• No results found

Effect of oxygen on formic acid decomposition over Pd catalyst

N/A
N/A
Protected

Academic year: 2021

Share "Effect of oxygen on formic acid decomposition over Pd catalyst"

Copied!
11
0
0

Bezig met laden.... (Bekijk nu de volledige tekst)

Hele tekst

(1)

Effect of oxygen on formic acid decomposition over Pd catalyst

Pengyu Xu, Fernando D. Bernal-Juan, Leon Lefferts

Catalytic Processes and Materials Group, Faculty of Science and Technology, MESA+ Institute for Nanotechnology, University of Twente, PO Box 217, 7500 AE Enschede, the Netherlands

a r t i c l e i n f o

Article history: Received 18 September 2020 Revised 30 October 2020 Accepted 31 October 2020 Available online xxxx Keywords:

Formic acid decomposition Oxygen

Liquid phase Pd/c-Al2O3catalyst

CO poisoning

a b s t r a c t

It is well known that Pd based catalyst deactivate during formic-acid decomposition in aqueous phase at mild temperatures. This study reports on a kinetic study of formic acid decomposition over Pd/c-Al2O3 catalysts including the effect of traces of oxygen, as well as pretreatment of the catalysts and supported by in-situ Attenuated Total Reflection Infrared Spectroscopy experiments. The results show that deactiva-tion of Pd/c-Al2O3catalysts can be suppressed by adding traces of oxygen. This is assigned to removal of adsorbed CO, poisoning the Pd surface, via oxidation to CO2. The activity of the catalyst during operation is maintained, promoting the H2production compared to operation in absence of any oxygen. Clearly, oxygen oxidizes CO preferentially over H2under the condition that the oxygen concentration is kept below 0.1 vol% in this study. Further increasing the oxygen concentration further increases conversion rate of formic-acid but also decreases the hydrogen yield significantly because formic acid oxidation and/or consecutive H2oxidation become dominating. The results of this study are important because the effect of traces of oxygen from ambient has not been considered in most of the reports in literature. Ó 2020 The Author(s). Published by Elsevier Inc. This is an open access article under the CC BY license (http://creativecommons.org/licenses/by/4.0/).

1. Introduction

In times of fossil fuel resources shortage and concern about CO2

emissions, the search for alternative and sustainable energy sources has become more pressing than ever. Hydrogen has attracted an increasing level of attention as an important energy vector and may play a significant role in power distribution in the future. However, due to the extremely low critical point and very low density of hydrogen gas, it is particularly difficult to store efficiently, especially on long term. Many molecules have been proposed as hydrogen carriers, e.g. ammonia [1], methanol [2], methane[3] as well as higher hydrocarbons[4]. Another option is formic-acid which can be produced from CO2and green

hydro-gen [5–7], resulting in a carbon-neutral process. Hydrogenation of CO2to formic acid requires either high pressure or operation

in aqueous solution, preferably at basic conditions, because of ther-modynamic limitations [6]. Formic-acid is a low-toxic chemical that can be easily stored, transported and handled. In addition, formic-acid is a significant by-product from biomass conversion. This work focusses on using liquid-phase formic-acid as a hydrogen storage material.

Obviously, in order to make the stored hydrogen available, formic-acid needs to be decomposed to CO2and H2according to

equation (1). Homogeneous catalysts for formic-acid

decomposition have been intensively studied [5,8,9]. Unfortu-nately, separation of the dissolved catalyst and the use of organic solvents, ligands and additives complicate the design of suitable devices[10]. Therefore, application of heterogeneous catalysts is preferred. Previous studies on heterogeneous catalysis have been performed in gas phase, requiring elevated temperatures [11– 13]. In this light, operation in aqueous phase at mild temperatures might be advantageous. Various Pd based mono-[14], bi-[15]and tri-metallic[16]catalysts have been identified as most active at low temperature and also resulting in high H2selectivity. The main

drawback of these Pd-based catalyst operated at mild temperature is poisoning by CO, which is formed via the dehydration reaction

[17–20], according to equation(2) [13,21–23].

HCOOH lð Þ ! CO2ð Þ þ Hg 2ð Þg

D

G¼ 48:4 kJ=mol;

D

H

¼ 31:2 kJ=mol ð1Þ

HCOOH lð Þ ! COðgÞ þ H2OðlÞ

D

G¼ 28:5 kJ=mol;

D

H

¼ 28:4 kJ=mol ð2Þ

So far, all heterogeneous catalysts studies suffer from deactiva-tion to some extent. However, the deactivadeactiva-tion mechanism is still under debate. Ruthven et al.[24]proposed that deactivation is due to formation of palladium hydride, supposedlyb-Pd hydride, which is suggested to be inactive. Hu et al.[25]suggested that catalyst deactivation is caused by occupation of active sites by protons, CO2, H2O and HCOO intermediate species. However, CO poisoning

https://doi.org/10.1016/j.jcat.2020.10.032

0021-9517/Ó 2020 The Author(s). Published by Elsevier Inc.

This is an open access article under the CC BY license (http://creativecommons.org/licenses/by/4.0/). ⇑Corresponding author.

E-mail address:l.lefferts@utwente.nl(L. Lefferts).

Contents lists available atScienceDirect

Journal of Catalysis

j o u r n a l h o m e p a g e : w w w . e l s e v i e r . c o m / l o c a t e / j c a t

Please cite this article as: P. Xu, F.D. Bernal-Juan and L. Lefferts, Effect of oxygen on formic acid decomposition over Pd catalyst, Journal of Catalysis,https:// doi.org/10.1016/j.jcat.2020.10.032

(2)

[17,19,20,26]is proposed most frequently to cause catalyst deacti-vation. CO adsorbs much more strongly on Pd than H2 and CO2,

considering adsorption enthalpies of 150 kJ/mol for CO [27] at low coverage, versus 100 kJ/mol [28]and 80 kJ/mol [29]for H2

and CO2, respectively. Jiang et al.[18]confirmed presence of

che-misorbed CO during formic-acid decomposition, using Attenuated Total Reflection Infrared Spectroscopy (ATR-IR).

CO adsorbed on Pd reacts easily with oxygen at room tempera-ture forming CO2, even in aqueous phase as e.g. observed with

ATR-IR[30]. Therefore, it may be expected presence traces of oxy-gen could influence the formic acid decomposition reaction. It is remarkable that in several studies on formic acid decomposition in aqueous phase, the gas composition in the reactor is not clearly defined[18,20,23,31–35], while other studies were performed in a reactor open to ambient[14,36–40]. Only in a few studies oxygen was rigorously removed[24,41,42], but the influence of the oxygen concentration on formic-acid decomposition has not been reported yet. This study reports on the influence of the oxygen concentra-tion on rate, selectivity and deactivaconcentra-tion in formic-acid decomposi-tion over Pd catalysts.

2. Experimental section 2.1. Materials

Commercial

c

-Al2O3powder purchased from BASF,

character-ized by a surface area of 195 m2/g, was used as catalyst supports

in this study. Tetraamminepalladium (II) nitrate solution (10 wt% in H2O, 99.99%) was purchased in Sigma-Aldrich was used as

cata-lyst precursor solution. Ammonium (50% v/v water), Formic-acid (98%) and sodium hydroxide were purchased from Sigma-Aldrich. Pre-mixed O2in Ar (0.1 vol%) and CO in Ar (10 vol%) was

purchased from Linde. All the aqueous solutions were prepared using ultra purified water obtained with a water purification sys-tem (Millipore, Synergy).

