You are on page 1of 12

Chemical Engineering Journal 431 (2022) 134310

Contents lists available at ScienceDirect

Chemical Engineering Journal


journal homepage: www.elsevier.com/locate/cej

High purity, self-sustained, pressurized hydrogen production from


ammonia in a catalytic membrane reactor
Jose L. Cerrillo a, 1, Natalia Morlanés a, 1, Shekhar R. Kulkarni a, Natalia Realpe a,
Adrian Ramírez a, Sai P. Katikaneni b, Stephen N. Paglieri b, Kunho Lee b, Aadesh Harale b,
Bandar Solami b, Aqil Jamal b, S. Mani Sarathy c, Pedro Castaño a, Jorge Gascon a, *
a
KAUST Catalysis Center (KCC), King Abdullah University of Science and Technology (KAUST), Thuwal 23955-6900, Saudi Arabia
b
Carbon Management R&D Division, Research and Development Center, Saudi Aramco, Dhahran 31311, Saudi Arabia
c
Clean Combustion Research Center, King Abdullah University of Science and Technology (KAUST), Thuwal 23955-6900, Saudi Arabia

A R T I C L E I N F O A B S T R A C T

Keywords: The combination of catalytic decomposition of ammonia and in situ separation of hydrogen holds great promise
Membrane reactor for the use of ammonia as a clean energy carrier. However, finding the optimal catalyst – membrane pair and
NH3 decomposition operation conditions have proved challenging. Here, we demonstrate that cobalt-based catalysts for ammonia
Pd-Au membrane
decomposition can be efficiently used together with a Pd-Au based membrane to produce high purity hydrogen at
High purity hydrogen
Pressurized hydrogen production
elevated pressure. Compared to a conventional packed bed reactor, the membrane reactor offers several oper­
ational advantages that result in energetic and economic benefits. The robustness and durability of the combined
system has been demonstrated for>1000 h on stream, yielding a very pure hydrogen stream (>99.97 % H2) and
recovery (>90 %). When considering the required hydrogen compression for storage/utilization and environ­
mental issues, the combined system offers the additional advantage of production of hydrogen at moderate
pressures along with full ammonia conversion. Altogether, our results demonstrate the possibility of deploying
high pressure (350 bar) hydrogen generators from ammonia with H2 efficiencies of circa 75% without any
external energy input and/or derived CO2 emissions.

capacity [12,13]. Already since its discovery by J. Priestley at 1774


1. Introduction [14,15] and its following industrial synthetic production in 1908 by the
famed Haber-Bosch process [16,17], ammonia is one of the most pro­
Over the last decades, the term ‘hydrogen economy’ has gained duced chemicals in the world in terms of volume [18,19]. This is because
increasing significance [1–3 4,5]. Hydrogen is an environmentally NH3 is an important raw material for the manufacture of different
friendly, COx-free, energy vector [6] with a high gravimetric energy products such as fertilizers, pharmaceuticals, plastics, refrigerants, nitric
storage density [7]. However, other properties such as its low energy acid or explosives, among others [9,18,20]. But not only its production is
volumetric density (1.5 kWh/L for liquid H2 and 0.8 kWh/L for gas H2) mature, the distribution and transport of ammonia through pipelines is
or its high flammability are not desirable characteristics for a fuel already available in countries such as USA, Russia or Ukraine, while
[8–10]. Overall, the scientific community agrees that hydrogen will be trucks, trains and ships are used worldwide for the distribution of this
key for the decarbonization of our current energy system, although very important commodity. For instance, the global maritime trade is 18⋅106
important challenges related to production, distribution and storage tones per year [21]. Furthermore, liquefaction of ammonia is cost-
need to be addressed [11]. affordable and possible through compression to 10 bar at room tem­
Ammonia is a promising energy carrier in the hydrogen economy. perature or by cooling to –33 ◦ C at atmospheric pressure. Additionally,
This chemical offers a couple of advantages in terms of storage potential: working with liquified ammonia provides a volumetric hydrogen den­
it has a volumetric energy density of 3 kWh/L and a 17.8 wt% hydrogen sity 45% higher than liquified hydrogen [8,22,23]. Lastly, this chemical

Abbreviations: WHSV, Weight hourly space velocity.


* Corresponding author at: King Abdullah University of Science and Technology, KAUST Catalysis Center (KCC), Advanced Catalytic Materials, Thuwal 23955-
6900, Saudi Arabia.
E-mail address: Jorge.gascon@kaust.edu.sa (J. Gascon).
1
These authors contribute equally to this work.

https://doi.org/10.1016/j.cej.2021.134310
Received 28 October 2021; Received in revised form 17 December 2021; Accepted 20 December 2021
Available online 24 December 2021
1385-8947/© 2021 The Authors. Published by Elsevier B.V. This is an open access article under the CC BY license (http://creativecommons.org/licenses/by/4.0/).
J.L. Cerrillo et al. Chemical Engineering Journal 431 (2022) 134310

Nomenclature

Da Damkohler number
Pe Peclet number
ΔHr ammonia decomposition reaction enthalpy (kWh/kg NH3)
ΔHvap ammonia evaporation enthalpy (kWh/kg NH3)
CpNH3 ammonia specific heat (kWh/K/kg NH3)
ΔT temperature difference (K)
LHVH2 hydrogen Lower heating value (kWh/kg H2)
LHVNH3 ammonia Lower heating value (kWh/kg H2)
MWH2 hydrogen molecular weight (kmol/kg H2)
MWNH3 ammonia molecular weight (kmol/kg NH3)
ηC ammonia cracker efficiency, (90–95%),
ηH2-rec hydrogen recovery (permeated hydrogen) from the
cracked ammonia, (85–95%),
ηheat-rec heat recovery from the retentate, (70–90%)

Fig. 1. Thermodynamic ammonia conversion versus temperature at


is less hazardous in terms of flammability and explosion limits than different pressures.
current fuels or even hydrogen [11,24]. In contrast, the toxicity of
ammonia is higher, being an important barrier for its implementation in H2-selective membranes with catalytic systems in packed-bed catalytic
mobile applications. Despite its toxicity, scientists argue that ammonia membrane reactors (PBCMR) can lead to overcoming equilibrium con­
risks are acceptable and similar to current fuels, [25] placing NH3 as one straints and speed up reaction kinetics [67,69,70]. The thin Pd layers
of the most important candidates for the decarbonization of the current used as H2 selective membranes present restrictions in terms of working
energy system [12,26]. temperature. At low temperature (T < 400 ◦ C), apart from the low
As stated above, ammonia can be directly used as a COx-free fuel, hydrogen permeation due to the limitation in the endothermic diffusion
either in fuel cells [8,27] or through conventional combustion [24,28]. of hydrogen through the membrane, the integrity of the membrane is
Main issues with the direct utilization are the high temperatures compromised due to hydrogen embrittlement. Meanwhile, at higher
required by ammonia fuel cells and the potential NOx emissions during temperatures T > 550 ◦ C, cracking at the ceramic–metal interface might
combustion. On the other hand, it can be also used as a hydrogen carrier occur due to a mistmatch in thermal expansion. In constrast, PdAu
to obtain carbon-free H2, which can be directly combusted or used in membranes offer improvements in terms of H2 permeation at low tem­
fuel cells [29,30]. Nowadays, different techniques such as plasma perature and higher resistance to embrittlement [71,72] In this spirit,
[31,32] or photo-driven catalysis [33,34] are investigated to obtain H2 we selected PdAu alloy membranes as the most adequate to be combined
from ammonia, although electrocatalytic and thermocatalytic processes with our recently reported barium promoted cobalt-cerium catalyst.
are the most studied. The electrolysis of ammonia is not viable presently [73]
due to insufficient energy efficiency of the process, formation of NOx, Apart from purity requirements, hydrogen should be supplied or
and NH3 slip during the electro-oxidation [23,35,36]. Hence, the ther­ stored at high pressure. Therefore, in order to avoid additional
mocatalytic decomposition of ammonia or ammonia cracking (Equation compression costs [74–76] and derived emissions (estimates are about
(1))is the most promising method. 6.0 kWh/kg for compression of H2 to 70 MPa, which leads to approxi­
2NH3 ↔ 3H2 + N2 (1) mately 1.3 kg of CO2/kg of hydrogen) [77], high pressure in the
permeate side of a PBCMR would be desirable.
Ammonia decomposition is an endothermic reaction thermody­ Here, we demonstrate the very efficient production of hydrogen by
namically limited at low temperatures. Accordingly, higher tempera­ combining our recently developed Co-based catalyst and a state-of-the-
tures than 400 ◦ C are needed to achieve equilibrium conversions above art H2-selective membrane. The membrane dramatically enhances con­
99% at atmospheric pressure, and even higher temperatures if the version over the catalyst, proving the ability of the PBCMR to both,
working pressure is higher (Fig. 1) [37,38]. accelerate the kinetics and overcome equilibrium constraints. A full
When it comes to catalyst design, ruthenium is the most active metal process optimization with the objective of minimizing energy re­
for the decomposition of ammonia [6,39–50]. However, the high cost quirements has been performed. The system offers the possibility of
and availability of this metal make its industrial implementation utilizing the energy of the hydrogen in the retentate to cover the thermal
doubtful [51]. Besides, when targeting 100% conversion to avoid any requirements for the process, developing in such a way a stand-alone
ammonia slip, thermodynamic limitations become more important than system with a H2 efficiency of ~ 78%. Furthermore, the possibility of
kinetics, making the use of Ru less desired (still almost 500 ◦ C are producing the H2-permeate stream at moderate pressure results in
needed for full ammonia decomposition at 1 bar [38,52,53]). Other enormous energy savings and prevents further efficiency losses, and also
authors have looked at more abundant and cost-effective elements and decreases CO2 emissions associated to the compression step. Last but not
their combinations (i.e. Co, Ni, Mn, Ir or Fe) supported in different least, we demonstrate the industrial potential of the system by running
materials [22,51,54–63]. the membrane reactor for over 1000 h.
In general, next to thermodynamic limitations, the considerable
endothermicity of the reaction,[6] the inhibitory effect of the produced 2. Result and discussion
hydrogen,[64] and the high purity H2 required for most applications
(higher than 99.97% [22]), they all point at the in situ removal of 2.1. Benchmarking the membrane reactor configuration
hydrogen as the most feasible strategy to efficiently decompose
ammonia.[65–68] Specifically, the combination of membranes and To elucidate the benefits of incorporating a H2-selective membrane
catalytic ammonia decomposition allows to separate high purity H2 in to a packed-bed catalytic reactor (PBCR), the catalytic activity was
situ. Various authors have already demonstrated that the combination of

