst
2% Ar, and balance N2. Activities and product selectivity for volumes were 0.1 and 0.2 mL for VPP and VWPO,
the 7 channels were compared at 420, 400, 380, and 360 °C respectively. The starting procedure and equilibration for the
with a holding time of 15 h at each temperature step. catalyst were identical to the parameters field study. MAN was
C DOI: 10.1021/acs.iecr.8b04328
Ind. Eng. Chem. Res. XXXX, XXX, XXX−XXX
Industrial & Engineering Chemistry Research Article

added to the reference gas stream by a saturator. Reference


measurements were done prior to and after cofeed experiments
to exclude changes in the active phase. Co-feed studies of CO,
ethylene, and water were performed during the parameters
field study prior to the end of the parameters field study.
Minimizing the effect of additional reactants, cofeed
compounds were matched to found concentrations under
reference conditions. Ethylene, CO, and water were added to
the reference feed. The testing procedure includes 5 GC
analyses per reactor. Before and after cofeed experiments,
reference measurements under standard conditions were run to
exclude changes in the active phase. Finally, the reactors were
cooled down to ambient temperature in 5% O2 in N2.
A simple power-law kinetics (eq 5) with Arrhenius-type
expressions for the rate constants (eq 6) was found to be
appropriate to describe the catalytic data.
Figure 1. Experimental analysis of mass and heat transfer limitations
ri = k ipa m pb n (5) for VPP. 2% n-butane, 20% O2, 3% H2O, balance N2, 2775 mL g−1
h−1, 360−420 °C.
k i = k 0 exp( −Ea,i /R /T ) (6)
Regarding the formation of CO and/or CO2 from catalyst bed. Negligible deviations from the ideal parity line can
combustion pathways of hydrocarbons and oxygenates it was likely be explained by unavoidable experimental errors such as
assumed that (i) carbon atoms in an oxidation state > +2 react slight deviations in reactor temperature, catalyst weight, or flow
to CO2, whereas (ii) for all other carbon atoms a common rate, respectively. In summary, the quality of catalytic
factor φ was established to describe the ratio of CO in the sum performance data can be regarded as reliable and consistent.
of carbon oxides formed by the combustion of this fragment VPP and VWPO catalysts provide specific surface areas of 30
(eq 7). Exemplarily, the MAN molecule contains 2 carbon and 3 m2 g−1, respectively, which are well preserved after the
atoms located at the anhydride functionality (oxidation state parameters field study. As confirmed by the similarly stable N2
+3), which are expected to form CO2. Instead, the two vinylic physisorption isotherms, the particles are free of considerable
C atoms (oxidation state −1) form 2φCO and 2(1 − φ)CO2. amounts of meso- or micropores. The structural stability is
The resulting stoichiometric equation for the combustion of confirmed by SEM overview images in different magnifications.
MAN is shown in eq 8. The EDX analysis averaged over 5 different locations reveals
that the P/V and P/(V+W) ratio of 1 ± 0.1 remains constant;
φ = CO/(CO + CO2 ) (7)
that is, a significant loss of P is not detected within the time-
scale of the study. Accordingly, the XRD analysis confirmed a
C4 H 2O3 + (3 − φ)O2 pronounced stability of both catalysts as well.
→ 2φCO + (4 − 2φ)CO2 + H 2O (8) For both samples the average C balance is 100 ± 3%
throughout the whole parameters field study, which lasted
The kinetic constants as well as φ were fitted to the entire approximately 60 days. The absence of gas phase reactions as
data set with the software Berkeley Madonna,37 using a fourth investigated in a separate reactor filled with inert steatite (Tab.
order Runge−Kutta algorithm and least-squares regression. S2) was confirmed by zero conversion of n-butane and O2,
The concentration profiles of the main components (n-butane, respectively, at each point of the testing program.
O2, MAN, CO, CO2) were weighted equally (factor 1), For the kinetic analysis of a catalytic reaction the stability of
whereas the trace components (AceA, AcrA) were weighted the catalyst over the extended duration of the parameters field
with a factor of 10. Values for rate orders and apparent study is a critical factor. Frequently changing reaction
activation energies obtained from graphical analyses via log− conditions can induce additional stress to the structural
log and Arrhenius plots, respectively, were taken for the integrity. However, only a consistent data set will allow
initialization of the numerical fitting procedure.