2.2. Catalyst preparation

The Pd/

c

-Al2O3catalyst containing 1 wt% of palladium, was

pre-pared by wet impregnating method. The method was described in elsewhere[43,44]. In brief, 10 g (0.098 mol) of the sieved support (size less than 20

l

m) was calcined at 600 °C in air for 4 h to remove any organic pollutants. Then the support was suspended in 100 mL millQ water, the pH of the solution was adjusted by add-ing 2 mL ammonia solution to maintain the pH around 9, checked with a pH meter in order to ensure electrostatic interaction of Pd (NH3)42+with the negatively charged alumina surface. Then 3 g of

the original palladium precursor solution (containing 1 mmol Pd (NO3)24NH3) was added in the suspension slowly and stirred at

room temperature for at least 1 h. Further, the solution was trans-ferred to the rotary evaporator to remove the liquid during 2 h at 70 °C. Then the catalyst was calcined in air at 400 °C for 3 h (5 K/min), switched to N2gas for 20 min, followed by reduction

in 50 vol% hydrogen diluted in nitrogen (total flow rate 60 mL/ min) at the same temperature for 3 h. The catalysts were charac-terized and used in catalytic tests as prepared and without any fur-ther treatment.

2.3. Catalyst characterization

The BET surface area of the catalyst, degassed at 300 °C for 1 day, was determined with N2-adsorption at 77 K (Micromeritics

Tristar). The Pd loading on the alumina support was determined with X-ray fluorescence spectroscopy (XRF, Philips PW 1480). The metal surface area that is accessible was determined with

CO chemisorption at room temperature (Chemisorb 2750, Micromeritics). Typically, the sample was reduced at room ature in hydrogen for 1 h and flushed with He at the same temper-ature for 0.5 h. Then, CO was introduced as pulses and the response was recorded using a TCD detector. Pd particle sizes are estimated assuming hemispherical metal particles and assuming that the sto-ichiometric ratio of adsorbed CO and Pd surface atoms is one. X-ray photoelectron spectroscopy (XPS) was conducted on as-prepared catalysts by an Omicron Nanotechnology GmbH (Oxford Instru-ments) surface analysis system with a photon energy of 1486.7 eV (Al K

a

X-ray source) with a scanning step size of 0.1 eV and a pass energy of 20 eV. Due to the poor electrical con-ductivity of sample surface, it is necessary to neutralize charge on the sample with an electron spray. The spectra were corrected using the binding energy of C 1s peak as a reference.

2.4. Catalytic tests

Activity and selectivity of the catalysts were measured in a 1 L batch reactor operated at 20°C at atmospheric pressure. The glass reactor (DURANÒBAFFLED, WIDE MOUTH BOTTLE GLS 80Ò) has a diameter of 10.1 cm and height 22.2 cm. The reactor has four con-nections on the reactor lid for gas-in, gas-out, sampling and a stir-ring shaft equipped with 4 stirstir-ring blades.

The reaction condition is shown inTable 1. Typically, 0.1 g cat-alyst was suspended in 0.3 L milli-Q water and stirred at 625 rpm while flushing with a mixture gas of Ar and oxygen (between 0 and 2 vol% oxygen) with a flow rate of 50 mL/min for at least one hour to remove any gasses dissolved in the water and gas in the 700 mL gas cap. The gas flow rates (NTP) were controlled using mass flow controllers (Brooks 5850S). The reaction was started by introduc-ing 60 mL pure formic-acid solution. The initial pH was varied between 2 and 10 by adding the appropriate amount of sodium hydroxide (1 M solution) to the formic-acid solution. The stability experiments were performed by injecting additional 15mL pure formic-acid, one hour after the first injection. This procedure was repeated three times.

The procedure described above includes experiments under Ar, in absence of any oxygen and this experiment is termed as inert-experiment. In addition, a semi-inert-experiment was conducted by flowing pure Ar (50 mL/min), like in the inert experiment, with the difference that the gas pipe inlet was positioned above the liq-uid level so that the dissolved oxygen as well as gasses produced during the experiment are very slowly removed from the water phase.

Additional experiments were performed after pretreatment with CO according the following procedure. The reactor was first flushed with inert for 90 min (99.9 vol% Ar, 60 mL/min), followed by reduction of the catalyst in 20 vol% H2/Ar for 50 min (Flowrate

50 mL/min). After that, the reactor was again flushed with inert gas until no H2was detected in the outlet stream of the reactor. Then a

gas mixture containing 10 vol% CO in Ar (Linde Gas Benelux B.V.,

Table 1

Operating conditions of the nitrite hydrogenation in slurry reactor.

Reaction temperature,°C 20

Reaction volume, L 0.3

pH of the solution 2–10

Stirring speed, rpm 625

Tested partially hydrophilic catalyst particle size,mm 0–20

Amount of catalyst, g 0.1

Initial formic-acid concentration, mmol/L 1.5–10

Total gas flow rate, mL/min NTP 50

Total operating pressure, bar 1

Oxygen partial pressure, bar 0–0.02

(3)

50 mL/min) was fed to the reactor for 30 min, followed by flushing once more with inert gas for about 40 min (99.9 vol% Ar, 50 mL/ min) to remove CO from the reactor. Finally, the experiment was initialized by adding formic-acid solution to the reactor while flowing with inert gas.

Samples were taken at different reaction times using a 2.5 mL syringe (BD Plastipak) and filtered through a syringe filter (PTFE, 0.2

l

m, Whatman) in order to remove the catalyst. Formic-acid concentrations were measured with ion-chromatography (DIO-NEX, ICS 3000) equipped with an Ultimate autosampler. The gas products were measured with an online micro-GC, sampling every 5 min, measuring the concentrations of H2, CO2and CO. The

num-ber of moles of H2, CO2and CO were obtained by the integration

over time of each of the observed flow rates, which were calculated using the concentration profile measured by GC and the total gas flow rate used in the experiment.

Formic-acid conversion and H2, CO2, CO yield were calculated

according to equation(3),(4),(5) and (6), respectively.

HCOOH con

v

ersion¼nHCOOH;t0 nHCOOH;t1

nHCOOH;t0  100 ð3Þ H2 yield¼ nH2;t1 nHCOOH;t0 100 ð4Þ CO2 yield¼ nCO2;t1 nHCOOH;t0 100 ð5Þ CO yield¼ nCO;t1 nHCOOH;t0 100 ð6Þ

nHCOOH;t0 and nHCOOH;t1 are the initial amount of formic-acid in

moles and at t1, respectively. Likewise, nH2;t1, nCO2;t1and nCO;t1are the integral amounts of H2, CO2 and CO formed at t1 in moles,

respectively.

The apparent turnover frequency (TOF) was calculated accord-ing equation(7), based on conversion at t1 lower than 20%.

TOF¼½con

v

ersion rate HCOOHt1mol s1

moles of a

v

ailable surface Pd mol ð7Þ

The number of available Pd surface-atoms was obtained with CO-chemisorption.

2.5. ATR-IR

The preparation of catalyst layer on the ATR crystal is described elsewhere[30,45–48]. Briefly, a suspension containing 0.1 g 1 wt% Pd/

c

-Al2O3dispersed in 20 mL 2-propanol was prepared. In order

to prevent cracking of the catalyst layer, the suspension was soni-cated with an ultrasonic processor (Fisher Scientific-705) for 1 h. Subsequently, the suspension was spray-coated on a trapezoidal ZnSe crystal (52.5 mm * 20 mm* 2 mm, facet angle 45°, Anadis instruments BV), which was placed on a hot plate at 150°C, result-ing in about 5 mg catalyst on the crystal. Then, the coated crystal was calcined at 300 °C (1 °C/min) for 1 h in N2 atmosphere

(20 mL/min). It was mounted in a home-build in-situ ATR-IR cell which has been described in detail elsewhere [45,46]. The cell was mounted in the sample compartment of an infrared spectrom-eter (Tensor 27, Bruker) equipped with a liquid nitrogen cooled MCT detector. All the liquid flows were pumped by a peristaltic pump (Verderflex) downstream of the ATR-IR cell to prevent for-mation of gas bubbles in the cell.