2
J.L. Cerrillo et al. Chemical Engineering Journal 431 (2022) 134310

firstly evaluated in the membrane module with the permeate stream utilization of both components. In this line, we investigated the effect of
opened (PBCMR) and closed (PBCR). In the second case, the reaction space velocity over membrane reactor performance (Fig. 4). The flow
mixture goes exclusively through the retentate or catalytic chamber, rate of ammonia was modified in the range 50–300 N mL min− 1, varying
mimicking a PBCR. The evolution of ammonia conversion using both the weight hourly space velocity (WHSV) in the range 300–1800 ml gcat-
1 − 1
approaches (PBCR and PBCMR) at different temperatures, contact times h . At low WHSV, full conversion and maximum hydrogen recovery
and pressures was analyzed (Fig. 2). The comparison of both experi­ are achieved. Under these conditions, the membrane is capable of
ments at different conditions reveals a significant improvement (10–30 permeating the produced H2 until similar PH2 are reached at both both
%) in ammonia conversion when the PBCMR was employed, in agree­ sides of the membrane. In contrast, at WSHVs higher than 1000 ml gcat-
1 − 1
ment with previous studies [69,70]. The presence of hydrogen has an h , the membrane is not able to permeate all produced hydrogen,
inhibitory effect on the kinetics of the reaction according to the Temkin- resulting in recovery rates below the maximum value. At higher WSHVs,
Pyzhev mechanism [78,79]. Therefore, the removal of H2 from the full conversion cannot be achieved and some NH3 slip is observed in the
retentate chamber improves ammonia conversion. This effect can be retentate.
seen only at temperatures higher than 400 ◦ C, minimum temperature to Additional experiments at increasing pressure in the retentate were
attain H2 transport through the Pd-Au membrane (Supplementary performed. As shown in Fig. 5, upon increasing the pressure in the
Figure S5). retentate chamber and consequently the driving force for permeation of
The purity of the recovered hydrogen in all the PBCMR experiments H2 through the membrane, full ammonia conversion with much higher
was 99.97 ± 0.03 %, demonstrating the absence of defects in the recovery rates can be achieved. This is the result of a much more effi­
membrane (see Figures S6-S8). Minor N2 impurities can be observed at cient use of the membrane that, in addition, strongly suppresses the
pressures higher than 9 bar (Supplementary Table S2), in agreement well-known inhibition of H2 on kinetics and further brings the reaction
with previous studies [64,67]. Overall, ammonia conversion increases far from equilibrium limitations [53,64,67,81]. The positive effect of
when temperature (250 – 500 ◦ C), contact time (2.7 – 16.3 g h mol− 1) increasing the pressure on ammonia conversion and hydrogen re­
and pressure (4 to 15 bar) are increased. A similar trend can be observed coveries is particularly interesting at higher space velocities (Supple­
in both reactor configurations in the case of temperature and contact mentary Figure S9).
time. In contrast, higher pressures are only beneficial in case of the
PBCMR [40,70]. This negative effect in the PBCR, when pressure is 2.2. Production of pressurized H2
increased, is the result of the previously mentioned inhibitory effect of
hydrogen on the reaction kinetics and of the strong thermodynamic A sets of experiments were conducted at 485 ◦ C and 200 ml NH3
limitations. [80 78 80 81] Finally, the PBCMR without catalyst did not min− 1, keeping constant the pressure in the permeate chamber at 1, 3, 5,
present any significant ammonia conversion, demonstrating no catalytic 10 and 15 bar, and increasing the pressure in the retentate chamber from
activity of the membrane (Supplementary Table S1). 1 to 50 bar. To the best of our knowledge, this is the first example in the
open literature reporting the production of hydrogen from ammonia at
2.1.1. Optimizing operation in the PBCMR pressures higher than atmospheric. Similarly, we could not find any
Fig. 3 shows the effect of temperature in ammonia conversion, publication reporting experiments at >6 bar in the retentate.
hydrogen recovery, and composition of the permeate and retentate, as As observed in Fig. 6, the pressure ratio retentate/permeate even­
well as the measured flows, when pressure in the retentate is fixed to 4 tually determines hydrogen recovery and productivity. However,
bar (ΔP = 3 bar). As described above, both ammonia conversion and reaching full ammonia conversion at higher permeate pressures be­
hydrogen recovery increase with the reaction temperature, reaching comes more difficult and may lead to undesirable ammonia slip (Sup­
total conversion and almost 80% H2 recovery at 500 ◦ C. Under these plementary Figures S10-S12), together with the increase in the
conditions, pure hydrogen is obtained in the permeate. However, not all impurities in the hydrogen permeate stream (Supplementary Table S3).
produced hydrogen can permeate through the membrane after partial In addition, similar experiment were carried out at lower temperatures
pressures of this gas (PH2) are equalized at both sides of the membrane (450 ◦ C) demonstrating similar trends at lower conversion and H2 re­
(see maximum recovery, that would correspond to similar PH2 in both covery values (Supplementary Figure S14). These results however,
permeate and retentate). Indeed, in membrane separation processes, the demonstrate the great potential of this configuration and offer a great
flow of the permeable gas is a function of the driving force (expressed as degree of flexibility for the optimization of the full process (vide infra).
difference in chemical potential) across the membrane.
Matching the permeation properties of the membrane and the ac­
tivity of the catalyst is of primary importance to ensure the full

Fig. 2. Effect of the reaction conditions (T, W/F and P) using PBCR (blue) and PBCMR (red). 0.5%Ba-CoCe catalyst: 2 g (350–500 μm) diluted with 10 g of SiC ~ 680
μm. Fixed conditions are ammonia flow rate 100 N mL min− 1, T = 485 ◦ C , Pretentate = 4 bar and Ppermeate = 1 bar. In each graph one variable was varied: Temperature
250 – 500 ◦ C (left), ammonia flow rate 50–300 N mL min− 1 (center), and Pretentate = 4–15 bar (right). (For interpretation of the references to colour in this figure
legend, the reader is referred to the web version of this article.)

