You are on page 1of 13

Journal of Membrane Science 614 (2020) 118483

Contents lists available at ScienceDirect

Journal of Membrane Science


journal homepage: http://www.elsevier.com/locate/memsci

A catalytic composite membrane reactor system for hydrogen production


from ammonia using steam as a sweep gas
Yongha Park a, Junyoung Cha a, Hyun-Taek Oh b, Taeho Lee a, Sung Hun Lee c, Myung Gon Park c,
Hyangsoo Jeong a, Yongmin Kim a, Hyuntae Sohn a, Suk Woo Nam a, d, Jonghee Han a, d,
Chang Won Yoon a, e, **, Young Suk Jo a, *
a
Center for Hydrogen and Fuel Cell Research, Korea Institute of Science and Technology (KIST), 5 Hwarang-ro 14-gil, Seongbuk-gu, Seoul, 02792, Republic of Korea
b
Department of Chemical and Biomolecular Engineering, Yonsei University, 50 Yonsei-ro, Seodaemun-gu, Seoul, 03722, Republic of Korea
c
Wonik Materials Co., 30 Yangcheong 3-gil, Ochang-eup. Chengwon-gu, Cheongju-si, Chungcheongbuk-do, 28125, Republic of Korea
d
Green School, Korea University, Anam-ro 145, Seongbuk-gu, Seoul, 02841, Republic of Korea
e
KHU-KIST Department of Converging Science and Technology, Kyung Hee University, 26 Kyungheedae-ro, Dongdaemun-gu, Seoul, 02447, Republic of Korea

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

Keywords: For catalytic reactions involving H2 extraction, the membrane reactor is an attractive option for enhancing the
Catalytic membrane reactor equilibrium and kinetics while eliminating excessive purification steps. In this study, a steam carrier adopted
Pure hydrogen production composite membrane reactor system is developed to produce pure H2 (>99.99%) from ammonia with high H2
Ammonia decomposition
productivity (>0.35 mol-H2 g−cat1 h− 1) and ammonia conversion (>99%) at a significantly reduced operating
Steam sweep gas
Membrane reactor process demonstration
temperature (<723 K). Coupling of a custom developed palladium/tantalum composite metallic membrane and
ruthenium on lanthanum-doped alumina catalysts allowed stable operation of the membrane system with sig­
nificant mass transfer enhancement. Various reactor assemblies involving as-fabricated membranes and catalysts
are experimentally compared to suggest the optimal configuration and operating conditions for future applica­
tions. Steam is adopted as a sweep gas, presenting efficient H2 recovery (>91%) while replacing conventionally
utilized noble carrier gases that require additional gas separation processes. The steam carrier presents similar
membrane reactor performance to that of noble gases, and the water reservoir used for steam generation acts as
an ammonia buffer via scrubbing effects. Finally, electricity generation is demonstrated using a commercial fuel
cell along with process simulation, substantiating potential of the proposed membrane system in practical ap­
plications for H2 production from ammonia and on-site power generation.

1. Introduction being actively investigated.


Recently, ammonia has been identified as one of the energy carriers
The renewable energy share is growing as a mean to mitigate the that allow efficient storage of renewable energy in a liquid form at a mild
global climate crisis and maintain sustainability. While renewable en­ pressure under atmospheric conditions (0.8 MPa, 293 K [1]). When
ergy has great potential, its unpredictable nature presents challenges for liquefied, it has a high energy density of 22.5 MJ kg− 1 (higher heating
direct integration into the grid. Additionally, either 1) it has a small value), with high gravimetric and volumetric H2 contents of 17.6 wt%
energy production scale or 2) large production sites are too distant from and 108 g L− 1 at 293 K [2], respectively. Additionally, well-established
energy utilization sites. Therefore, a large amount of renewable energy storage and distribution infrastructure makes ammonia an economically
remains untapped, with limited logistics to store and transport the en­ feasible mass-scale energy carrier. While ammonia synthesis is
ergy efficiently and economically. To resolve these temporal and spatial energy-intensive, efforts have been directed toward making the process
discrepancies between the supply and demand of renewable energy, “greener” by adopting water electrolysis in conjunction with a conven­
energy carriers with high energy densities and economic feasibilities are tional Haber Bosch process [3] or by electrochemical synthesis [4]. This