Once the cell was assembled in the IR spectrometer, it was flushed with Ar/H2O with a flow rate of 0.5 mL/min until a stable

water spectrum was obtained. Once the water spectrum was stable, the background spectrum was collected. Subsequently, the cell with the catalyst layer was flushed with formic-acid solution

(10 mM) at pH 5 or pH 3, degassed with Ar before the experiment for at least 3 h. The catalysts in the ATR cell was exposed to the liq-uid flow during 10 min, where after the liqliq-uid flow was stopped, mimicking a batch reactor. ATR-IR spectra were recorded at room temperature (20 ± 1°C) in an air-conditioned room. Each spectrum was acquired by averaging of 128 scans taken with a resolution of 4 cm1. The collected spectra were averaged over 60 s. The catalyst layers were reused a few time and comparable results were obtained, indicating that the catalyst layer is stable during the experiments.

3. Results

3.1. Catalyst characterization

Table 2reports the properties of the Pd/

c

-Al2O3catalyst. The

high surface area

c

-Al2O3support assisted to achieve high metal

dispersion, as determined with CO-chemisorption. XPS measure-ments showed that the sample stored in ambient conditions con-tains 30% of the Pd in oxidized state (shown in Figure S1), in reasonable agreement with literature[19,39].

3.2. Formic-acid decomposition under inert atmosphere

Fig. 1 shows a typical result of a formic-acid decomposition experiment, showing the decreasing concentration of formic-acid as well as the concentrations of H2and CO2in the gas-stream

dur-ing the experiment. CO was never detected in any experiment in this study. Note that the differences in the shape of the profiles are caused by the fact that the liquid phase can be considered as a batch reactor, the formic-acid concentration is converted with a decreasing rate in time. In contrast, the gas phase products are measured in the gas stream passing through the reactor and the low concentrations observed in the first half hour are due to the fact that the concentrations of H2and CO2in the gas cap have to

build up first. Consequently, information on gas-phase products is delayed. Nevertheless, the amount of H2and CO2produced and

removed from the reactor at a certain time can be calculated by the integration over time of the flow rate of each product, which is based on its concentration profile and the Ar flow rate, as is shown inFig. 2a.

Fig. 2a shows the amount of formic-acid converted (based on the formic-acid concentration in Fig. 1) and the amounts of H2

and CO2produced (calculated as explained above).Fig. 2b shows

in more detail the same for the initial phase of the experiment. The results clearly show that the reaction rate is constant during the first half our, whereas the amounts of gas-phase products show a clear delay, as discussed above. Therefore, the initial turnover fre-quencies are calculated based on conversion of formic-acid at low conversion. Furthermore, the reaction becomes extremely slow after typically one hour, as can be seen inFig. 2a where both pro-files for the consumption of formic-acid and the cumulative amount of gas-phase products are flattened. Additionally, H2and

CO2 are produced in equal amount, in agreement with the

stoi-chiometry of the decomposition reaction. Also, the number of

Table 2

Catalyst characterization.

Catalyst Pd/c-Al2O3

Catalyst specific surface area, m2

/g 195

Metal loading, wt% 0.9

CO uptake,mmol/g 31.3

Metal dispersion, % 38

Mean metal particle size, nm 2.8

(4)

moles formic-acid (0.5 mmol) converted after 3 h agree well with the amount of H2and CO2formed.

The results are not influenced by internal mass transfer accord-ing to the calculation of the Weisz Prater number as shown in the supporting information section internal mass transfer, as the esti-mated value is much lower than 1. However, we cannot rule out any effects of formation of bubbles in the catalyst pores, which might both slow down or enhance internal transport, the latter by causing chaotic movement of the liquid in the catalyst pore according to the ‘‘oscillation theory” [49–51]. Also, any influence of external mass transfer can be ruled out based on the observation that the stirring rate (above 375 rpm) has no influence on the reac-tion rate, as well as based on calculareac-tions, as shown in supporting information section on external mass transfer.

3.3. Effect of oxygen

Figure S2 in supporting information shows the conversion of formic-acid in time for experiments with oxygen concentrations between 0 and 2 vol%. Clearly, the presence of oxygen enhances conversion. Fig. 3a shows more clearly the effect of oxygen at low concentrations, i.e. below 0.1%.Fig. 3b shows that the initial

apparent TOF increases with oxygen concentration between 0 and 2 vol% with almost a factor 3. However, the initial TOF is con-stant within experimental error at O2concentration below 0.1 vol

%, in agreement with the fact that conversion seems independent of oxygen concentration during the first 30 min inFig. 3a.

Fig. 4a shows the formic-acid conversion, and H2and CO2yields

after three hours of reaction, calculated by integrating the H2and

CO2 concentrations in gas-phase as explained in section 3.2 and

based on the final conversions shown in Figures 3 and S2.Fig. 4a shows that the hydrogen yield is maximal at 0.1 vol%. In contrast, the CO2yield increases with oxygen concentration until

formic-acid is completely converted within 3 h. A similar trend in H2

and CO2yield is also observed after 30 min reaction time as shown

in Figure S3.Fig. 4a also shows that the CO2yield is always higher

than the H2yield; the difference clearly increases with increasing

oxygen concentration as shown inFig. 4b. Figure S4 shows how that the ratio between H2yield and CO2yield decreases with

reac-tion time in presence of 0.1% O2, suggesting that the contribution of

H2oxidation increases with time.

The mass balance inFig. 4a is not completely closed and about 10% of the carbon seems lost, caused by the delay in the gas phase analysis data as discussed inSection 3.2. This is illustrated by Fig-ure S5, showing that when formic-acid is completely decomposed, i.e. no further H2and CO2are produced, detection of H2and

espe-cially CO2in the gas phase continues for more than two hours. This

is caused by slowly flushing out the gasses and the effect is much stronger for CO2because of the relatively high solubility of CO2in

water. The accuracy of the carbon mass balance is probably most affected by CO2remaining in the water in the reactor.

Fig. 5shows the formic-acid conversion profiles of an experi-ment in inert with an experiexperi-ment under semi-inert conditions, i.e. by flushing only the gas-cap with inert gas without bubbling through the liquid. The red dots show 17% conversion in three hours under inert-experiment condition. In contrast, a much higher conversion (68%) is obtained under semi-inert conditions. The difference is caused by oxygen dissolved in the water as it is not removed with Ar under semi-inert conditions.

3.4. Effect of formic-acid concentration

Fig. 6shows that the initial apparent TOF, in presence of 0.1 vol % oxygen, remains almost constant on changing the formic-acid concentration. Note that varying the initial formic-acid concentra-tion between 2 and 10 mM also changed the initial soluconcentra-tion pH Fig. 1. H2and CO2volume percent in the composition gas analyzed by a micro-GC

sampling every 5 min and formic-acid concentration profile (5 mM formic-acid, 50 mL/min NTP Ar flow through, 100 mg catalyst).