3
J.L. Cerrillo et al. Chemical Engineering Journal 431 (2022) 134310

Fig. 3. Effect of the reaction temperature in the ammonia conversion and hydrogen recovery (left), the molar composition and retentate flow (right-up) and molar
composition and permeate flow (right-down). Reaction conditions: 0.5%Ba-CoCe catalyst: 10 g (350–500 mm) diluted with 20 g of SiC ~ 680 mm. Ammonia flow
rate = 200 N mL min− 1, WHSV = 1200 ml gcat-1h− 1, T = 400–500 ◦ C, Pretentate = 4 bar; Ppermeate = 1 bar.

Fig. 4. Effect of the weight hourly space velocity in the ammonia conversion and hydrogen recovery (left), the molar composition and retentate flow (right-up) and
molar composition and permeate flow (right-down). Reaction conditions: 0.5%Ba-CoCe catalyst: 10 g (350–500 μm) diluted with 20 g of SiC ~ 680 μm. Ammonia
flow rate = 50–310 N mL min− 1, WHSV = 300–1860 ml gcat-1h− 1, T = 485 ◦ C, Pretentate = 4 bar; Ppermeate = 1 bar.

2.3. Controlling regimes ( )


Ea
k0 exp − RT
P(a+b)
p W
In order to gain further insights into the controlling regimes during Da =
membrane operation, we made use of two well known dimensionless FNH3 ,0
numbers in the field of membrane reactors: the Damkohler (Da) and the
FNH3 ,0
Peclet (Pe) numbers. The definitons are equivalent to the ones proposed Pe =
JH2 APp
by Gomez-garcia et al. [82] who modelled a similar shell-tube type
membrane reactor for ammonia decomposition. The choice of catalyst Pf
mass in the Da definition arrives from the definition of k0 unlike the Pr =
Pp
original work, where volume based definition were employed. An
additional dimensionless number was introduced for the regimes eval­ Where k0 is the pre-exponential factor (8.2 × 107 mol kgcat− 1 s− 1);
uation at higher pressures. This number was derived as the ratio be­ Ea the activation energy (116.11 kJ mol− 1); a the ammonia reaction
tween the feed and the permeate pressures. order (0.73); b the hydrogen reaction order (-1.62); Pf and Pp are feed

4
J.L. Cerrillo et al. Chemical Engineering Journal 431 (2022) 134310

Fig. 5. Effect of the retentate pressure in the ammonia conversion and hydrogen recovery (left), the molar composition and retentate flow (center) and molar
composition and permeate flow (right). Reaction conditions: 0.5%Ba-CoCe catalyst: 10 g (350–500 μm) diluted with 20 g of SiC ~ 680 μm. Ammonia flow rate = 200
N mL min− 1, WHSV = 1200 ml gcat-1h− 1, T = 485 ◦ C, Pretentate = 4–9 bar, Ppermeate = 1 bar.

Fig. 6. Effect of the permeate pressure using different retentate pressures on conversion and H2-recovery. Reaction conditions: 0.5%Ba-CoCe catalyst: 10 g (350–500
μm) diluted with 10 g of SiC ~ 680 μm. Ammonia flow rate 200 N mL min− 1, T = 485 ◦ C , Pretentate = 4–50 bar, Ppermeate = 1–15 bar.

and permeate pressures respectively; W the catalyst mass (g); FNH3,0 is transfer: convection through the catalytic bed against diffusion through
the initial ammonia flow rate (mol s− 1); JH2 the hydrogen permeance the membrane. High Da values (obtained for either a high catalyst
(1.17 × 10-7 mol m− 2 pa-1 s− 1) and A the membrane surface area loading or a low reactant feed flow) would indicate a fast reaction that is
(0.00716 m2). The values of reaction parameters are estimated with close to equilibrium and will be controlled by the convective mass
similar methodology followed by Sayas et al. [80] transfer from the gas to the catalyst. Low Pe values would indicate
Da compares the reaction rate against the convective mass transfer relatively faster diffusion through the membrane on account of high
rate, while the Pe number correlates the rates of two modes of mass permeability. As a consequence, a combination of high Da and low Pe

5
J.L. Cerrillo et al. Chemical Engineering Journal 431 (2022) 134310

Energy of the hydrogen recovered in the permeate − Net Energy requirements


η= (2)
Energy of the hydrogen recovered in the permeate

maximizes the potential of membrane reactors against their packed bed observed in Figure S15, where ammonia conversion and H2-recovery are
counterparts. On the contrary, low Da along with high Pe values are represented in terms of different pressure ratios. Once again, a signifi­
obtained for membrane reactors operating close to the performance of cant rise in Da values overcomes the slight increase in resistance to
packed bed reactors. permeation due to increased Pe numbers and allows for a higher re­
The Da and Pe numbers have been calculated using the equations covery of hydrogen on the permeate side of the reactor. Overall, this
described in the supplementary information and the values of reaction analysis shows that, under the studied reaction conditions, optimal
parameters have been estimated following a similar methodology as that usage of both catalyst and membrane is possible and highlights the
previously reported by Sayas et al. [80]. Fig. 7 collectively compares Da importance of a good matching between both reactor components.
and Pe values for various experiments performed in the packed bed
catalytic membrane reactor. In two instances where different catalyst 2.3.1. Efficiency of the process
loadings have been used (2 g represented by rhombuses and 10 g by Energy efficiency of the process is an issue of the utmost importance
stars). At high temperatures, Da values are sufficiently high to reach in the field energy storage. The overall efficiency in this case was
almost full ammonia conversion, as noted experimentally in Fig. 2. calculated using Equation (2).[19,83,84 85]
Simultaneously, the Pe values are constant and sufficiently low to pro­
mote hydrogen permeation, resulting in high H2-recoveries on the
permeate side. Even at higher Pe numbers, such as in the case of 10 g The energy requirements include: (i) ammonia evaporation (ΔHvap),
catalyst loading, the ammonia conversion and H2-recovery results are (ii) energy required to reach the reaction temperature (400–500 ◦ C,
still promising, due to the Da number values increasing roughly by an CpNH3 × ΔT), and (iii) the energy to convert ammonia into hydrogen in
order of magnitude when the catalyst amount is increased. the catalytic bed (ΔHr). The thermal energy losses in the system, or the
The effect of WHSV is shown in Fig. 7. At 485˚C and 2 g of catalyst thermal energy that could be recovered from the reaction products
loading, when WHSV varies from 1800 to 9000 ml gcat-1h− 1 the Da and during the cooling step from the reaction temperature (400–500 ◦ C) to
Pe numbers change (represented by squares). The evaluated operating room temperature are not considered here. [19]
conditions cover the low-to-moderate Pe range, and show conversions The heat required for the ammonia decomposition in order to
higher than 80% and high (>80%) recovery as Pe decreases; in good maintain hydrogen production can be supplied by gas turbine, high
agreement with the experimental data reported earlier in Fig. 2. Simi­ temperature fuel cell, or diesel power plants. However, for standalone
larly, at the same temperature, 10 g of catalyst loading and WHSV systems, a fraction of the hydrogen produced could be used to cover the
ranging from 360 to 1800 ml gcat-1h− 1 (represented by circles) the re­ energetic demands of the whole process, for instance using the hydrogen
sults cover the same Pe range but higher Da numbers; obtaining higher which is not recovered in the permeate. Thermodynamically, ammonia
conversions and higher recoveries, as expected (Fig. 4). In any case, both decomposition requires around 15–20 % of the hydrogen calorific value
experiments follow the same trend, where conversion and H2-recovery [84,86]. Therefore, with H2 recoveries around 80–85%, the energy
increase as the ratio between Da and Pe increases. Lastly, the effect of content of the retentate will be sufficient to satisfy the endothermic
pressure ratio studied in Fig. 6 is further explored in terms of dimen­ requirement for NH3 decomposition. In such a stand-alone system, the
sionless numbers, symbolized by triangles. Both, Da and Pe numbers are net energy requirements can be calculated according to Equation (3).
kept constant at different pressure ratios. This effect ban be better

( )
MWH2
Net energy requirements = ΔHvap + CpNH3 × ΔT + ΔHr × ηC − LHVH2 × × ηC × (1 − ηH2− rec ) × ηHeat− rec − LHVNH3 × (1 − ηC ) × ηHeat− rec (3)
MWNH3

Fig. 7. Comparison of (a) ammonia conversion


and (b) hydrogen recovery as a function of Da
and Pe for various experimental conditions uti­
lized in Figs. 2-4. (■) 2 g catalyst, 485 ◦ C, 1800
to 9000 ml gcat-1h− 1, Ppermeate = 1 bar and Pre­
tentate = 4 bar; (●) 10 g catalyst, 485 C, 360 to

1800 ml gcat-1h− 1, Ppermeate = 1 bar and Pretentate


= 4 bar; (◆) 2 g catalyst, 400 to 500 ◦ C, 3000 ml
gcat-1h− 1, Ppermeate = 1 bar and Pretentate = 4 bar;
(★) 10 g catalyst, 400 to 500 ◦ C, 1800 ml gcat-
1 − 1
h , Ppermeate = 1 bar and Pretentate = 4 bar; (▴)
10 g catalyst, 485 ◦ C, 1200 ml gcat-1h− 1, Ppermeate
= 1 to 45 bar and Pretentate = 1 to 15 bar.