creation of a meaningful kinetic and/or mechanistic model. To
quantitatively follow the catalyst deactivation, the catalytic
RESULTS AND DISCUSSION performance under reference reaction conditions (2% n-
Figure 1 illustrates the pretest results for mass and heat transfer butane, 20% O2, 3% H2O, balance N2, 420 °C) was frequently
effects, which were only examined for the VPP sample due to tested (Figure 2). Clearly, in terms of n-butane and O2
its much higher activity. Here, the reference (abscissa) is the consumption the systems are sufficiently stable up to a TOS
undiluted 100−200 μm fraction. The impact of pore diffusion of 1000 h (VWPO: 600 h, see caption of Figure 2), except for a
on n-butane conversion is negligible and agrees with the short initial activation period of VPP. Due to the pronounced
absence of micro- and mesopores in the VPP catalyst. Even at stability in n-butane conversion the product selectivities can be
the upper temperature limit of 420 °C (X ≈ 80%) the directly compared. Thus, the slight loss of MAN selectivity at
conversion drops only by <5% for the bigger particles as the extended TOS can possibly be correlated with a marginal
compared to the reference. Instead, the lower conversions loss of (near-surface) P, which cannot be detected by EDX and
observed for the smallest particle size fraction (50−100 μm) is not replenished by a gaseous P source in this study. For both
are likely explained by channeling effects. The dilution of the catalysts the selectivities for CO, CO2, acetic acid, and acrylic
catalyst bed with quartz has only a minor effect on the activity acid, which are the compounds considered in the kinetic
and also confirms the absence of a significant hot spot in the model, remain stable as well. These results nicely confirm the
D DOI: 10.1021/acs.iecr.8b04328
Ind. Eng. Chem. Res. XXXX, XXX, XXX−XXX
Industrial & Engineering Chemistry Research Article

activation rather than a secondary combustion of the activated


product. However, beyond a conversion of n-butane of 80%
the combustion of MAN becomes increasingly relevant over
VPP and the decreasing CO/CO2 ratio indicates that the
secondary combustion forms more CO2 than the direct alkane
combustion. Unfortunately, due to the low activity of VWPO a
conversion higher than 70% is hardly achieved; thus, deeper
information about MAN combustion cannot be extracted from
this data set. Similarly, all other oxygenates (acetic acid, acrylic
acid, acetaldehyde, furan, acrolein) are formed in primary
reactions from n-butane. These compounds, however, are
much less stable than MAN and combust rapidly, with acetic
acid being the most abundant oxygenate at high n-butane
conversion. For most compounds a connection to MAN
cannot be found. Regarding acetic acid, however, a pretest with
VPP indicated that in the absence of n-butane and in a H2O-
rich feed MAN may react to acetic acid, whereas codosing of
MAN to the regular feed yields exclusively CO and CO2 in a
ratio of almost 1:1 (Table S3). In fact, an apparent stability of
acetic acid could originate from its regeneration as a secondary
product from MAN, which superimposes the direct formation
and degradation pathways. In principle, furan may react to
MAN; however, its concentration is far too low to represent a
main pathway of the network. Traces of alkenes and alkynes
are also present. 1-Butene is formed via oxidative dehydrogen-
ation of n-butane and as for furan its low concentration
questions its role as a significant intermediate in MAN
formation, although a consecutive oxidation to MAN is likely.
Indeed for some data sets the selectivity of MAN weakly
increases in the range of 10−40% of n-butane conversion (e.g.,
Figure 3d), which could be attributed to such selective side
Figure 2. Aging behavior of conversions and product selectivities
reactions. The partial oxidation of n-butane leads to <0.03%
during the testing sequence over VWPO (top) and VPP (bottom).
2% n-butane, 20% O2, 3% H2O, balance N2, VWPO: 500 h−1, VPP: propene which can further be oxidized to acetaldehyde, acetic
4000 h−1, 420 °C. Note that a power shut-down at t = 600 h caused acid, acrylic acid, or COx. With regard to the C2 hydrocarbons
an activity increase of VWPO, whereas VPP returned to its previous ethylene and acetylene a significant difference appears. As for
activity, thus catalytic data measured at t > 600 h (gray) is not the oxygenates and other alkenes, ethylene is a primary
considered in the discussion for VWPO. product from the degradation of n-butane over both catalysts.
For VWPO this is valid also for acetylene (Figure 3c). Instead,
absence of structural changes in the used catalyst samples as when plotting the acetylene selectivities vs n-butane con-
evidenced by the physicochemical characterization. versions for VPP over the entire parameters field (not shown)
For both catalysts the main products of the reaction are an intercept near 0% is visible. This is indicative for a
MAN, CO, and CO2, with the higher MAN selectivity for the consecutive product likely arising from the decomposition of
VPP sample. The lower MAN selectivity of 20% observed over MAN, although the cofeed of MAN predominantly yields CO
VWPO as compared to our previous study29 (30%) is the and CO2 (Figure 3f). Ethylene and acetylene formed in the
stable value after initial equilibration for 2 weeks. The reaction network are much more stable than propene and 1-
formation of byproducts is detected throughout the test in butene. The resulting general reaction network for both
low quantities. Acetic and acrylic acids are the most important catalysts is similar and shown in Scheme 3.
byproducts over VPP and frequently discussed in literature. As mentioned above for VWPO it turned out that the
Their abundance, however, is lower over VWPO. According to maximum conversion of n-butane obtained at 420 °C and 1%
the detection limit of our GC, trace amounts of acetaldehyde, n-butane/20% O2 is still too low to identify a significant
ethylene, acetylene, propene, furan, and acrolein are formed. degradation of MAN in the S/X plot. Consequently, kinetic
The systematic variation of contact times allows for the parameters for MAN combustion as described later could not
comparison of product selectivities at different degrees of n- be obtained. To overcome this, a cofeed study was performed
butane conversion under the same reaction conditions, adding MAN at different concentrations (0.2 and 0.4%), O2
respectively (Figure 3). concentrations (10 and 20%), and temperatures (400 and 420
On the basis of S/X trajectories exemplarily shown in Figure °C), as listed in Table S3. With regard to the reaction network
3 for the reference reaction conditions the reaction network it is clearly seen that MAN mainly reacts to CO and CO2 in a
can be discussed. As the MAN and COx trajectories cannot be ratio of approximately 1:1; however, small amounts of acetic
extrapolated to an intercept of zero, all main products stem and acrylic acids are formed as well from MAN.
from primary reactions. The constant (VWPO) or slowly For VPP the reaction network was further investigated by
decreasing (VPP) selectivity of MAN shows the remarkably cofeeding of selected reaction products in concentrations
high stability of this molecule over both catalysts and suggests similar to their appearances in the product stream, respectively
that the selectivity issue is related to a primary step of C−H (Table S3). As discussed above, the cofeed of MAN is
E DOI: 10.1021/acs.iecr.8b04328
Ind. Eng. Chem. Res. XXXX, XXX, XXX−XXX
Industrial & Engineering Chemistry Research Article