Fig. 2. a) Formic-acid, H2and CO2content profile during the reaction, b) the zoom-inFig. 2a the initial 20% conversion of formic-acid data (5 mM formic-acid, 50 mL/min NTP

(5)

between 3.2 and 2.9[52]. As shown in Figure S6, the rate of reac-tion decreases with increasing pH, but the change within the pH window between 2.9 and 3.2 appears smaller than experimental accuracy. Therefore, it can be concluded that the apparent order in formic-acid is zero under our experimental conditions. Different reaction orders in formic-acid are reported in literature[20,23,53]

but unfortunately the oxygen concentration was not well con-trolled in these studies. The importance of oxygen will be dis-cussed later.

3.5. Catalyst stability

Catalyst stability was tested in both inert (Ar) as well as in the presence of 0.1 vol% O2. A single batch of the catalyst was tested

during four hours by adding the same amount of the formic-acid every 60 min to the batch reactor.Fig. 7a shows the results under inert atmosphere. Clearly, formic-acid is almost fully converted in the first hour. The activity decreased significantly in the second hour after dosing additional formic-acid solution, and even more so after dosing for the third and fourth time. The catalyst clearly deactivates under inert condition. On the contrary, in the same experiments performed in 0.1 vol% oxygen, as shown in Fig. 7b, the catalyst maintained its activity in three runs, only showing mild deactivation in the fourth run.

Fig. 8 shows the formic-acid conversion profile after pretreat-ment of the catalysts with CO and standard inert experipretreat-ment. Clearly, pretreatment with CO causes complete deactivation, which is confirmed by the observation that no gas products (H2, CO2, CO)

could be detected with GC.

3.6. ATR experiments on formic-acid decomposition

Fig. 9presents the ATR-IR spectra obtained with bare ZnSe, a bare alumina layer and a catalyst layer, exposed to formate solu-tions for at least 10 min at pH 3 and 5. The dark yellow line shows the spectrum on bare ZnSe at pH 5, showing three peaks at respec-tively 1581, 1380 and 1350 cm1. These three peaks are also observed in the experiments with the Al2O3 layer and the Pd/

Al2O3layer at different pH condition and are assigned to free

for-mate in the bulk solution, in agreement with literature[18]. Note that the absolute intensities inFig. 9cannot be compared because the optical properties of the layers are different. The red line shows the spectrum obtained with Pd/Al2O3at pH 5, showing an

addi-tional peak at 2350 cm1which is assigned to CO2[19,54–57],

con-firming the formation of CO2 via Pd-catalyzed formate

decomposition. The black spectrum presents the result of the same experiment at pH 3, revealing both a larger CO2peak at 2350 cm1

as well as a clear shoulder peak around 1610 cm1, which however Fig. 3. a) Formic-acid conversion profile at extreme low O2concentration (0–0.1 vol%), b) the initial apparent TOF plot with the O2concentration (0–2 vol%) applied (5 mM

formic-acid, 50 mL/min NTP Ar/O2flow through, 100 mg catalyst).

Fig. 4. a) Formic-acid conversion, integrated yields of H2and CO2under different oxygen concentration after 3 h reaction, b) H2to CO2ratio plot with the O2concentration

(6)

cannot be assigned at this time and might be an artifact caused by small changes in the background spectrum of water.

Fig. 10a shows how the peaks in ATR-IR spectra develop during exposure to formic-acid (pH = 3) for 5 min, finally resulting in the spectrum shown inFig. 9.Fig. 10b shows a zoom-in of the window between 1700 and 2200 cm1, revealing two peaks at 2110 cm1 and 1830 cm1, assigned to linear bonded and bridged bonded CO

on the Pd surface [30], demonstrating the formation of adsorbed CO. The same experiment at pH 5 (Figure S7c) does not result in detection of CO during exposure to formic acid.

4. Discussion

4.1. Catalyst characterization

A well-dispersed supported Pd catalyst was obtained with a mean metal particle size of 2.8 nm. The small Pd particles oxidized easily in air as observed with XPS (Table 2). The catalysts were

used as prepared without any further pretreatment before the reaction, thus Pd and PdO co-exist initially. It is well known Pd2+

can be reduced to Pd0 metal state by formic-acid[14,35,58,59]. According to recent studies[25,40], in-situ reduction with H2does

not influence catalyst activity. In any case, any formic-acid con-sumption due to PdO reduction in our experiments can maximally convert 0.2% of the initial amount of formic-acid. So, this would not influence the catalyst activity.

4.2. Activity in inert atmosphere

As shown inFig. 1, the catalyst is deactivated during the reac-tion when oxygen is absent, as only a small fracreac-tion of formic-acid is converted and the reaction rate is very small after two hours. The same is also observed in the experiment with repeat-edly dosing of formic acid inFig. 7a. Furthermore, the result shown in Fig. 8, obtained after pretreatment of the catalyst with CO, clearly demonstrates that CO deactivates the catalyst significantly, in agreement with claims in literature [18,20,23–25,34,38,60]. However, CO was never detected with GC, as CO absorbs very strongly on Pd and cannot desorb at room temperature. The amount of CO required to completely cover the Pd surface is as small as 3.2 * 103mmol CO, which can be produced by dehydra-tion (Eq.(2)) of only 0.2% of the initial amount of formic-acid.

Formation of adsorbed CO during formic acid decomposition is confirmed by ATR-IR experiments inFig. 10b. This agrees well with Jiang et al.[18], reporting similar results with ATR-IR. The peaks are slightly blue-shifted compared to our previous study[30], pos-sibly caused by differences in pH, the surface coverage of CO or interaction with formate ions. Clearly, in-situ characterization is required to detect CO as ex-situ IR spectroscopy failed as reported by Hu et al.[25], probably caused by oxidation of adsorbed CO in air.

In previous studies, TOFs were calculated based on the produc-tion rate of H2[14,23,31,53,61]or total gas production rate (H2and

CO2) [25,36,37], resulting in TOFs typically between 100 and

1000 h1. Unfortunately, the level of conversion is not reported, making direct comparison impossible. On top of that, the presence of oxygen was not clearly reported, which is important as will be discussed below.

Table 3summarizes the results of studies performed in well-controlled inert atmosphere. Clearly, the initial apparent TOF in our work is in the same order of magnitude or somewhat larger as reported in literature[42,62–64], despite differences in precise conditions and uncertainty about any undesired effect of mass transfer, except for our data. The fact that high TOF is observed at low formic-acid concentration is in line with the order zero in formic-acid (Fig. 6), also confirming that formic-acid mass transfer cannot be limiting in our experiments.

4.3. Influence of pH

The activity of the Pd catalyst decreases with increasing pH, independent of the oxygen concentration as shown Figure S6 in presence of 0.1 vol% O2and under inert conditions. This agrees well

with the observations in ATR-IR experiments that both CO2and

adsorbed CO form faster at pH 3 (Fig. 10) than at pH 5 (Figure S7). The results suggest that un-dissociated formic acid rather than for-mate ions react, but this observation can also be interpreted in terms of the electrochemical potential of the Pd particles, which decreases with increasing pH.

Many reports in literature[20,36,39]report on the influence of the ratio of formic-acid and sodium-formate in the reaction mix-ture, without considering that the actual concentration of formate-ions is determined by the acid-base equilibrium of the Fig. 5. The conversion of 5 mM formic-acid with 50 mg 1 wt% Pd/c-Al2O3in

presence of different amounts of traces of oxygen in the reaction system at the start of the batch experiment (5 mM formic-acid, 50 mg catalyst, inert-experiment: Ar gas flow through liquid 50 mL/min NTP, semi-inert-experiment: Ar gas flow above liquid 50 mL/min NTP).