6
J.L. Cerrillo et al. Chemical Engineering Journal 431 (2022) 134310

η in terms of the available options to cover the energetic demand for the
whole process. If no external energy inputs are available, a maximum
recovery of ~ 70–75% would be the optimum value for an energy
neutral process.
Finally, an extra consideration could be added to the overall effi­
ciency of the process by including the the energy requirements for the
compression of hydrogen produced in the permeate. Here we take 3
different examples: fuel cells (up to 10 bar), vehicles (up to 350 bar), and
hydrogen storage (up to 700 bar). Energy requirements for compression
of hydrogen for these three scenarios are depicted in Fig. 9 considering
hydrogen in the permeate is directly produced in a range between 1 and
20 bar. As it can be observed, there is a considerable difference in the
required energy as a function of the starting pressure of H2. This trans­
lates into important energy savings and highlights the importance of
operating membrane reactors at elevated pressures, not only from a
productivity point of view. For example, in the case of producing
hydrogen in the permeate at 5 bar, this will reduce the energetic re­
quirements to compress this hydrogen up to 350 bar, by circa a 30%
(compared to the case in which the compression was done from atmo­
spheric pressure). The savings could reach up to almost 60% if hydrogen
Fig. 8. Energetic analysis of the hydrogen production and purification process was produced at 20 bar prior to compression at 350 bar.
by ammonia decomposition conducted in a membrane reactor. Fig. 10 shows the efficiency of the process considering the energy
needed for compression. For example, if the hydrogen produced is
required to be compressed up to 350 bar, the efficiency would drop from
Where ηC is the ammonia cracker efficiency, (90–95%), ηH2-rec is the 77.7% to ~ 68–73%, when hydrogen in the permeate is produced at
hydrogen recovered (permeated hydrogen) from the cracked ammonia, 1–20 bar, respectively. Fig. 10 clearly reveals the benefits of producing
(85–95%) and ηheat-rec is the heat recovered from the retentate, the hydrogen permeate stream a certain value above the atmospheric
(70–90%). LHVi and MWi represent the lower heating value and the pressure, if the hydrogen produced requires further compression. The
molecular weight of the components, respectively. All the energetic drop in efficiency is smaller when increasing the pressure in the
terms in Equations 2–3 are expressed in kWh per kg of ammonia. The permeate (for all the scenarios considered here of compression up to 25,
energy of the hydrogen recovered in the permeate is given by Equation 350 or 700 bar). Only a 5% of efficiency loss is observed when the
(4) permeate is produced at 20 bar and requires further compression to 350
bar, compared to the 12% loss in the case of producing the permeate at 1
Energy of the hydrogen recovered in the permeate = LHV H2 × bar.
( ) Another benefit of producing hydrogen in the permeate at certain
MW NH3 × ηC × ηH2− rec (4) pressure above the atmospheric pressure is an important decrease in CO2
MW H2

emissions associated to the compression step (considering that these


Fig. 8 summarizes the parameters considered to determine the effi­
compressors are ran by non renewable energy sources). Fig. 11 shows
ciency of the process. According to equations 2–4, the overall efficiency
smaller CO2 emissions when hydrogen permeate is directly produced at
for the hydrogen production-purification system would be 77.8 %, when
increasing pressures in the permeate stream. A decrease in the CO2
considering the use of the H2 in the retentate (and eventually the non-
emissions in the range of 30–60% would be achieved when hydrogen is
converted NH3) to provide a large part of the energy needs; for an
compressed up to 350 bar from 5 to 20 bar, with respect to the case in
85% H2-recovery and 95% the efficiency of the cracker. It has to be
which the hydrogen was produced at atmospheric pressure.
stressed that this overall efficiency considers that the rest of the required
In the previous section we have demonstrated the possibility of
energy to drive the process is provided by waste streams, something that
producing the permeate at several pressures by simply increasing further
may not be realistic in case of stand-alone systems. Indeed, the optimum
the pressure in the retentate (Fig. 6). It is important to highlight that
H2-recovery for such stand-alone units would depend on their location
hardly any compression work is required to conduct the reaction at

Fig. 9. Energy requirements for hydrogen compression from 1 to 20 bar in the permeate up to 700 bar (left), and the energy savings resulting from the production of
a pressurized permeate streams in the range 5–20 bar (right).

7
J.L. Cerrillo et al. Chemical Engineering Journal 431 (2022) 134310

Fig. 10. Efficiency and efficiency loss when including the energetic compression requirements up to higher pressures in the range 25–700 bar, for the cases in which
the permeate is produced at 1–20 bar (80% H2-recovery was used for the calculation).

Fig. 11. CO2 emission associated to the compression step of the hydrogen produced in the permeate.

elevated pressure, because pressurizing liquid ammonia is simple and power savings in the membrane-based process are found due to the
cheap. absence of the PSA and the higher pressure of the final H2 stream after
the reactor (Supplementary Table S4). As a result, a 31% utility cost
2.3.2. Hydrogen purification technologies: Membrane reactor versus savings could be achieved. Moreover, we also need to point out that, due
conventional pressure swing Adsorption. to the difficulty of achieving 100% conversion in standard reactors and
The feasibility of the ammonia decomposition process to produce despite the NH3 recovery unit that works at − 100 ◦ C, close to 500 ppm of
high purity hydrogen has been additionally investigated by comparing NH3 slip is observed in the PSA scenario (stream S9, in supplementary
the membrane reactor to a conventional pressure swing adsorption Scheme S2). This will most likely imply the need for an additional NH3
(PSA) technology using Aspen plus simulations. Here we compare the slip reaction unit (ASC) [89] as, apart from being one order of magni­
energy consumption for the two hydrogen purification technologies tude higher than the Permissible Exposure Limit (50 ppm), small traces
(Supplementary Table S4). of NH3 can compromise the performance of the PSA and compression
The conventional process uses an equilibrium reactor (GIBBS), fol­ units [90]. Furthermore, Fuel Cells require NH3 concentrations below
lowed by an NH3 recovery unit, a PSA purification unit and a H2 0.1 ppm[91]. As a result, extra energy requirements will be needed in
compression unit (Supplementary Scheme S2). The GIBBS reactor the conventional scenario for this ASC unit. On the other hand, in our
operates at 20 bar and 500 ◦ C in the reactor. The PSA adsorption is membrane system, the H2 stream directly comes out virtually free of
simplified as a theoretical separator with 90% H2 recovery and 0.1% N2 NH3 (see Fig. 3), revealing another important advantage of our system.
slip [87]. The energy consumption of the vacuum pump (0.5 bar) and Moreover, the retentate stream still rich in H2, and could be used in a
the further recompression are considered the main energy inputs for this cogeneration cycle to generate 3 MW of the electric power needed to run
technology [88]. On the other side, the membrane reactor is modeled as the plant, reducing further the energy consumption and utility costs
a combination of a GIBBs equilibrium reactor and a theoretical separator (Supplementary Scheme S4).
with 95% recovery (Supplementary Scheme S3). No PSA unit is needed Following, we evaluated the accuracy of the above estimations of
in this case. The reaction conditions for the membrane reactor are energy efficiency for an energy stand-alone membrane system by per­
selected from experimental data, in particular 20 bar of feed, 3 bar of H2 forming additional simulations with 85% and 75% H2 recovery in the
permeate, and 95% H2 recovery with 99.97% purity (see Fig. 6). membrane. While reducing the recovery will lessen the hydrogen pro­
The hydrogen purification technology coupled to the NH3 decom­ duction rate, it will result in higher H2 content of the retentate stream for
position was successfully modeled with Aspen plus for the membrane the cogeneration cycle and, at the same time, higher pressures in the
reactor and the conventional PSA, delivering the target 99.9% H2 purity. permeate stream (see Fig. 6), both aspects being beneficial for the stand-
The initial case study has been considered at 20 bar and 500 ◦ C in the alone energetic operation of the plant. The results can be observed in the
reactor, with H2 delivery also at 20 bar. Interestingly, 51% electric Supporting Information at Table S4 and, indeed, lowering the H2