Figure 3. S/X plots by GHSV variation (2% n-butane, 20% O2, 3% H2O, balance N2, 420 °C) as obtained for VWPO (a−c, 500−2,000 h−1) and
VPP catalysts (d−f, 2,000−16,000 h−1). The product spectrum is subdivided into main products (a,d), oxygenates (b,e), and hydrocarbons (c,f).

predominantly converted to COx. Instead, CO and C2H4 are (Table S4). Clearly, when using the MS scan the limit of
not oxidized and totally inert with respect to the n-butane fragment deconvolution for minor byproducts is much lower
conversion. A significant influence, however, is observed for than the detection limit of regular GC analysis used in the
H2O, which increases the MAN selectivity, however, on the parameters field test.
cost of activity. Notably, the formation of acetic and acrylic This finding is similar to other C4 compounds, which are
acids is much more strongly promoted by H2O. Furthermore, frequently discussed as the intermediates of the multistep
transient reactant pulsing was performed over the working VPP conversion of n-butane to MAN, such as 1-butene or 2,5-
catalyst. Figure 4 exemplarily shows the characteristic MS dihydrofuran (Table S4). The MAN selectivity obtained with
traces (after correction for fragments of other species) for the these substrates is lower than with n-butane. A possible
furan pulse. Furan itself is fully converted during the pulse; interpretation of the experimental results is that the surface
thus, no signal at m/z 68 is observed during the transient intermediates once desorbed into gas phase under certain
experiment. The products detected are MAN, CO, CO2, and reaction conditions cannot completely readsorb in a way to
H2O, however, with an unexpectedly low selectivity to MAN continue the original reaction sequence starting from n-butane.
F DOI: 10.1021/acs.iecr.8b04328
Ind. Eng. Chem. Res. XXXX, XXX, XXX−XXX
Industrial & Engineering Chemistry Research Article