* Corresponding author.
** Corresponding author. Center for Hydrogen and Fuel Cell Research, Korea Institute of Science and Technology (KIST), 5 Hwarang-ro 14-gil, Seongbuk-gu, Seoul,
02792, Republic of Korea.
E-mail addresses: cwyoon@kist.re.kr (C.W. Yoon), yjo@kist.re.kr (Y.S. Jo).

https://doi.org/10.1016/j.memsci.2020.118483
Received 24 April 2020; Received in revised form 2 July 2020; Accepted 4 July 2020
Available online 25 July 2020
0376-7388/© 2020 Elsevier B.V. All rights reserved.
Y. Park et al. Journal of Membrane Science 614 (2020) 118483

green ammonia can be shipped from the renewable energy producer to is constantly being improved, reducing the reaction temperature,
the renewable energy consumer, whereby it can be cracked into H2 and potentially to the point where the equilibrium NH3 decomposition
N2. H2 production from ammonia, shown in Fig. 1, is a well-known conversion starts to drop significantly (below 98% at 623 K [12]). While
catalytic process involving catalytic chemical reaction and H2 purifica­ theoretical benefits of CMR are widely accepted, it is important to note
tion steps. that the increase in conversion and catalytic activity might be marginal
For the ammonia decomposition reaction, Ru is known to be the most in some cases. Detailed analyses are thus required to truly benefit from
active metal, and various support materials, such as alumina [5,6], the adoption of CMR. In the following, the proposed membrane reactor
mesoporous silica [7,8], activated carbon [9], metal oxides [10], and concepts for NH3 decomposition are reviewed, with references to studies
carbon nanotubes [11] are being explored as effective support materials performed from 1994 to the present.
[12,13]. Additionally, there have been significant efforts to replace Ru Early studies on NH3 decomposition membrane reactors focused on
with non-noble metals as active sites owing to its cost and scarcity [5,14, application for the reduction of NOX emissions. Collins et al. designed a
15]. Even after decades of research, a large amount of work is being membrane reactor using a Pd-ceramic membrane with Ni/Al2O3 cata­
performed to reduce the reaction temperature and the metal loading of lysts. The NH3 conversion increased from 53% (packed bed reactor,
expensive Ru. After the catalytic NH3 decomposition reaction, there are denoted as PBR) to 94% (catalytic membrane reactor, denoted as CMR)
two purification steps required to obtain H2 with a high purity: removal at 873 K [23]. Gobina et al. modeled an NH3 membrane reactor using
of unreacted NH3 and separation of N2. Even at a high temperature and experimental H2 permeation data of Pd–Ag membrane and an NH3
extensive catalyst loading, there remains a small amount of unreacted decomposition kinetic model for Ni/Al2O3 catalysts [24], and Abashar
NH3 in the product stream [16], which can be detrimental to a et al. performed simulations to determine the equilibrium shift via in situ
proton-exchange membrane fuel cell (PEMFC) having a trace level (~10 H2 extraction [25]. Later studies focused more on the use of NH3 as a H2
ppm) of NH3 tolerance [17,18]. While there are numerous reports on the carrier, to produce pure H2 as an energy carrier. Zhang et al. demon­
development of catalysts for ammonia decomposition, few studies have strated the production of 99.96% pure H2 with a recovery yield of 77%
investigated the whole process, including H2 purification. Kojima et al. using an electroless plated Pd membrane with a Ni/La–Al2O3 catalyst
studied the effects of Li-exchange X-type zeolite (Li-X) as an adsorbent [26]. García-García et al. fabricated a CMR utilizing a Pd membrane and
material for unreacted ammonia [19], and Dolan et al. purified the Ru/C catalysts for producing high-purity H2, in an attempt to exceed the