8
J.L. Cerrillo et al. Chemical Engineering Journal 431 (2022) 134310

Fig. 12. Composition and flow at


retentate (a), ammonia conversion and
H2 recovery (b) and composition and
flow at permeate (c) at long-term sta­
bility test with Pd-Au membrane module
Reaction conditions: 0.5%Ba-CoCe cata­
lyst: 10 g (350–500 μm) diluted with 20
g of SiC ~ 680 μm; Ammonia flow rate
= 170 N ml min− 1; T = 485 ◦ C, Pretentate
= 4 bar; Ppermeate = 1 bar. Ammonia
conversion, H2 recovery and H2 purity of
the permeate (d) using the same condi­
tions in cycles of heating/cooling.

recovery in the membrane results in higher energy produced in the several cooling /heating cycles were conducted. The membrane module
coreneration cycle with lower electricity needed for the final compres­ was purged with inert gas (Ar, 100 N mL min− 1 for 1 h), and kept under
sion stem. While in neither of the cases we generate enough electric Ar to protect the H2-selective membrane and to prevent the reoxidation
power for a stand-alone operation, we need to stress that the power cycle of the Co-based catalyst. Then, cooling and heating ramps were set at
of Scheme S4 (Supplementary Information) is solely designed for elec­ 1 ◦ C min− 1 to prevent the stress between membrane and ceramic sup­
tric power generation and, in our particular case, we could omit the port. The cooling /heating cycles were followed by the ammonia reac­
second turbine with the water-steam cycle and used instead the heat of tion at the same operating conditions as in the long-term stability test,
the low pressure hot gas (Supplementary Scheme S4, 18.25 MW in for 24 h each cycle. The catalytic performance for six cycles (Fig. 12 and
HEATX for a 75% H2 recovery) to provide the required energy for the Supplementary Figure S17,) clearly showed the stability of the mem­
reactor (19.1 MW). In this case, the first turbine will provide more than brane, with no significant differences in conversion and recovery.
enough power (8.5 MW) for the rest of the utilities, effectively reaching Overall, the total working time of the system under several reaction
the desired energy and stand-alone operation. conditions reached ~ 1000 h (when including the long term and addi­
tional experiments performed with the same membrane module and
2.3.3. Stability of the PBCMR catalyst).
Last but not least, the stability of the system was further analyzed,
since it could become one of the limiting steps to the development of this 3. Conclusions
technology [83,92]. Fig. 12 shows the catalytic performance over the
reaction time for the long-term stability test conducted. No detrimental The production of high purity H2 from NH3 using packed bed
effects were observed on ammonia conversion, hydrogen recovery, membrane reactors holds great promise. The robustness and durability
composition or flow rates of the permeate and retentate. From the cat­ of such system has been demonstrated for >1000 h on stream, yielding a
alytic point of view, no ammonia was detected in the chromatographic very pure hydrogen stream (>99.97 % H2) and H2 recovery (>90 %). In
analysis of the retentate, neither in the permeate stream. To uncover any situ separation of hydrogen results in faster cracking kinetics and helps
unreacted ammonia under the detection limit of the chromatographs overcome equilibrium limitations. However, a good match between the
(detection limit 0.01–0.02 %vol.), a sulfuric acid trap (0.5 M) was properties of catalyst and membrane needs to be achieved: Pd based
installed to capture ammonia at the ppm level (in the permeate and membranes are very effective in a relatively narrow operation window
retentate sides). Aliquots of the sulfuric acid trap were daily analyzed (namely 400 to 550 ◦ C), implying that the use of catalysts active and
following the indophenol blue protocol [93]. We can unequivocally selective in this temperature range is mandatory and that productivity of
assure that no ammonia was detected in any of the exhausts of the membrane an catalysts need to be well matched. This is the case for the
system, even at the ppm level. combination studied in this work.
In order to validate the robustness of the PBCMR, the system was set When considering NH3 as potential H2 carrier, efficiency of the
at the maximum operating conditions specified by the supplier, 500 ◦ C cracking has to account for the final application of the produced H2,
and 16 bar for 24 h. After that, the conditions were set back to the including compression to the desired pressure and important environ­
previous long-term stability test for 24 extra hours. The results (Sup­ mental aspects such as potential NH3 slip. Our work demonstrates that
plementary Figure S16) were almost identical to the previous ones, nearly thermodynamic efficiencies can be achieved when the appro­
corroborating the stability and the robustness of the PBCMR. Finally, priate membrane reactor is employed. Indeed, using this technology, we