Scheme 3. Speculative Reaction Network as Discusseda byproducts approach zero selectivity at full n-butane
conversion (Figures 3b, c, e, and f). Notably, the expected
transformations from acetaldehyde and acrolein to acetic acid
and acrylic acid, respectively, could not be detected. Instead,
acrolein forms traces of ethylene and traces of acetic acid are
formed from acetylene, whereas all other C 2 and C 3
components exclusively react to CO and CO2.
Under the selected reference conditions, acetic and acrylic
acids are fully converted indicating primary formation from n-
butane. However, acetylene is converted to acetic acid with low
selectivities indicating a new pathway. At least for the route
from butenes to acetic acid a reaction pathway via 2-butanol is
proposed,38 with the hydration step likely being catalyzed by
acidic P−OH groups as present on VWPO and VPP catalysts.
However, with regard to this route it should be noted that in
our study neither 1-butene nor acetaldehyde and acrolein was
converted to acetic acid. The compounds MAN, CO, C2H4,
and H2O were introduced in steady-state cofeed mode (420
°C, 20% O2, 2% n-butane) in a concentration range that covers
the typical concentration of each compound, respectively, in
the parameters field test. Table S3 shows the results of the
cofeed study with the given concentrations of cofeed.
As discussed above MAN is predominantly formed in a
primary reaction from n-butane as well as by a minor
contribution from intermediary formed 1-butene and 2,5-
dihydrofuran, respectively. The latter pathways likely occur
also on VWPO. For VPP, MAN combustion becomes
a significant as the conversion of n-butane reaches 80−90% as
Solid lines: suggested by experiments, dashed lines: assumed/based
on literature. A reduced version of this figure (w/o COx) can be indicated by a sharp decline of MAN selectivity. In general, the
found in the Supporting Information (Figure S1). S/X curve is relatively stable and within the entire parameters
field the selectivity for MAN does not differ by more than 10
percentage points from the curve obtained under the reference
conditions. Whereas for VWPO the MAN selectivity appears
to be totally independent from reaction conditions (within the
range covered by the parameters field test) some trends are
observed for VPP: (i) The influence of temperature on
selectivity of MAN is quite small. Temperatures were varied in
the range of 360−420 °C. A parallel shift of S/X trajectories is
observed with the highest MAN selectivity at the lower end of
reaction temperatures. (ii) The MAN selectivity also steadily
increases with increasing n-butane (0.5−2 vol %) and also (iii)
with increasing H2O (0−10 vol %) feed concentrations,
respectively, whereas (iv) the concentration of O2 has no effect
on the S/X trajectory.
Over both catalysts acetic and acrylic acids are formed with a
Figure 4. Product traces after pulsing of furan over VPP under selectivity below 3%. Both acids are susceptible to combustion.
reaction conditions. DHF = 2,5-dihydrofuran. Over VPP, however, acetic acid is apparently more stable than
acrylic acid as indicated by the different signs of curvature,
This could be due to an immediate reoxidation of vacant which are typically concave and convex for acrylic and acetic
surface vanadia sites by gas phase O2 or bulk-to-surface acid, respectively (Figure 3e). Convex shaping, however, could
diffusion of mobile O species (Scheme 4), resulting in an also be induced by secondary formation of acetic acid from
oxidized state of the active site that might favor the total MAN, e.g., according to a potential reaction sequence depicted
oxidation of activated hydrocarbons such as alkenes and in Scheme 5. This pathway is clearly identified for both
oxygenates. catalysts (Table S3). The significance of hydrolysis pathways is
The results of the series of pulsing experiments are suggested by the increased formation of acetic acid in the
assembled in Table S4. For all substrates except ethylene presence of H2O in the feed as well as the pulse studies.
and acetylene full conversion is observed; that is, these Similarly, monodecarboxylation of hydrolyzed maleic acid,
compounds react much faster with the catalyst than n-butane. which is already known from patent literature,39,40 may serve as
The relative stability of ethylene and acetylene agrees well with a secondary source for acrylic acid (Table S3). Taking such
their detection as byproducts (S > 0.1%) under typical reaction hydrolysis pathways of MAN into account, it should be noted
conditions, and in particular with the plateau-like shape of their that in the presence of additional H2O in the feed the
S−X curves (Figure 3c and f), whereas all other C 2+ selectivity to MAN is still higher than in the dry feed.
G DOI: 10.1021/acs.iecr.8b04328
Ind. Eng. Chem. Res. XXXX, XXX, XXX−XXX
Industrial & Engineering Chemistry Research Article

Scheme 4. Suggested Interaction of Presumed Reaction Intermediates with Surface V Sites

Scheme 5. Speculative Reaction Routes from MAN to AceA selectivity is still under discussion41 and a focus of our
successive report.33 In our parameters field test the variation of
water feed was done at the end of the study after TOS of 50 d.
It should be mentioned that selectivity and conversion are
slightly different from the first reference point due to small
aging effects. Conversion of n-butane increases by 3% whereas
the selectivity of MAN decreases by 3%. During the long test
period VWPO and VPP may suffer from the loss of surface
phosphorus, which explains the found trends.

■ KINETIC ANALYSIS
The differential and integral kinetic data of the parameters field
studies were used to establish steady-state kinetic models for
both catalysts, which enable the simulation of the concen-
tration profiles of reactants and products within the reactor as a
function of residence time. Besides reactor modeling it can also
However, for both catalysts the main sources for acetic and be used for extrapolation to indicate attractive reaction
acrylic acids are the direct formation pathways from n-butane, conditions, which have not been covered by the widespread
as the trajectories have a positive intercept at zero n-butane parameters field study. The formal kinetic models include the
conversion. Interestingly, for the variation of H2O and n- main reaction products MAN, CO, CO2. Acetic and acrylic
butane, the trends of formation of organic acids are similar to acids are included only for VPP, whereas their traces over
MAN, i.e., their selectivities increase with the increasing VWPO were found to be insufficiently consistent with regard
content of H2O and/or n-butane, respectively, in the feed. It to the reaction conditions. The latter is likely due to their
should be noted that the higher hydrocarbon concentration lower overall concentrations and possibly a result of a lower
inevitably leads to higher H2O concentrations at similar time-on-stream stability for these byproducts. As discussed
degrees of conversion; thus, the effect seen for the n-butane above, a factor φ was introduced to describe the ratio of CO/
variation may be (in part) linked to a surface modification by COx arising from the combustion of nonoxygenated carbon
H2O. atoms in any combustible reaction product. Thus, the
The decrease in activity can be explained by competitive stoichiometric equations considered in the model are the
adsorption of H2O and n-butane. The positive effect on following:

Table 1. Kinetic Parameters of n-Butane Oxidation as Derived from Formal Kinetic Analysis
r0c r1f r1c r2f r2c r3f r3c
VWPO
Ea/kJ mol−1 85 67 72b
k0/mol kg−1s−1 1.14 × 104 3.22 × 103 1.65 × 103b
nCa 0.90 0.93 0.99b
nO2 0.13 0.00 0.49b
VPP
Ea/kJ mol−1 101 82 86 156 144 52 92
k0/mol kg−1 s−1 4.34 × 104 1.46 × 104 4.36 × 103 1.84 × 109 1.70 × 109 2.21 × 10° 8.52 × 102
nCa 0.34 0.50 0.50 0.83 0.68 0.63 0.21
nO2 0.37 0.42 0.24 0.53 0.33 0.17 0.11

a
Carbonaceous substrate. bEstimated from cofeed study (Table S3).