product stream from ammonia decomposition using a Pd-coated tubular equilibrium conversion at a temperature lower than 653 K [27]. Israni
V membrane to get >99.99% pure H2 [20]. Our group recently et al. adopted a hollow fiber membrane reactor with Pd nanopore
demonstrated operation of a 1-kW-class fuel cell using H2 produced by membranes and Ni catalysts with 10% conversion increase with CMR
an ammonia decomposition reformer, along with the removal of NH3 [28].
using 13X zeolite, resulting in <1 ppm of NH3 in the product stream While most studies utilized Pd membranes, Li et al. fabricated a
[16]. bimodal CMR by doping Ru on a silica membrane support, which served
Finally, after the removal of NH3, high-purity H2 can be produced by as a catalytic support layer. The ammonia conversion was significantly
employing pressure swing adsorption (PSA) or membrane processes to higher for the bimodal configuration (95% at 723 K) than for the
separate N2, which accounts for 25 vol.% of the produced gas. This monomodal configuration (45% at 723 K) [29]. In a follow-up study, Li
adsorption-based H2 purification steps can comprise a significant et al. examined the effects of the γ-Al2O3 impregnation amount, pore
portion of the fuel reforming process, leading to high H2 costs [21]. structure, and catalytic activity of Ru on the performance of a bimodal
While it is possible to operate a PEMFC without separating N2 (as it is membrane reactor. The purity of the permeate H2 stream was 84%, with
harmless to the fuel-cell stack), this reduces the H2 partial pressure and 4.5% NH3 [30]. While the silica membrane is one of viable membrane
prohibits H2 recirculation, adversely affecting the stack performance. options for NH3 CMRs, its low H2 selectivity necessitates an additional
Also, the N2 content in the product stream will reduce the energy density adsorptive device to remove NH3 and N2. More recent works again
of the compressed H2 produced from ammonia to be used in the H2 adopted advanced Pd membrane fabrication and catalyst synthesis
refueling station. techniques. Itoh et al. fabricated an NH3 CMR using a 0.2-mm-thick
The catalytic reaction and H2 purification can be conducted sepa­ tubular Pd membrane with a Ru/SiO2 catalyst and achieved a 15%
rately, but this makes the NH3 decomposition process complex and conversion increase compared to a PBR at 723 K [31]. Rizzuto et al.
energy-intensive. To address this issue, membrane reactor concepts have utilized a commercial CMR with a Pd membrane and Ru catalyst in an
been suggested. Membrane reactors are particularly beneficial for the attempt to identify the operation condition where equilibrium conver­
NH3 decomposition reaction, for several reasons: (i) high-purity H2 can sion could be achieved with no use of sweep gas [32]. Zhang et al.
be produced with no extra extensive NH3 purification or N2 separation produced >99.7% pure H2 using a CMR with a 6-μm-thick Pd film
steps, allowing power generation using fuel cells with a low NH3 toler­ deposited on a Ru-impregnated yttria-stabilized zirconia tube, allowing
ance or use as a pure H2 source for H2 charging stations; (ii) enhanced almost complete NH3 conversion at temperatures as low as 673 K [33].
NH3 decomposition reaction kinetics can be achieved with in situ In addition to experimental studies, a few simulation studies have been
extraction of H2, as described by the Temkin–Pyzhev equation [22], performed, with developed models coupling the NH3 reaction kinetics,
reducing the reaction temperature; (iii) the thermodynamic equilibrium gas diffusion in the catalyst bed, and H2 permeation through the mem­
can be shifted, which is becoming important as the catalyst performance brane via the solution-diffusion mechanism [34–37].

Fig. 1. Ammonia as a renewable energy carrier: H2 production process from ammonia.