Fig. 6. The initial apparent TOF plots with different formic-acid concentration (2– 10 mM) under 0.1 vol% oxygen concentration (50 mL/min NTP Ar/O2flow through,

(7)

dissociation of formic-acid (Eq.(8)) as also argued for the first time in[53].

HCOOH$ HCOOþ Hþ ð8Þ

The consequence is that the reactant concentration, i.e. the sum of formate and formic acid, is not constant so that an unambiguous observation on the effect of pH on the reaction rate is not obtained. Nevertheless, the qualitative observation that basic solutions result in very low reaction rates, is in agreement with observations in lit-erature[32,36,39,53,65].

4.4. Activity in presence of oxygen

Fig. 3b shows that the apparent initial TOF increases with increasing oxygen concentration, although the effect is insignifi-cant when varying the oxygen concentration in a narrow window between 0 and 0.1 vol%. Fig. 3a shows that the catalyst activity remains higher during the experiment on increasing the oxygen concentration, achieving much higher conversion. In other words, catalyst deactivation is suppressed by oxygen, even if the oxygen concentration is very low, i.e. below 0.1 vol%. The result in

Fig. 7b confirms that catalyst stability is significantly improved

by introducing trace amounts of oxygen. However, it is also clear that deactivation occurs in the fourth run, suggesting that a second deactivation mechanism is in operation. In any case, low oxygen concentration improves stability of Pd catalysts for formic-acid decomposition.

Oxygen may be involved in three reactions as presented in

Scheme 1, all leading to formation of H2O and CO2. Firstly, oxygen

may react with CO to form CO2(step 4), termed as CO oxidation.

Secondly oxygen may oxidize H2to H2O (step 3), termed as H2

oxi-dation, which are both consecutive reactions. Thirdly, formic-acid may react directly (step 6) with oxygen dissociated on the Pd sur-face (step 5), termed as formic-acid oxidation.

The effect of small concentrations of oxygen on catalyst perfor-mance is attributed to step 4, decreasing the CO coverage of the Pd surface and suppressing or even preventing deactivation, as is argued below.

In principle, the conversion rate of formic-acid could also be boosted via the reaction in step 5 and 6, i.e. direct deep oxidation Fig. 7. Pd catalyst stability for formic-acid decomposition under a) inert gas (50 mL/min NTP Ar flow through); b) 0.1 vol% O2(50 mL/min NTP Ar/O2flow through);15mL pure

formic-acid solution was added to the batch reactor every 60 min.

Fig. 8. Formic-acid conversion profile after pretreatment under CO and inert conditions (5 mM formic-acid, 50 mL/min NTP flow through, 100 mg catalyst).

Fig. 9. ATR-IR spectra after exposure to formate solutions: Dark yellow represents formate at pH 5 spectrum on bare ZnSe (the intensity was multiplied by five), blue line represents formate at pH 5 spectrum on Al2O3layer, red line and black line

represent formate spectrum on Pd/Al2O3layer at pH 5 and pH 3, respectively (5 mg

(8)

of formic-acid. There are three arguments against this proposition. Firstly, if steps 5 and 6 would take over completely after one hour in the experiment with 0.1 vol% O2(seeFig. 3a), i.e. forming CO2

and H2O exclusively during the second and third hour, the H2/

CO2ratio would decrease from 0.73 (Figure S4) after 1 h to 0.32

after 3 h. However, the observed H2/CO2ratio after 3 h is 0.58

(Fig-ure S4), implying that H2formation via reaction(1)continued

dur-ing the whole experiment. Note that the H2/CO2ratio is calculated

based on the integral amounts of the H2and CO2produced.

Sec-ondly, the H2yield increases with increasing oxygen concentration

below 0.1 vol% as can be seen inFig. 4a and S3b, implying that H2

formation rate increases over the whole experiment. Thirdly, the presence of a very small amount of O2in the reactor, i.e. by

remov-ing oxygen only in the gas cap before the experiment and leavremov-ing dissolved oxygen in water behind (Fig. 5), increases the conversion after 3 h from 17% (in inert conditions) to 68%. Stoichiometric reac-tion of O2dissolved in water in equilibrium with air could account

for an increase in conversion of maximal 11%. The real contribution is even smaller because the oxygen concentration in water will decrease somewhat during flushing the gas cap with inert gas. Fig. 10. a) ATR-IR spectra of formic-acid decomposition at pH = 3 flowing for 5 min, b) zoom-in a) in the window of 1700 and 2300 cm1(5 mg 1 wt% catalyst on ZnSe, 10 mM formic-acid solution, 0.5 mL/min flow rate).

Table 3

Literature data on rate of formic-acid decomposition over supported Pd catalysts in inert atmosphere.

Catalyst Temperature Formic-acid concentration pH Pd metal size Support size TOF

Pd/C[42] 20–25°C 1000 mM 1.88 1.9–3.2 nm N.A. 60–100 h1

Pd/H-BETA(0.5)[62] 50°C 1000 mM 1.88 3.4 nm N.A. 59.2 h1

Pd/SiO2[63] 30°C 1000 mM 1.88 3.5 nm N.A. 23.4 h1

Pd/g-C3N4[64] 30°C 1000 mM 1.88 4.6 nm N.A. 35 h1

This work 20°C 5 mM 3 2.8 nm Below 20mm 230 h1

(9)

These three observations demonstrate that preventing deactiva-tion via removal of adsorbed CO by oxidadeactiva-tion is the dominant mechanism, whereas any effect via direct oxidation of formic-acid is minor.

It should be noted that no distinction is possible between ‘‘di-rect oxidation of formic-acid (5 and 6)” and the combination of ‘‘dehydrogenation followed by H2 oxidation (1 and 3)” and

‘‘de-composition followed by CO oxidation (2 and 4)”. Therefore, the same arguments discussed above could also be used to show that CO oxidation dominated over H2 oxidation at O2 concentration

up to 0.1%.

On the other hand, the situation is quite different at higher O2

concentrations, where step 3 and/or 6 clearly dominate, decreasing the H2yield, as shown inFig. 4a. It is clear that oxygen not only

suppresses deactivation, but also influences the rate of conversion of formic-acid by opening an additional reaction pathway, influ-encing also the product distribution.

There are three types of experiments reported in literature for studying formic-acid decomposition. First, in many studies the experiments were conducted in air, implying that oxygen is pre-sent in both the gas above the solution as well as in the solution

[16,18,20,23,25,31–37,39,40,60,61,66–69]. Second, in some studies the air above the solution was removed by flushing with inert before the reaction was initiated, implying that oxygen dissolved in the liquid might still be present[19,53,70–72]. This is the same, somewhat poorly defined, situation as we applied inFig. 5, demon-strating a strong influence on the reaction rate. Third, in a few studies air was completely removed from the reactor including the solution, which is the standard method in this study

[24,41,42,62–64,73,74]. The results confirm that Pd catalysts deac-tivate [24,41,64,73], in agreement with our results in inert. Our study shows not only that presence of oxygen at low concentra-tion, i.e. 0.1%, suppresses catalyst deactivaconcentra-tion, but also that the observed reaction rates are strongly influenced by the presence of oxygen in the concentrations window between 0.2 and 2%. Therefore, the effect of oxygen should be rigorously considered in future research on efficient catalysts for formic-acid decomposition.