H DOI: 10.1021/acs.iecr.8b04328
Ind. Eng. Chem. Res. XXXX, XXX, XXX−XXX
Industrial & Engineering Chemistry Research Article

r0c: C4 H10 + (6.5 − 2φ)O2 both catalysts VWPO and VPP the stability of the crystal
structure and specific surface area were confirmed by XRD and
→ 4φCO + (4 − 4φ)CO2 + 5H 2O (9) SEM measurements. The product spectrum and reaction
network of the selective oxidation of n-butane appeared similar
r1f : C4 H10 + 3.5O2 → C4 H 2O3 + 4H 2O (10) over both catalyst systems and are cosistent with the reports
published in the literature with the exception of acetylene
r1c: C4 H 2O3 + (3 − φ)O2 which was not reported so far. Notably, MAN appears to be
→ 2φCO + (4 − 2φ)CO2 + H 2O (11) more stable over VWPO, however, on the cost of a much lower
initial selectivity. The assignment of these findings to electronic
r2f : C4 H10 + (3.5 − 0.5φ)O2 or structural properties of the model catalysts might lead to
further improvements. The appearances of ethylene and
→ C3H4O2 + φCO + (1 − φ)CO2 + 3H 2O (12) acetylene, as well as acetic and acrylic acids were found to
correlate with selectivity of MAN, indicating them (partially)
r2c: C3H4O2 + (3 − φ)O2 as byproducts of MAN decomposition. New reaction pathways
for acetic and acrylic acids were proposed based on parameters
→ 2φCO + (3 − 2φ)CO2 + 2H 2O (13)
field, pulse, and cofeed studies. Acrylic acid can be formed by
r3f : C4 H10 + 2.5O2 → 2C2H4O2 + H 2O (14) hydrolysis of MAN, which is subsequently monodecarboxy-
lated. Further, decarboxylation leads to ethene. Ethene is
r3c: C2H4O2 + (2 − 0.5φ)O2 oxidized to acetic acid via ethanol and acetaldehyde. Another
pathway of acetic acid formation is based on acetylene, which
→ φCO + (2 − φ)CO2 + 2H 2O (15) is formed from MAN. Hydrolysis of acetylene leads to
Here, the indices 0, 1, 2, 3, f, and c stand for n-butane, MAN, acetaldehyde, which can be selectively oxidized to acetic
acrylic acid, acetic acid, formation, and combustion, acid. Furan, 2,5-dihydrofuran, and 1-butene are discussed as
respectively. The kinetic constants extracted from the crucial reaction intermediates in the literature. Pulse studies
simultaneous fit are listed in Table 1. As a representative over equilibrated catalyst showed that MAN is formed from all
example, experimental data at the reference conditions assumed intermediates, but selectivites are lower than
together with simulated concentration profiles are shown in expected. The finding might be attributed to readsorption on
Figure 5, whereas the respective parity plots are given in the the surface, which leads to nonselective pathways. Never-
Supporting Information (Figure S2). theless, the results are in good agreement with the literature.
Carbon monoxide is not readsorbed on the active site and
cannot be oxidized to carbon dioxide. Temperature and water
and n-butane partial pressure have a positive impact on n-
butane concentration whereas oxygen partial pressure has
none. We implemented a kinetic model which includes the
main reaction products MAN, CO, CO2 as well as byproducts
like acetic acid and acrylic acid. Further, the role of oxygen
activation is included and reflects experimental evidence.
As a very important result the success in kinetic modeling
supports the reaction network proposed. The resulting model
can now be used for further experiments, i.e., advanced
modeling of profile reactors as already done by Horn et al.,42 or
extrapolation for the search of optimized reaction conditions.
The straightforward discussion of kinetic parameters in terms
of the reaction mechanism is not possible due to the formal
character of the kinetic model. In general, the apparent
activation energies for n-butane combustion (r0c) as well as
Figure 5. Kinetic modeling of n-butane oxidation. Exemplarily shown formation and combustion of MAN (r1f, r1c) are approximately
for the reference conditions over VPP: 2% n-butane, 20% O2, and 3%
15 kJ/mol lower for VWPO than for VPP, which is possibly
H2O at 420 °C, 2,000−16,000 h−1. Symbols: experimental data, lines:
kinetic model. due to the higher average oxidation state of VWPO. Such a
phenomenon has recently been reported for the oxidative
dehydrogenation over carbon-based catalysts.43 In fact, the