9
J.L. Cerrillo et al. Chemical Engineering Journal 431 (2022) 134310

have demonstrated that stand alone systems with a H2 efficiency of circa Additional experiments were performed using 10 g of BaCoCe and SiC in
75% can be achieved in the production of H2 at 350 bar. We expect our a 1:2 ratio. The effect of the operational conditions (P, T and WHSV) on
results to pave the way towards the the development of this tehcnology ammonia conversion, hydrogen recovery and productivity were evalu­
in a scalable manner through parallelization, enabling a pathway for ated using the PBCMR. Long-term stability tests were carried out for
distributed use of hydrogen from ammonia for a wide range of >1000 h, including heating/cooling cycles and diverse operational
applications. conditions to determine the robustness of the membrane and the sta­
bility of the catalyst. Preliminary experiments to evaluate the hydrogen
4. Experimental details permeability and capability of the membrane module (Supplementary
Figures S1-S4).
The barium promoted cobalt-cerium catalyst (BaCoCe, 80/20 M ratio
Co/Ce) was synthesized by coprecipitation followed by the addition of Declaration of Competing Interest
barium as promoter using incipient wetness impregnation (0.5 wt%. Ba).
Experimental details of the synthesis procedure and characterization of The authors declare that they have no known competing financial
the catalyst can be found in our previous work [73]. Prior to the ex­ interests or personal relationships that could have appeared to influence
periments, the catalyst was reduced under pure hydrogen (at atmo­ the work reported in this paper.
spheric pressure) at 485 ◦ C for 8 h.
The membrane reactor (PBCMR) was supplied by GaoQ Functional
Acknowledgment
Materials Co. (Nanjing, China). The reaction system consists of a Pd-Au
alloy membrane (i.d., 8 mm; o.d, 12 mm; effective membrane area, 70
The authors gratefully acknowledge the financial support provided
cm2; thickness, 8 µm; Au content, 15 wt%) in a stainless-steel module.
by Saudi Aramco, and the resources and facilities provided by the King
The thin Pd-Au alloy layer is supported on the external surface of a
Abdullah University of Science and Technology.
porous ceramic tube. The shell and tube sides are the retentate and
permeate chambers, respectively. The catalyst (2–10 g) is incorporated
in the annular space (retentate chamber) together with SiC in a catalyst: Author Contributions
SiC weight ratio of 1:2.
The ammonia decomposition tests were performed using a PID N.M. and J.G. conceived this work. J.L.C and N.M. designed, char­
Microactivity Reference system. The experimental setup is shown in acterized and conducted synthesis and catalytic experiments. N.R. and S.
Scheme S1 in the Supplementary Information. The original furnace has R.K. performed control regime studies. A.R. performed the technology
been replaced with a longer one to ensure isothermal operation in the simulation comparison and provided experimental support. J.G., P.C.
membrane reactor. The experiments were carried out at different tem­ and S.M.S. provided revision suggestions. J.L.C and N.M. wrote the
peratures (300 to 600 ℃), pressures (1 to 25 bar) and WHSV (300–1800 original draft of the manuscript. All authors have given the approval to
gcat mL-1h− 1) using different amounts of catalyst (2 to 10 g) diluted with the final version of the manuscript.
SiC in diverse ratios (catalyst: SiC = 1:10 and 1:2). Each set of experi­
mental conditions was kept for two hours. The flow rates of NH3, H2, N2 Appendix A. Supplementary data
and Ar were carefully fixed by mass flow controller in order to explore
those different reaction conditions. Gases from the two chambers were Supplementary data to this article can be found online at https://doi.
analyzed online by two Micro-GC in an online gas chromatograph org/10.1016/j.cej.2021.134310.
(3000A Micro-GC gas analyzer, Agilent) equipped with thermal con­
ductivity detectors and using He as internal standard (2 ml min− 1). One References
channel equipped with a PLOT-U precolumn, followed by a Molsieve
column using Ar as the carrier gas was used for detection of N2 and H2; a [1] C.H. Christensen, T. Johannessen, R.Z. Sørensen, J.K. Nørskov, Towards an
ammonia-mediated hydrogen economy? Catalysis Today 111 (1-2) (2006)
second channel with a PLOT-U column and He as the carrier gas, was 140–144.
used for NH3 detection. [2] M. Yadav, Q. Xu, Liquid-phase chemical hydrogen storage materials, Energy &
The experimental setup is designed to feed liquified ammonia with Environmental Science 5 (12) (2012) 9698–9725.
[3] G.W. Crabtree, M.S. Dresselhaus, M.V. Buchanan, The hydrogen economy, Physics
the possibility of performing reactions at high pressure. This feature of
today 57 (12) (2004) 39–44.
the system enables high-pressure experiments and could potentially be [4] J.O. Abe, A. Popoola, E. Ajenifuja, O. Popoola, Hydrogen energy, economy and
used to generate pressurized pure hydrogen on the permeate side. storage: review and recommendation, International Journal of Hydrogen Energy
Experimental verification of the concept was done at feed pressure up to 44 (29) (2019) 15072–15086.
[5] G.E. Ioannatos, X.E. Verykios, H2 storage on single-and multi-walled carbon
50 bar and permeate pressure up to 20 bar. nanotubes, International Journal of Hydrogen Energy 35 (2) (2010) 622–628.
During heating, both the tube/retentate side and the shell/permeate [6] T. Bell, L. Torrente-Murciano, H 2 production via ammonia decomposition using
side of the module were fed with Ar to avoid hydrogen embrittlement of non-noble metal catalysts: A review, Topics in Catalysis 59 (15) (2016) 1438–1457.
[7] C.D. Demirhan, W.W. Tso, J.B. Powell, E.N. Pistikopoulos, Sustainable ammonia
Pd membrane at low temperatures. The temperature was carefully production through process synthesis and global optimization, AIChE Journal 65
raised with 1–2 ◦ C/min. (7) (2019) e16498, https://doi.org/10.1002/aic.16498.
The maximum flow rate was limited by the system’ s ability to [8] A. Valera-Medina, F. Amer-Hatem, A.K. Azad, I.C. Dedoussi, M. de Joannon, R.
X. Fernandes, P. Glarborg, H. Hashemi, X. He, S. Mashruk, J. McGowan,
maintain a constant catalyst bed temperature, as the high flow rate and C. Mounaim-Rouselle, A. Ortiz-Prado, A. Ortiz-Valera, I. Rossetti, B. Shu, M. Yehia,
endothermic reaction may result in a significant temperature gradient H. Xiao, M. Costa, Review on ammonia as a potential fuel: from synthesis to
across the catalyst bed. economics, Energy & Fuels 35 (9) (2021) 6964–7029.
[9] M. Benés, G. Pozo, M. Abián, Á. Millera, R. Bilbao, M.U. Alzueta, Experimental
The use of a sweep gas could decrease the partial pressure of Study of the Pyrolysis of NH3 under Flow Reactor Conditions, Energy & Fuels 35
hydrogen in the permeate side, thereby, increasing the driving force for (9) (2021) 7193–7200.
permeation, but for hydrogen-production applications, dilution by a [10] T. Sinigaglia, F. Lewiski, M.E. Santos Martins, J.C. Mairesse Siluk, Production,
storage, fuel stations of hydrogen and its utilization in automotive applications-a
sweep gas is, in principle, not desired. Because of this, all experiments
review, International journal of hydrogen energy 42 (39) (2017) 24597–24611.
were conducted without sweep gas to obtain pure hydrogen in the [11] A. Klerke, C.H. Christensen, J.K. Nørskov, T. Vegge, Ammonia for hydrogen
permeate. storage: challenges and opportunities, Journal of Materials Chemistry 18 (20)
A set of experiments was carried out using 200 mg of pelletized (2008) 2304–2310.
[12] N. Morlanés, S.P. Katikaneni, S.N. Paglieri, A. Harale, B. Solami, S.M. Sarathy,
BaCoCe between 300 μm and 500 μm and diluted with 2 g of SiC, J. Gascon, A technological roadmap to the ammonia energy economy: Current state
mimicking the experimental conditions reported in our previous work. and missing technologies, Chemical Engineering Journal 408 (2021), 127310.