From a practical point of view and considering formic-acid as a possible way to store hydrogen, efficient recovery of hydrogen by formic-acid decomposition is important. This efficiency can be cal-culated based on the integral H2/CO2ratio, as CO2is the only

C-containing product detected and the amount of chemisorbed CO can be neglected. The data in Fig. 4b show that the efficiency is 90% in the absence of any oxygen, whereas the apparent loss of 10% is assigned to CO2staying behind in the reactor dissolved in

water, as discussed in Section 3.3. In any case, the efficiency decreases to 58% in the presence of 0.1% of oxygen. Consequently, addition of traces of oxygen can prevent deactivation, but only with the disadvantage that efficiency decreases significantly. Fur-ther research would be required to solve this problem.

5. Conclusion

The kinetics of formic-acid decomposition over Pd catalyst sup-ported on alumina is strongly influenced by deactivation during the batch experiment, dominantly caused by CO poisoning. Deacti-vation can be suppressed by dosing trace amounts of oxygen. How-ever, oxygen reacts not only with adsorbed CO, preventing deactivation, but also with H2simultaneously. Operation at oxygen

concentrations below 0.1 vol% enhances the production of hydro-gen, as efficient prolonging catalyst activity dominates over con-secutive oxidation of hydrogen. Furthermore, oxygen concentrations between 0.1 vol% and 2 vol% cause significant increase in the rate of conversion of formic-acid and influences also

the product distribution, i.e. decreasing the hydrogen yield, which in many cases is not accounted for in reported batch experiments on Pd catalyzed decomposition of formic-acid in literature. Declaration of Competing Interest

The authors declare that they have no known competing finan-cial interests or personal relationships that could have appeared to influence the work reported in this paper.

Acknowledgment

The authors gratefully acknowledge financial support from China Scholarship Council and the National Mexican Council of Science and Technology (CONACYT) grant number 329977. We are grateful to K. Altena–Schildkamp and T.M.L Velthuizen for chemical analysis. We thanks Yang Wang from Inorganic Materials Science group University of Twente for the XPS measurement. We acknowledge B. Geerdink for technical support.

Appendix A. Supplementary data

Supplementary data to this article can be found online at

https://doi.org/10.1016/j.jcat.2020.10.032. References

[1]W. Wang, J.M. Herreros, A. Tsolakis, A.P.E. York, Ammonia as hydrogen carrier for transportation; investigation of the ammonia exhaust gas fuel reforming, Int. J. Hydrogen Energy. 38 (2013) 9907–9917.

[2] M. Bertau, H. Offermanns, L. Plass, F. Schmidt, H.J. Wernicke, Methanol: The basic chemical and energy feedstock of the future: Asinger’s vision today, 2014.

[3]E. Rivard, M. Trudeau, K. Zaghib, Hydrogen storage for mobility: a review, Materials (Basel) 12 (2019).

[4] U. Bossel, B. Eliasson, Energy and Hydrogen Economy, Eur. Fuel Cell Forum, Lucerne. (2002) 36. https://afdc.energy.gov/files/pdfs/hyd_economy_bossel_ eliasson.pdf.

[5] R. van Putten, T. Wissink, T. Swinkels, E.A. Pidko, Fuelling the hydrogen economy: Scale-up of an integrated formic acid-to-power system, Int. J. Hydrogen Energy. (2019).

[6]K. Grubel, H. Jeong, C.W. Yoon, T. Autrey, Challenges and opportunities for using formate to store, transport, and use hydrogen, J. Energy Chem. 41 (2020) 216–224.

[7]D.A. Bulushev, J.R.H. Ross, Heterogeneous catalysts for hydrogenation of CO2 and bicarbonates to formic acid and formates, Catal. Rev. - Sci. Eng. 60 (2018) 566–593.

[8]M. Grasemann, G. Laurenczy, Formic acid as a hydrogen source - recent developments and future trends, Energy Environ. Sci. 5 (2012) 8171–8181. [9]M. Yadav, Q. Xu, Liquid-phase chemical hydrogen storage materials, Energy

Environ. Sci. 5 (2012) 9698–9725.

[10]A. Boddien, H. Junge, Catalysis: acidic ideas for hydrogen storage, Nat. Nanotechnol. 6 (2011) 265–266.

[11]D.A. Bulushev, M. Zacharska, Y. Guo, S. Beloshapkin, A. Simakov, CO-free hydrogen production from decomposition of formic acid over Au/Al2O3 catalysts doped with potassium ions, Catal. Commun. 92 (2017) 86–89. [12]M. Sridhar, D. Ferri, M. Elsener, J.A. Van Bokhoven, O. Kröcher, Promotion of

ammonium formate and formic acid decomposition over Au/TiO<inf>2</inf> by support basicity under SCR-relevant conditions, ACS Catal. 5 (2015) 4772– 4782.

[13]F. Sánchez, D. Motta, N. Dimitratos, Catalytic decomposition of carbon-based liquid-phase chemical hydrogen storage materials for hydrogen generation under mild conditions, Appl. Petrochem. Res. 6 (2016) 269–277.

[14]Z.L. Wang, J.M. Yan, H.L. Wang, Y. Ping, Q. Jiang, Pd/C synthesized with citric acid: an efficient catalyst for hydrogen generation from formic acid/sodium formate, Sci. Rep. 2 (2012) 598.

[15]W.Y. Yu, G.M. Mullen, D.W. Flaherty, C.B. Mullins, Selective hydrogen production from formic acid decomposition on Pd-Au bimetallic surfaces, J. Am. Chem. Soc. 136 (2014) 11070–11078.

[16]Z. Dong, F. Li, Q. He, X. Xiao, M. Chen, C. Wang, X. Fan, L. Chen, PdCoNi nanoparticles supported on nitrogen-doped porous carbon nanosheets for room temperature dehydrogenation of formic acid, Int. J. Hydrogen Energy 44 (2019) 11675–11683.

[17]Y. Wang, Y. Qi, D. Zhang, C. Liu, New insight into the decomposition mechanism of formic acid on pd(111): competing formation of co2 and co, J. Phys. Chem. C 118 (2014) 2067–2076.

(10)

[18] K. Jiang, K. Xu, S. Zou, W. Bin Cai, B-doped pd catalyst: Boosting room-temperature hydrogen production from formic acid-formate solutions, J. Am. Chem. Soc. 136 (2014) 4861–4864.

[19]Q. Lv, Q. Meng, W. Liu, N. Sun, K. Jiang, L. Ma, Z. Peng, W. Cai, C. Liu, J. Ge, L. Liu, W. Xing, Pd-PdO interface as active site for HCOOH selective dehydrogenation at ambient condition, J. Phys. Chem. C 122 (2018) 2081–2088.

[20] M. Yurderi, A. Bulut, M. Zahmakiran, M. Kaya, Carbon supported trimetallic PdNiAg nanoparticles as highly active, selective and reusable catalyst in the formic acid decomposition, Appl. Catal. B Environ. 160–161 (2014) 514–524. [21]H. Alamgholiloo, S. Zhang, A. Ahadi, S. Rostamnia, R. Banaei, Z. Li, X. Liu, M.

Shokouhimehr, Synthesis of bimetallic 4-PySI-Pd@Cu(BDC) via open metal site Cu-MOF: Effect of metal and support of Pd@Cu-MOFs in H2 generation from formic acid, Mol. Catal. 467 (2019) 30–37.

[22]T. Feng, J.M. Wang, S.T. Gao, C. Feng, N.Z. Shang, C. Wang, X.L. Li, Covalent triazine frameworks supported CoPd nanoparticles for boosting hydrogen generation from formic acid, Appl. Surf. Sci. 469 (2019) 431–436.