■ CONCLUSIONS
The parameters field study inside a parallel test setup was
pre-exponential factors of VWPO are only a factor of 3−4
smaller than those of VPP, although the specific surface area is
a factor of 10 smaller, indicating that the net activity of VWPO
found to be an advantageous method for acquiring steady state could be substantially increased by increasing the specific
kinetic data in a short time scale. An important aspect is the surface area (3 m2/g), e.g., by advanced synthesis protocols or
absence of mass and heat transfer limitations. The repeated by supporting the active phase. Notably, the apparent
measurement of the catalyst performance under reference activation energies of acetic acid formation and combustion,
reaction conditions shows a low activation with minor decrease respectively, are much higher than the values for the other
in MAN selectivity for VPP, whereas VWPO unintentionally reactions over VPP. As we discussed previously, acetic acid
activates in the middle of the parameters field test while may also be formed in a secondary step from MAN. Thus, the
keeping the selectivity pattern. The data unsuitable for kinetic exceptionally high apparent energies could point at a more
fitting could be easily identified by control experiments. For complex reaction network as the basis for kinetic modeling. To
I DOI: 10.1021/acs.iecr.8b04328
Ind. Eng. Chem. Res. XXXX, XXX, XXX−XXX
Industrial & Engineering Chemistry Research Article

solve this issue, dedicated cofeed studies with MAN and acetic Present Address
acid are required. The apparent rate orders for most reactions ⊥
(C.S.) hte GmbH, the high throughput experimentation
lie in between 0.1 and 0.9, with the rate order of the company, Kurpfalzring 104, 69123 Heidelberg, Germany.
carbonaceous compound being always higher that the rate
Notes
order of O2. This can be tentatively interpreted as an indication
that both steps, the catalyst oxidation and the hydrocarbon The authors declare no competing financial interest.
conversion, contribute equally to the overall rate with the
catalyst being in a rather oxidized state. Mechanistic models to
be developed and fitted to the data set can possibly shed light
■ ACKNOWLEDGMENTS
This research was conducted in the framework of the
on this aspect. MAN formation over VWPO (r1f) is a single BasCatJointLab between BASF SE, the TU Berlin, and the
example with an almost first order reaction with respect to n- FHI and received no external funding. We thank the Berlin
butane and zero rate order with respect to O2. Here, we can Cluster of Excellence UniCat for support.


conclude that the surface active site for the rate determining
step is fully oxidized and the overall rate is only limited by the
activation of n-butane. However, for all other reactions the REFERENCES
situation is more complex. (1) Bergman, R. I.; Frisch, N. W. Production of maleic anhydride by
If the product pattern and kinetic aspects of a catalyst under oxidation of n-butane. U.S. Patent 3,293,268, 1964.
certain reaction conditions can be regarded as a sensitive (2) Centi, G. Vanadyl pyrophosphate - a critical overview. Catal.
fingerprint for the nanostructure of the active site of a catalyst, Today 1993, 16, 5.
then the apparent similarities between VWPO and VPP point (3) Hutchings, G. J. Vanadium phosphate: a new look at the active
at a similarly structured active domain. This aspect should be components of catalysts for the oxidation of butane to maleic
anhydride. J. Mater. Chem. 2004, 14, 3385.
discussed in terms of the novel mechanistic reaction model (4) Coulston, G. W. The kinetic significance of V5+ in n-butane
suggested recently by Goddard et al.22−24 The ROA oxidation catalyzed by vanadium phosphates. Science 1997, 275, 191.
mechanism is based on the redox-transformation of reduced (5) Volta, J.-C. Vanadium phosphorus oxides, a reference catalyst for
(VO)2P2O7 and oxidized X1-VOPO4 phases. Such trans- mild oxidation of light alkanes: a review. C. R. Acad. Sci., Ser. IIc:
formation cannot easily be transferred to the αII phase of our Chim. 2000, 3, 717.
VWPO catalyst, although the surface composition and (6) Ballarini, N.; Cavani, F.; Cortelli, C.; Ligi, S.; Pierelli, F.; Trifirò,
structure of both VPP and VWPO systems are still under F.; Fumagalli, C.; Mazzoni, G.; Monti, T. VPO catalyst for n-butane
debate and unknown, respectively. We will address this oxidation to maleic anhydride. Top. Catal. 2006, 38, 147.
apparent contradiction in future surface-sensitive studies. (7) Bartholomew, C. H.; Farrauto, R. J. Fundamentals of industrial
Further development and validation of the kinetic model is catalytic processes; John Wiley & Sons: Hoboken, NJ, USA, 2005.
crucial. Experiments including macro kinetics such as heat and (8) Hodnett, B. K. Vanadium-phosphorus oxide catalysts for the
mass transport limitation are advisible based on this micro selective oxidation of C4 hydrocarbons to maleic anhydride. Catal.
Rev.: Sci. Eng. 1985, 27, 373.
kinetic model. The reaction network is quite simple and should
(9) Pepera, M. A.; Callahan, J. L.; Desmond, M. J.; Milberger, E. C.;
include all other byproducts. Therefore, detailed reaction Blum, P. R.; Bremer, N. J. Fundamental study of the oxidation of
network studies are necessary. Reaction intermediates have to butane over vanadyl pyrophosphate. J. Am. Chem. Soc. 1985, 107,
be introduced over equilibrated catalysts. The role of bulk 4883.
oxygen is quite controversial, so isotope labeling studies should (10) Guliants, V. V.; Benziger, J. B.; Sundaresan, S.; Wachs, I. E.;
be taken into consideration. For VPP we identified the reaction Jehng, J.-M.; Roberts, J. E. The effect of the phase composition of
temperature as the most important factor to realize changes in model VPO catalysts for partial oxidation of n-butane. Catal. Today
MAN selectivity. A lower reaction temperature, however, is 1996, 28, 275.
accompanied by a significant loss of activity; thus, future efforts (11) Eichelbaum, M.; Hävecker, M.; Heine, C.; Karpov, A.; Dobner,
should address structural factors such as specific surface area, C.-K.; Rosowski, F.; Trunschke, A.; Schlögl, R. The intimate
e.g., by supporting VPP or developing templating techniques. relationship between bulk electronic conductivity and selectivity in