10
J.L. Cerrillo et al. Chemical Engineering Journal 431 (2022) 134310

[13] A. Valera-Medina, H. Xiao, M. Owen-Jones, W.I.F. David, P.J. Bowen, Ammonia for producing COx-free hydrogen from ammonia decomposition, Applied Catalysis B:
power, Progress in Energy and combustion science 69 (2018) 63–102. Environmental 211 (2017) 167–175.
[14] M.A. Sutton, J.W. Erisman, F. Dentener, D. Möller, Ammonia in the environment: [45] X. Li, W. Ji, J. Zhao, S. Wang, C. Au, Ammonia decomposition over Ru and Ni
From ancient times to the present, Environmental Pollution 156 (3) (2008) catalysts supported on fumed SiO2, MCM-41, and SBA-15, Journal of Catalysis 236
583–604. (2) (2005) 181–189.
[15] R.E. Schofield, The enlightenment of Joseph Priestley: a study of his life and work [46] C. Huang, Y. Yu, J. Yang, Y. Yan, D. Wang, F. Hu, X. Wang, R. Zhang, G. Feng, Ru/
from 1733 to 1773, Penn State Press, 1997. La2O3 catalyst for ammonia decomposition to hydrogen, Applied Surface Science
[16] V. Smil, Detonator of the population explosion, Nature 400 (6743) (1999) 415. 476 (2019) 928–936.
[17] J.W. Erisman, M.A. Sutton, J. Galloway, Z. Klimont, W. Winiwarter, How a century [47] I. Lucentini, A. Casanovas, J. Llorca, Catalytic ammonia decomposition for
of ammonia synthesis changed the world, Nature Geoscience 1 (10) (2008) hydrogen production on Ni, Ru and NiRu supported on CeO2, International Journal
636–639. of Hydrogen Energy 44 (25) (2019) 12693–12707.
[18] I. Matito-Martos, J. García-Reyes, A. Martin-Calvo, D. Dubbeldam, S. Calero, [48] X. Ju, L. Liu, X. Zhang, J.i. Feng, T. He, P. Chen, Highly Efficient Ru/MgO Catalyst
Improving ammonia production using zeolites, The Journal of Physical Chemistry with Surface-Enriched Basic Sites for Production of Hydrogen from Ammonia
C 123 (30) (2019) 18475–18481. Decomposition, ChemCatChem 11 (16) (2019) 4161–4170.
[19] S. Giddey, S.P.S. Badwal, C. Munnings, M. Dolan, Ammonia as a renewable energy [49] W. Zheng, J. Zhang, B. Zhu, R. Blume, Y. Zhang, K. Schlichte, R. Schlögl, F. Schüth,
transportation media, ACS Sustainable Chemistry & Engineering 5 (11) (2017) D.S. Su, Structure–function correlations for Ru/CNT in the catalytic decomposition
10231–10239. of ammonia, ChemSusChem: Chemistry & Sustainability Energy & Materials 3 (2)
[20] X. Wen, J. Guan, Recent advancement in the electrocatalytic synthesis of ammonia, (2010) 226–230.
Nanoscale 12 (15) (2020) 8065–8094. [50] K. Nagaoka, T. Eboshi, N. Abe, S.-I. Miyahara, K. Honda, K. Sato, Influence of basic
[21] M. Goff, Distribution of amonia as ana energy carrier, Ammonia Energy Coference dopants on the activity of Ru/Pr6O11 for hydrogen production by ammonia
2020, Ammonia energy association (2020). decomposition, International journal of hydrogen energy 39 (35) (2014)
[22] K.E. Lamb, M.D. Dolan, D.F. Kennedy, Ammonia for hydrogen storage, A review of 20731–20735.
catalytic ammonia decomposition and hydrogen separation and purification, [51] X. Duan, J. Ji, G. Qian, C. Fan, Y. Zhu, X. Zhou, D.e. Chen, W. Yuan, Ammonia
International Journal of Hydrogen Energy 44 (7) (2019) 3580–3593. decomposition on Fe (1 1 0), Co (1 1 1) and Ni (1 1 1) surfaces: A density functional
[23] N.M. Adli, H. Zhang, S. Mukherjee, G. Wu, ammonia oxidation electrocatalysis for theory study, Journal of Molecular Catalysis A: Chemical 357 (2012) 81–86.
hydrogen generation and fuel cells, Journal of the Electrochemical Society 165 [52] J.P. Collins, J.D. Way, Catalytic decomposition of ammonia in a membrane reactor,
(15) (2018) J3130. Journal of membrane science 96 (3) (1994) 259–274.
[24] C. Zamfirescu, I. Dincer, Using ammonia as a sustainable fuel, Journal of Power [53] F.R. García-García, Y.H. Ma, I. Rodríguez-Ramos, A. Guerrero-Ruiz, High purity
Sources 185 (1) (2008) 459–465. hydrogen production by low temperature catalytic ammonia decomposition in a
[25] N.J.M. Duijm, F.; Paulsen, J. L.;, Safety assessment of ammonia as a transport fuel - multifunctional membrane reactor, Catalysis Communications 9 (3) (2008)
Risoe-R No. 1504, Risø National Laboratory, Denmark, 2005. 482–486.
[26] Z. Cesaro, M. Ives, R. Nayak-Luke, M. Mason, R. Bañares-Alcántara, Ammonia to [54] G. Lanzani, K. Laasonen, NH3 adsorption and dissociation on a nanosized iron
power: Forecasting the levelized cost of electricity from green ammonia in large- cluster, International journal of hydrogen energy 35 (13) (2010) 6571–6577.
scale power plants, Applied Energy 282 (2021), 116009. [55] H. Muroyama, C. Saburi, T. Matsui, K. Eguchi, Ammonia decomposition over Ni/
[27] A. Afif, N. Radenahmad, Q. Cheok, S. Shams, J.H. Kim, A.K. Azad, Ammonia-fed La2O3 catalyst for on-site generation of hydrogen, Applied Catalysis A: General
fuel cells: a comprehensive review, Renewable and Sustainable Energy Reviews 60 443-444 (2012) 119–124.
(2016) 822–835. [56] C. Plana, S. Armenise, A. Monzón, E. García-Bordejé, Ni on alumina-coated
[28] J. Li, H. Huang, N. Kobayashi, Z. He, Y. Osaka, T. Zeng, Numerical study on effect cordierite monoliths for in situ generation of CO-free H2 from ammonia, Journal of
of oxygen content in combustion air on ammonia combustion, Energy 93 (2015) Catalysis 275 (2) (2010) 228–235.
2053–2068. [57] F. García-García, A. Guerrero-Ruiz, I. Rodríguez-Ramos, A. Goguet, S. Shekhtman,
[29] I.P. Jain, Hydrogen the fuel for 21st century, International journal of hydrogen C. Hardacre, TAP studies of ammonia decomposition over Ru and Ir catalysts,
energy 34 (17) (2009) 7368–7378. Physical Chemistry Chemical Physics 13 (28) (2011) 12892–12899.
[30] I. Staffell, D. Scamman, A. Velazquez Abad, P. Balcombe, P.E. Dodds, P. Ekins, [58] J. Polanski, P. Bartczak, W. Ambrozkiewicz, R. Sitko, T. Siudyga, A. Mianowski,
N. Shah, K.R. Ward, The role of hydrogen and fuel cells in the global energy J. Szade, K. Balin, J. Lelątko, S. Bhargava, Ni-supported Pd nanoparticles with Ca
system, Energy & Environmental Science 12 (2) (2019) 463–491. promoter: a new catalyst for low-temperature ammonia cracking, PloS one 10 (8)
[31] L. Wang, Y. Zhao, C. Liu, W. Gong, H. Guo, Plasma driven ammonia decomposition (2015) e0136805.
on a Fe-catalyst: eliminating surface nitrogen poisoning, Chemical communications [59] D.A. Hansgen, D.G. Vlachos, J.G. Chen, Using first principles to predict bimetallic
49 (36) (2013) 3787–3789. catalysts for the ammonia decomposition reaction, Nature chemistry 2 (6) (2010)
[32] M. Akiyama, K. Aihara, T. Sawaguchi, M. Matsukata, M. Iwamoto, Ammonia 484–489.
decomposition to clean hydrogen using non-thermal atmospheric-pressure plasma, [60] Z.-W. Wu, X. Li, Y.-H. Qin, L. Deng, C.-W. Wang, X. Jiang, Ammonia decomposition
International Journal of Hydrogen Energy 43 (31) (2018) 14493–14497. over SiO2-supported Ni–Co bimetallic catalyst for COx-free hydrogen generation,
[33] H. Yuzawa, T. Mori, H. Itoh, H. Yoshida, Reaction mechanism of ammonia International Journal of Hydrogen Energy 45 (30) (2020) 15263–15269.
decomposition to nitrogen and hydrogen over metal loaded titanium oxide [61] A. Srifa, K. Okura, T. Okanishi, H. Muroyama, T. Matsui, K. Eguchi, Hydrogen
photocatalyst, The Journal of Physical Chemistry C 116 (6) (2012) 4126–4136. production by ammonia decomposition over Cs-modified Co3Mo3N catalysts,
[34] A. Utsunomiya, A. Okemoto, Y. Nishino, K. Kitagawa, H. Kobayashi, K. Taniya, Applied Catalysis B: Environmental 218 (2017) 1–8.
Y. Ichihashi, S. Nishiyama, Mechanistic study of reaction mechanism on ammonia [62] Ö. Akarçay, S.F. Kurtoğlu, A. Uzun, Ammonia decomposition on a highly-dispersed
photodecomposition over Ni/TiO2 photocatalysts, Applied Catalysis B: carbon-embedded iron catalyst derived from Fe-BTC: Stable and high performance
Environmental 206 (2017) 378–383. at relatively low temperatures, International Journal of Hydrogen Energy 45 (53)
[35] F. Chang, W. Gao, J. Guo, P. Chen, Emerging Materials and Methods toward (2020) 28664–28681.
Ammonia-Based Energy Storage and Conversion, Adv. Mater. 33 (50) (2021) [63] X. Duan, G. Qian, X. Zhou, Z. Sui, D. Chen, W. Yuan, Tuning the size and shape of
2005721. Fe nanoparticles on carbon nanofibers for catalytic ammonia decomposition,
[36] C. Zhong, W.B. Hu, Y.F. Cheng, Recent advances in electrocatalysts for electro- Applied Catalysis B: Environmental 101 (3–4) (2011) 189–196.
oxidation of ammonia, Journal of Materials Chemistry A 1 (10) (2013) 3216–3238. [64] V. Cechetto, L. Di Felice, J.A. Medrano, C. Makhloufi, J. Zuniga, F. Gallucci, H2
[37] M. Feyen, C. Weidenthaler, R. Güttel, K. Schlichte, U. Holle, A.-H. Lu, F. Schüth, production via ammonia decomposition in a catalytic membrane reactor, Fuel
High-Temperature Stable, Iron-Based Core-Shell Catalysts for Ammonia Processing Technology 216 (2021), 106772.
Decomposition, Chemistry–A, European Journal 17 (2) (2011) 598–605. [65] N. Sazali, W.N.W. Salleh, A.F. Ismail, Carbon tubular membranes from
[38] F. Schüth, R. Palkovits, R. Schlögl, D.S. Su, Ammonia as a possible element in an nanocrystalline cellulose blended with P84 co-polyimide for H2 and He separation,
energy infrastructure: catalysts for ammonia decomposition, Energy & International Journal of Hydrogen Energy 42 (15) (2017) 9952–9957.
Environmental Science 5 (4) (2012) 6278–6289. [66] C. Zhao, B. Sun, J. Jiang, W. Xu, H2 purification process with double layer bcc-
[39] A.K. Hill, L. Torrente-Murciano, Low temperature H2 production from ammonia PdCu alloy membrane at ambient temperature, International Journal of Hydrogen
using ruthenium-based catalysts: Synergetic effect of promoter and support, Energy 45 (35) (2020) 17540–17547.
Applied Catalysis B: Environmental 172-173 (2015) 129–135. [67] Z. Zhang, S. Liguori, T.F. Fuerst, J.D. Way, C.A. Wolden, Efficient ammonia
[40] A. Di Carlo, L. Vecchione, Z. Del Prete, Ammonia decomposition over commercial decomposition in a catalytic membrane reactor to enable hydrogen storage and
Ru/Al2O3 catalyst: An experimental evaluation at different operative pressures utilization, ACS Sustainable Chemistry & Engineering 7 (6) (2019) 5975–5985.
and temperatures, international journal of hydrogen energy 39(2) (2014) 808-814. [68] N.W. Ockwig, T.M. Nenoff, Membranes for hydrogen separation, Chemical reviews
[41] L. Li, Z. Zhu, G. Lu, Z. Yan, S. Qiao, Catalytic ammonia decomposition over CMK-3 107 (10) (2007) 4078–4110.
supported Ru catalysts: Effects of surface treatments of supports, Carbon 45 (1) [69] A. Di Carlo, A. Dell’Era, Z. Del Prete, 3D simulation of hydrogen production by
(2007) 11–20. ammonia decomposition in a catalytic membrane reactor, International Journal of
[42] L. Li, Z. Zhu, Z. Yan, G. Lu, L. Rintoul, Catalytic ammonia decomposition over Ru/ Hydrogen Energy 36 (18) (2011) 11815–11824.
carbon catalysts: The importance of the structure of carbon support, Applied [70] S.H. Israni, B.K.R. Nair, M.P. Harold, Hydrogen generation and purification in a
Catalysis A: General 320 (2007) 166–172. composite Pd hollow fiber membrane reactor: Experiments and modeling, Catalysis
[43] S.-F. Yin, B.-Q. Xu, C.-F. Ng, C.-T. Au, Nano Ru/CNTs: a highly active and stable Today 139 (4) (2009) 299–311.
catalyst for the generation of COx-free hydrogen in ammonia decomposition, [71] N.S. Patki, S.-T. Lundin, J.D. Way, Apparent activation energy for hydrogen
Applied Catalysis B: Environmental 48 (4) (2004) 237–241. permeation and its relation to the composition of homogeneous PdAu alloy thin-
[44] X. Ju, L. Liu, P. Yu, J. Guo, X. Zhang, T. He, G. Wu, P. Chen, Mesoporous Ru/MgO film membranes, Sep. Purif. Technol. 191 (2018) 370–374.
prepared by a deposition-precipitation method as highly active catalyst for