[23]K. Tedsree, T. Li, S. Jones, C.W.A. Chan, K.M.K. Yu, P.A.J. Bagot, E.A. Marquis, G.D. W. Smith, S.C.E. Tsang, Hydrogen production from formic acid decomposition at room temperature using a Ag-Pd core-shell nanocatalyst, Nat. Nanotechnol. 6 (2011) 302–307.

[24]D.M. Ruthven, R.S. Upadhye, The catalytic decomposition of aqueous formic acid over suspended palladium catalysts, J. Catal. 21 (1971) 39–47. [25]C. Hu, J.K. Pulleri, S.W. Ting, K.Y. Chan, Activity of Pd/C for hydrogen generation

in aqueous formic acid solution, Int. J. Hydrogen Energy 39 (2014) 381–390. [26]M. Navlani-García, D. Salinas-Torres, D. Cazorla-Amorós, Hydrogen production

from formic acid attained by bimetallic heterogeneous pdag catalytic systems, Energies 12 (2019).

[27]M. Peter, S. Adamovsky, J.M. Flores Camacho, S. Schauermann, Energetics of elementary reaction steps relevant for CO oxidation: CO and O2 adsorption on model Pd nanoparticles and Pd(111), Faraday Discuss. 162 (2013) 341–354. [28]M. Lischka, A. Groß, Hydrogen on palladium: a model system for the

interaction of atoms and molecules with metal surfaces, Res. Signpost. 661 (2003) 111–132.

[29]F. Solymosi, A. Berkó, Adsorption of CO2 on clean and potassium-covered Pd (100) surfaces, J. Catal. 101 (1986) 458–472.

[30] S.D. Ebbesen, B.L. Mojet, L. Lefferts, The influence of water and pH on adsorption and oxidation of CO on Pd/Al2O3-an investigation by attenuated total reflection infrared spectroscopy, Phys. Chem. Chem. Phys. 11 (2009) 641– 649.

[31]X. Zhou, Y. Huang, C. Liu, J. Liao, T. Lu, W. Xing, Available hydrogen from formic acid decomposed by rare earth elements promoted Pd-Au/C catalysts at low temperature, ChemSusChem 3 (2010) 1379–1382.

[32]L. Yang, X. Hua, J. Su, W. Luo, S. Chen, G. Cheng, Highly efficient hydrogen generation from formic acid-sodium formate over monodisperse AgPd nanoparticles at room temperature, Appl. Catal. B Environ. 168–169 (2015) 423–428.

[33]J. Liu, L. Lan, R. Li, X. Liu, C. Wu, Agglomerated Ag-Pd catalyst with performance for hydrogen generation from formic acid at room temperature, Int. J. Hydrogen Energy. 41 (2016) 951–958.

[34] F. Sanchez, M.H. Alotaibi, D. Motta, C.E. Chan-Thaw, A. Rakotomahevitra, T. Tabanelli, A. Roldan, C. Hammond, Q. He, T. Davies, A. Villa, N. Dimitratos, Hydrogen production from formic acid decomposition in the liquid phase using Pd nanoparticles supported on CNFs with different surface properties, Sustain. Energy Fuels. (2018).

[35]S. Zhang, B. Jiang, K. Jiang, W. Bin Cai, Surfactant-free synthesis of carbon-supported palladium nanoparticles and size-dependent hydrogen production from formic acid-formate solution, ACS Appl. Mater. Interfaces 9 (2017) 24678–24687.

[36]Q. Zhu, N. Tsumori, Q. Xu, Immobilizing extremely catalytically active palladium nanoparticles to carbon nanospheres: a weakly-capping growth approach, J. Am. Chem. Soc. 137 (2015) 11743–11748.

[37]S.J. Li, Y. Ping, J.M. Yan, H.L. Wang, M. Wu, Q. Jiang, Facile synthesis of AgAuPd/graphene with high performance for hydrogen generation from formic acid, J. Mater. Chem. A 3 (2015) 14535–14538.

[38]Q.Y. Bi, J.D. Lin, Y.M. Liu, H.Y. He, F.Q. Huang, Y. Cao, Dehydrogenation of formic acid at room temperature: boosting palladium nanoparticle efficiency by coupling with pyridinic-nitrogen-doped carbon, Angew. Chemie - Int. Ed. 55 (2016) 11849–11853.

[39]J. Li, W. Chen, H. Zhao, X. Zheng, L. Wu, H. Pan, J. Zhu, Y. Chen, J. Lu, Size-dependent catalytic activity over carbon-supported palladium nanoparticles in dehydrogenation of formic acid, J. Catal. 352 (2017) 371–381.

[40] Y. Kim, D.H. Kim, Understanding the effect of Pd size on formic acid dehydrogenation via size-controlled Pd/C catalysts prepared by NaBH4 treatment, Appl. Catal. B Environ. 244 (2019) 684–693.

[41]K. Mandal, D. Bhattacharjee, S. Dasgupta, Synthesis of nanoporous PdAg nanoalloy for hydrogen generation from formic acid at room temperature, Int. J. Hydrogen Energy 40 (2015) 4786–4793.

[42]M. Navlani-García, K. Mori, A. Nozaki, Y. Kuwahara, H. Yamashita, Investigation of size sensitivity in the hydrogen production from formic acid over carbon-supported Pd nanoparticles, ChemistrySelect. 1 (2016) 1879– 1886.

[43]P. Xu, S. Agarwal, L. Lefferts, Mechanism of nitrite hydrogenation over Pd/c -Al2O3 according a rigorous kinetic study, J. Catal. 383 (2020) 124–134. [44]P. Xu, S. Agarwal, J.F. Albanese, L. Lefferts, Enhanced transport in

Gas-Liquid-Solid catalytic reaction by structured wetting properties: nitrite hydrogenation, Chem. Eng. Process. - Process Intensif. 148 (2020).

[45]S.D. Ebbesen, B.L. Mojet, L. Lefferts, In Situ attenuated total reflection infrared (ATR-IR) study of the adsorption of NO2-, NH2OH, and NH 4+ on Pd/Al2O3 and Pt/Al 2O3, Langmuir 24 (2008) 869–879.

[46]S.D. Ebbesen, B.L. Mojet, L. Lefferts, In situ ATR-IR study of nitrite hydrogenation over Pd/Al2O3, J. Catal. 256 (2008) 15–23.

[47]S.D. Ebbesen, B.L. Mojet, L. Lefferts, Mechanistic investigation of the heterogeneous hydrogenation of nitrite over Pt/Al2O3 by attenuated total reflection infrared spectroscopy, J. Phys. Chem. C 113 (2009) 2503–2511. [48]Y. Zhao, N. Koteswara Rao, L. Lefferts, Adsorbed species on Pd catalyst during

nitrite hydrogenation approaching complete conversion, J. Catal. 337 (2016) 102–110.

[49]L.B. Datsevich, Some theoretical aspects of catalyst behaviour in a catalyst particle at liquid (liquid-gas) reactions with gas production: oscillation motion in the catalyst pores, Appl. Catal. A Gen. 247 (2003) 101–111.

[50]L.B. Datsevich, Alternating motion of liquid in catalyst pores in a liquid/liquid-gas reaction with heat or liquid/liquid-gas production, Catal. Today 79–80 (2003) 341–348. [51]L.B. Datsevich, Oscillations in pores of a catalyst particle in exothermic liquid (liquid-gas) reactions: analysis of heat processes and their influence on chemical conversion, mass and heat transfer, Appl. Catal. A Gen. 250 (2003) 125–141.

[52] Zumdahl, Table of Acids with Ka and pKa Values, Lab Man. Zumdahl/Zumdahl’s Chem. (2003) 656.