the catalytic oxidation of n-butane. Angew. Chem., Int. Ed. 2012, 51,
ASSOCIATED CONTENT 6246.
(12) Guliants, V. V.; Benziger, J. B.; Sundaresan, S. The oxidation of
* Supporting Information
S C4 molecules on vanadyl pyrophosphate catalysts. Stud. Surf. Sci.
The Supporting Information is available free of charge on the Catal. 1996, 101, 991.
ACS Publications website at DOI: 10.1021/acs.iecr.8b04328. (13) Teles, J. H.; Hermans, I.; Franz, G.; Sheldon R. A Oxidation. In
Ullmann’s Encyclopedia of Industrial Chemistry; Wiley-VCH: Wein-
Details of catalytic testing, catalytic data of pulse and heim, 2015; pp 1−103.
cofeed studies, reduced reaction network, and parity (14) Zazhigalov, V. A.; Haber, J.; Stoch, J.; Pyatnitzkaya, A. I.;
plots of kinetic fits (PDF) Komashko, G. A.; Belousov, V. M. Properties of cobalt-promoted


(VO)2P2O7 in the oxidation of butane. Appl. Catal., A 1993, 96, 135.
AUTHOR INFORMATION (15) Hutchings, G. J.; Higgins, R. Effect of promoters on the
selective oxidation of n-butane with vanadium-phosphorus oxide
Corresponding Author catalysts. J. Catal. 1996, 162, 153.
*Fax: (+49) 030-314-27331, E-mail: b.frank@bascat.tu-berlin. (16) Hutchings, G. J.; Sananes, M. T.; Sajip, S.; Kiely, C. J.; Burrows,
de. A.; Ellison, I. J.; Volta, J.-C. Improved method of preparation of
ORCID vanadium phosphate catalysts. Catal. Today 1997, 33, 161.
(17) Johnson, J. W.; Johnston, D. C.; Jacobson, A. J.; Brody, J. F.
Matthias Driess: 0000-0002-9873-4103 Preparation and characterization of vanadyl hydrogen phosphate
Robert Glaum: 0000-0001-5805-1466 hemihydrate and its topotactic transformation to vanadyl pyrophos-
Benjamin Frank: 0000-0001-6928-9568 phate. J. Am. Chem. Soc. 1984, 106, 8123.

J DOI: 10.1021/acs.iecr.8b04328
Ind. Eng. Chem. Res. XXXX, XXX, XXX−XXX
Industrial & Engineering Chemistry Research Article