11
J.L. Cerrillo et al. Chemical Engineering Journal 431 (2022) 134310

[72] L. Shi, A. Goldbach, G. Zeng, H. Xu, Preparation and performance of thin-layered [82] M.Á. Gómez-García, I. Dobrosz-Gómez, J. Fontalvo, J.M. Rynkowski, Membrane
PdAu/ceramic composite membranes, Int. J. Hydrogen Energy 35 (9) (2010) reactor design guidelines for ammonia decomposition, Catalysis Today 191 (1)
4201–4208. (2012) 165–168.
[73] N. Morlanés, S. Sayas, G. Shterk, S.P. Katikaneni, A. Harale, B. Solami, J. Gascon, [83] L. Lin, Y. Tian, W. Su, Y. Luo, C. Chen, L. Jiang, Techno-economic analysis and
Development of a Ba–CoCe catalyst for the efficient and stable decomposition of comprehensive optimization of an on-site hydrogen refuelling station system using
ammonia, Catal. Sci. Technol. 11 (9) (2021) 3014–3024. ammonia: hybrid hydrogen purification with both high H 2 purity and high
[74] J.B. Von Colbe, J.-R. Ares, J. Barale, M. Baricco, C. Buckley, G. Capurso, N. recovery, Sustainable Energy & Fuels 4 (6) (2020) 3006–3017.
Gallandat, D.M. Grant, M.N. Guzik, I. Jacob, Application of hydrides in hydrogen [84] J.L. Nieminen, Experimental investigation of hydrogen production from ammonia
storage and compression: Achievements, outlook and perspectives, international decomposition, 2011.
journal of hydrogen energy 44(15) (2019) 7780-7808. [85] J.H. Kim, O.C. Kwon, A micro reforming system integrated with a heat-
[75] C. Deng, M. Zhu, Y. Zhou, X. Feng, Novel conceptual methodology for hydrogen recirculating micro-combustor to produce hydrogen from ammonia, International
network design with minimum compression work, Energy 159 (2018) 203–215. journal of hydrogen energy 36 (3) (2011) 1974–1983.
[76] G. Sdanghi, G. Maranzana, A. Celzard, V. Fierro, Review of the current [86] K.E. Lamb, D.M. Viano, M.J. Langley, S.S. Hla, M.D. Dolan, High-purity H2
technologies and performances of hydrogen compression for stationary and produced from NH3 via a ruthenium-based decomposition catalyst and vanadium-
automotive applications, Renewable and Sustainable Energy Reviews 102 (2019) based membrane, Industrial & Engineering Chemistry Research 57 (23) (2018)
150–170. 7811–7816.
[77] J.B. Von Colbe, J.-R. Ares, J. Barale, M. Baricco, C. Buckley, G. Capurso, N. [87] Air-Liquid, Pressure Swing Adsorption (PSA) - Hydrogen Purification, 2021. htt
Gallandat, D.M. Grant, M.N. Guzik, I. Jacob, Application of hydrides in hydrogen ps://www.engineering-airliquide.com/pressure-swing-adsorption-psa-hydr
storage and compression: Achievements, outlook and perspectives, international ogen-purification. (Accessed 08/2021).
journal of hydrogen energy 44(15) (2019) 7780-7808. [88] M.T. Ho, G.W. Allinson, D.E. Wiley, Reducing the cost of CO2 capture from flue
[78] S. Armenise, E. García-Bordejé, J. Valverde, E. Romeo, A. Monzón, A Langmuir- gases using pressure swing adsorption, Industrial & Engineering Chemistry
Hinshelwood approach to the kinetic modelling of catalytic ammonia Research 47 (14) (2008) 4883–4890.
decomposition in an integral reactor, Physical Chemistry Chemical Physics 15 (29) [89] Umicore, Ammonia Slip Catalyst (ASC). https://ac.umicore.com/en/technolo
(2013) 12104–12117. gies/ammonia-slip-catalyst/. (Accessed 08-2021 2021).
[79] G. Djéga-Mariadassou, C.-H. Shin, G. Bugli, Tamaru’s model for ammonia [90] G.F.B. Brendel, J E; Rathbone, R F; Frey, Jr, R N, INVESTIGATION OF AMMONIA
decomposition over titanium oxynitride, Journal of Molecular Catalysis A: ADSORPTION ON FLY ASH DUE TO INSTALLATION OF SELECTIVE CATALYTIC
Chemical 141 (1–3) (1999) 263–267. REDUCTION SYSTEMS, 2000.
[80] S. Sayas, N. Morlanés, S.P. Katikaneni, A. Harale, B. Solami, J. Gascon, High [91] Z. Du, C. Liu, J. Zhai, X. Guo, Y. Xiong, W. Su, G. He, A Review of Hydrogen
pressure ammonia decomposition on Ru–K/CaO catalysts, Catalysis Science & Purification Technologies for Fuel Cell Vehicles, Catalysts 11 (3) (2021) 393.
Technology 10 (15) (2020) 5027–5035. [92] S. Adhikari, S. Fernando, Hydrogen membrane separation techniques, Industrial &
[81] G. Li, M. Kanezashi, H.R. Lee, M. Maeda, T. Yoshioka, T. Tsuru, Preparation of a Engineering Chemistry Research 45 (3) (2006) 875–881.
novel bimodal catalytic membrane reactor and its application to ammonia [93] M. Wang, S. Liu, T. Qian, J. Liu, J. Zhou, H. Ji, J. Xiong, J. Zhong, C. Yan, Over
decomposition for COx-free hydrogen production, International journal of 56.55% Faradaic efficiency of ambient ammonia synthesis enabled by positively
hydrogen energy 37 (17) (2012) 12105–12113. shifting the reaction potential, Nature, Communications 10 (1) (2019) 341.

12

You might also like