[53]S. Akbayrak, Y. Tonbul, S. Özkar, Nanoceria supported palladium(0) nanoparticles: superb catalyst in dehydrogenation of formic acid at room temperature, Appl. Catal. B Environ. 206 (2017) 384–392.

[54]H.A. Bullen, S.A. Oehrle, A.F. Bennett, N.M. Taylor, H.A. Barton, Use of attenuated total reflectance fourier transform infrared spectroscopy to identify microbial metabolic products on carbonate mineral surfaces, Appl. Environ. Microbiol. 74 (2008) 4553–4559.

[55]J.Y. Wang, H.X. Zhang, K. Jiang, W. Bin Cai, From HCOOH to CO at Pd electrodes: a surface-enhanced infrared spectroscopy study, J. Am. Chem. Soc. 133 (2011) 14876–14879.

[56]M.F. Baruch, J.E. Pander, J.L. White, A.B. Bocarsly, Mechanistic insights into the reduction of CO<inf>2</inf> on tin electrodes using in situ ATR-IR spectroscopy, ACS Catal. 5 (2015) 3148–3156.

[57]J.E. Pander, M.F. Baruch, A.B. Bocarsly, Probing the mechanism of aqueous CO2Reduction on post-transition-metal electrodes using ATR-IR spectroelectrochemistry, ACS Catal. 6 (2016) 7824–7833.

[58]Q. Wang, Y. Wang, P. Guo, Q. Li, R. Ding, B. Wang, H. Li, J. Liu, X.S. Zhao, Formic acid-assisted synthesis of palladium nanocrystals and their electrocatalytic properties, Langmuir 30 (2014) 440–446.

[59]W.J. Lee, Y.J. Hwang, J. Kim, H. Jeong, C.W. Yoon, Pd 2+ -Initiated formic acid decomposition: plausible pathways for C-H activation of formate, ChemPhysChem 20 (2019) 1382–1391.

[60]M. Caiti, D. Padovan, C. Hammond, Continuous production of hydrogen from formic acid decomposition over heterogeneous nanoparticle catalysts: from batch to continuous flow, ACS Catal. 9 (2019) 9188–9198.

[61]Q.Y. Bi, J.D. Lin, Y.M. Liu, F.Q. Huang, Y. Cao, Promoted hydrogen generation from formic acid with amines using Au/ZrO2 catalyst, Int. J. Hydrogen Energy 41 (2016) 21193–21202.

[62]M. Navlani-García, M. Martis, D. Lozano-Castelló, D. Cazorla-Amorós, K. Mori, H. Yamashita, Investigation of Pd nanoparticles supported on zeolites for hydrogen production from formic acid dehydrogenation, Catal. Sci. Technol. 5 (2015) 364–371.

[63]J. García-Aguilar, M. Navlani-García, Á. Berenguer-Murcia, K. Mori, Y. Kuwahara, H. Yamashita, D. Cazorla-Amorós, Evolution of the PVP-Pd surface interaction in nanoparticles through the case study of formic acid decomposition, Langmuir 32 (2016) 12110–12118.

[64]Y. Wu, M. Wen, M. Navlani-García, Y. Kuwahara, K. Mori, H. Yamashita, Palladium nanoparticles supported on titanium-doped graphitic carbon nitride for formic acid dehydrogenation, Chem. - An Asian J. 12 (2017) 860– 867.

[65]Q.F. Deng, Z.F. Zhang, F.J. Cui, L.H. Jia, Highly dispersed Pd-MnOx nanoparticles supported on graphitic carbon nitride for hydrogen generation from formic acid-formate mixtures, Int. J. Hydrogen Energy. 42 (2017) 14865–14871. [66]M. Hattori, D. Shimamoto, H. Ago, M. Tsuji, AgPd@Pd/TiO2 nanocatalyst

synthesis by microwave heating in aqueous solution for efficient hydrogen production from formic acid, J. Mater. Chem. A 3 (2015) 10666–10670. [67]Y. Kim, J. Kim, D.H. Kim, Investigation on the enhanced catalytic activity of a

Ni-promoted Pd/C catalyst for formic acid dehydrogenation: effects of preparation methods and Ni/Pd ratios, RSC Adv. 8 (2018) 2441–2448. [68]T.Y. Ding, Z.G. Zhao, M.F. Ran, Y.Y. Yang, Superior activity of Pd nanoparticles

confined in carbon nanotubes for hydrogen production from formic acid decomposition at ambient temperature, J. Colloid Interface Sci. 538 (2019) 474–480.

[69]J. Sun, H. Qiu, W. Cao, H. Fu, H. Wan, Z. Xu, S. Zheng, Ultrafine Pd particles embedded in nitrogen-enriched mesoporous carbon for efficient H 2 production from formic acid decomposition, ACS Sustain. Chem. Eng. 7 (2019) 1963–1972.

[70]M.H. Jin, D. Oh, J.H. Park, C.B. Lee, S.W. Lee, J.S. Park, K.Y. Lee, D.W. Lee, Mesoporous silica supported Pd-MnOxcatalysts with excellent catalytic activity in room-temperature formic acid decomposition, Sci. Rep. 6 (2016) 33502.

[71]Y. Yu, X. Wang, C. Liu, F. Vladimir, J. Ge, W. Xing, Surface interaction between Pd and nitrogen derived from hyperbranched polyamide towards highly effective formic acid dehydrogenation, J. Energy Chem. 40 (2020) 212–216.

(11)

[72]M. Navlani-García, D. Salinas-Torres, K. Mori, A.F. Léonard, Y. Kuwahara, N. Job, H. Yamashita, Insights on palladium decorated nitrogen-doped carbon xerogels for the hydrogen production from formic acid, Catal. Today 324 (2019) 90–96.

[73]M. Navlani-García, K. Mori, A. Nozaki, Y. Kuwahara, H. Yamashita, Screening of carbon-supported PdAg nanoparticles in the hydrogen production from formic acid, Ind. Eng. Chem. Res. 55 (2016) 7612–7620.

[74]M. Wen, K. Mori, Y. Futamura, Y. Kuwahara, M. Navlani-García, T. An, H. Yamashita, PdAg nanoparticles within core-shell structured zeolitic imidazolate framework as a dual catalyst for formic acid-based hydrogen storage/production, Sci. Rep. 9 (2019) 1–10.

Referenties

GERELATEERDE DOCUMENTEN

lijk van grote oppervlakten te schat­ ten zodat deze methode aileen voor kleine oppervlakten in aanmerking komt en plaatsel ijk in aanvulling op de methode Tansley

“The cognitive linguistic view of emotions is capable of integrating a methodologically sound analysis of the linguistic richness and complexity of emotion language in a

In de groeve is de grens tussen de Hochheim Formatie (vroeger: Mittleren Cerithien-Schichten) en de Oppen- heim Formatie (vroeger: Obere Cerithien-Schichten, on- derste deel)

-Een positief punt voor wat betreft de uitslag van de enquête was verder dat er verschillende aardige aanknopingspunten werden aangedragen voor. artikelen die we in de toekomst

Most importantly, we report, for the first time, a mass-transport-limited formic acid oxidation rate on the Pd ML Pt(100) rotating electrode in perchloric acid, setting a

The oxidation activity of methanol was found to decrease with the particle size, as was also observed for the CO oxidation, while the formic acid oxidation rate increases with

The HEPA industry group and HEWP occupation group were found to have longer average weighted distances from the Sydney City Centre: 19 and 20km, respectively

Copyright and moral rights for the publications made accessible in the public portal are retained by the authors and/or other copyright owners and it is a condition of