(18) Ghelfi, F.; Mazzoni, G.; Fumagalli, C.; Cavani, F.; Pierelli, F. (41) Lesser, D.; Mestl, G.; Turek, T. Transient behavior of vanadyl
Niobium-doped vanadium/phosphorus mixed oxide catalyst. U.S. pyrophosphate catalysts during the partial oxidation of n-butane in
Patent 7,638,457 (B2), 2004. industrial-sized, fixed bed reactors. Appl. Catal., A 2016, 510, 1.
(19) Weiguny, J.; Storck, S.; Duda, M.; Dobner, C. Catalyst and (42) Dong, Y.; Geske, M.; Korup, O.; Ellenfeld, N.; Rosowski, F.;
method for producing maleic anhydride. European Patent 1,487,576 Dobner, C.; Horn, R. What happens in a catalytic fixed-bed reactor for
(A1), 2003. n-butane oxidation to maleic anhydride? Chem. Eng. J. 2018, 350, 799.
(20) Trifirò, F.; Grasselli, R. K. How the yield of maleic anhydride in (43) Khavryuchenko, O. V.; Frank, B. Impact of edge oxidation state
n-butane oxidation, using VPO catalysts, was improved over the years. on red-ox barriers during hydrocarbon oxydehydrogenation over
Top. Catal. 2014, 57, 1188. carbon nanotube catalysts. J. Phys. Chem. C 2017, 121, 3958.
(21) Contractor, R. M.; Sleight, A. W. Selective oxidation in riser
reactor. Catal. Today 1988, 3, 175.
(22) O’Leary, W. C.; Goddard, W. A.; Cheng, M.-J. Dual-phase
mechanism for the catalytic conversion of n-butane to maleic
anhydride by the vanadyl pyrophosphate heterogeneous catalyst. J.
Phys. Chem. C 2017, 121, 24069.
(23) Cheng, M.-J.; Goddard, W. A.; Fu, R. The reduction-coupled
oxo activation (ROA) mechanism responsible for the catalytic
selective activation and functionalization of n-butane to maleic
anhydride by vanadium phosphate oxide. Top. Catal. 2014, 57, 1171.
(24) Cheng, M.-J.; Goddard, W. A. The critical role of phosphate in
vanadium phosphate oxide for the catalytic activation and
functionalization of n-butane to maleic anhydride. J. Am. Chem. Soc.
2013, 135, 4600.
(25) Centi, G.; Trifirò, F.; Ebner, J. R.; Franchetti, V. M.
Mechanistic aspects of maleic anhydride synthesis from C4 hydro-
carbons over phosphorus vanadium oxide. Chem. Rev. 1988, 88, 55.
(26) Buchanan, J. S.; Sundaresan, S. Kinetics and redox properties of
vanadium phosphate catalysts for butane oxidation. Appl. Catal. 1986,
26, 211.
(27) Mestl, G.; Lesser, D.; Turek, T. Optimum performance of
vanadyl pyrophosphate catalysts. Top. Catal. 2016, 59, 1533.
(28) Chen, B.; Munson, E. J. Evidence for two competing
mechanisms for n-butane oxidation catalyzed by vanadium
phosphates. J. Am. Chem. Soc. 1999, 121, 11024.
(29) Schulz, C.; Roy, S. C.; Wittich, K.; d’Alnoncourt, R. N.; Linke,
S.; Strempel, V. E.; Frank, B.; Glaum, R.; Rosowski, F. αII-
(V1‑xWx)OPO4 catalysts for the selective oxidation of n-butane to
maleic anhydride. Catal. Today, DOI: 10.1016/j.cattod.2018.05.040.
(30) Aït-Lachgar, K.; Abon, M.; Volta, J. C. Selective oxidation of n-
butane to maleic anhydride on vanadyl pyrophosphate. J. Catal. 1997,
171, 383.
(31) Roy, S. C.; Glaum, R.; Abdullin, D.; Schiemann, O.; Quang
Bac, N.; Lii, K.-H. Solid solution formation between vanadium(V)
and tungsten(V) oxide phosphate. Z. Anorg. Allg. Chem. 2014, 640,
1876.
(32) Snyder, R. L.; Fiala, J.; Bunge, H.-J. Defect and microstructure
analysis by diffraction; Oxford science publications 10; Oxford
University Press: Oxford, 1999.
(33) Schulz, C.; Kraehnert, R.; Rosowski, F.; Frank, B. Selective
oxidation of n-butane over vanadium-phosphorus oxide: oxygen
activation and dynamics. Submitted for publication in ChemCatChem.
(34) Geupel, S.; Pilz, K.; van Smaalen, S.; Büllesfeld, F.; Prokofiev,
A.; Assmus, W. Synchrotron-radiation study of the two-leg spin-ladder
(VO)2P2O7 at 120 K. Acta Crystallogr., Sect. C: Cryst. Struct. Commun.
2002, 58, i9.
(35) Heine, C.; Hävecker, M.; Stotz, E.; Rosowski, F.; Knop-
Gericke, A.; Trunschke, A.; Eichelbaum, M.; Schlögl, R. Ambient-
pressure soft x-ray absorption spectroscopy of a catalyst surface in
action. J. Phys. Chem. C 2014, 118, 20405.
(36) Abon, M. Evolution of a VPO catalyst in n-butane oxidation
reaction during the activation time. J. Catal. 1995, 156, 28−36.
(37) Macey, R.; Oster, G. Berkeley Madonna 2009.
(38) Yamashita, T.; Ninagawa, S.; Kato, T. Synthesis of acetic acid
by catalytic oxidation of butenes (part 1). Bull. Jpn. Pet. Inst. 1976, 18,
167.
(39) Fruchey, O. S.; Reeves, C. T.; Brooks, W. C.; Yang, E. J.;
Roundy, R. L. Method of decarboxylating maleic acid to acrylic acid.
European Patent 2,786,980 (A1), 2008.
(40) Lloyd, A. P. Decarboxylation. U.S. Patent 3,476,803, 1966.

K DOI: 10.1021/acs.iecr.8b04328
Ind. Eng. Chem. Res. XXXX, XXX, XXX−XXX

You might also like