2
Y. Park et al. Journal of Membrane Science 614 (2020) 118483

Previous studies clearly indicate the effectiveness of coupling H2 in an RuCl3∙xH2O solution, followed by drying in a vacuum oven. A
separation using a membrane with NH3 decomposition to increase NH3 quantitative analysis of the metal contents in the Ru/La–Al2O3 catalyst
conversion. However, for the practical application of the NH3 CMR was performed using inductively coupled plasma optical emission
technology, further investigation of various aspects is required. First, the spectroscopy (720 ICP-OES, Agilent Technologies, Inc.). The results
optimal membrane types for reducing the usage of Pd and increasing the indicated a composition of approximately 0.65 wt% Ru and 10 mol% La
purity of the produced H2 must be identified. Then, the productivity or ions.
throughput of the CMR must be improved to increase the H2 production
per Ru metal and membrane area. Finally, the use of a vacuum pump or 2.2. Design and operation of CMR
inert gas carrier in the permeate side must be avoided, or there must be a
novel way to increase the H2 partial pressure difference between the A schematic of the system used in both the PBR for the catalytic-
feed and permeate sides. Our group recently investigated the viability of activity test and the CMR is shown in Fig. 2. The detailed characteris­
a Pd coated body-centered cubic (V, Nb, Ta) composite dense metallic tics of the apparatus and operating conditions are presented in Table 1.
membranes for high-purity H2 production from mixed gas separation The system was automatically operated and recorded by an AIO/DIO
[38,39], and applied it to an NH3 decomposition, benefiting from the PLC interface and a control program (LabVIEW, National Instruments
matched operation region between H2 permeation and the NH3 Corp., USA).
decomposition reaction [40]. However, to further substantiate the The ammonia decomposition reaction was conducted with a
application of the composite membrane reactor concept, enhanced mass concentric double-pipe reactor configuration: the outside shell was a 3/
transfer for higher H2 productivity, performance analysis over wide 4′′ stainless-steel tube, and the inside was a 1/4′′ dummy Ta membrane
range of operation conditions, and removal of commonly used inert tube for the PBR experiments and a 1/4′′ Pd/Ta/Pd composite mem­
sweep gases (e.g. N2, Ar, etc), were mandatory. brane for the CMR experiments. It was confirmed that H2 permeation
In this study, new CMR configurations utilizing a custom fabricated through Ta tube was not detected without Pd coatings in a temperature
Pd/Ta/Pd composite membrane and optimized Ru/La–Al2O3 pellet range tested. For a fair comparison between the PBR and CMR opera­
catalysts are developed to substantially increase the H2 productivity and tions, the reactor geometries were kept the same (Fig. S1 shows the
recovery yield while reducing the Ru catalyst loading for better utili­ membrane assembly). Ru(0.65 wt%)/La(10 mol%)-Al2O3 catalysts were
zation of the limited membrane surface area and Ru catalysts. Two CMR packed in a dummy Ta tube or a Pd/Ta/Pd composite membrane, and
configurations and one PBR configuration are compared at different the pressure drops between the inlet and outlet streams under various
temperatures, pressures, NH3 feed rates, and sweep gas flowrates to reaction conditions were confirmed to be insignificant. The feed-side
determine the optimal operation conditions for an NH3 CMR adopting flowrates were controlled by electrical mass flow controllers (MFCs,
dense metallic composite membranes. Then, steam is adopted as a sweep F–201C, Bronkhorst, High-Tech), and an electrical backpressure
gas to reduce the permeate-side H2 pressure while acting as an NH3 trap controller (EPC, P–702C, Bronkhorst, High-Tech) was used to maintain a
in the case of membrane failure or leakage. Finally, as-produced H2 from constant feed pressure under varying reactant flowrates. To supply the
the CMR was directly fed to a small-scale PEMFC to demonstrate onsite sweep gas to the permeate side, an MFC for the N2 carrier and a high-
H2 and electricity generation applications, and a membrane process pressure liquid chromatography pump (Series II Legacy HPLC pump,
simulation was further conducted to balance the heat requirement of the TELEDYNI SSI) for the steam carrier were utilized. To evaporate
steam generation and endothermic NH3 decomposition. deionized (DI) water and preheat the steam supplied to the permeate
side, a 1-m-long 1/16′′ stainless-steel tube was installed in the electric
2. Experimental furnace. The flowrate of the separated H2 through the Pd/Ta/Pd com­
posite membrane (permeated side) was measured by a mass flow meter
2.1. Fabrication of membranes and catalysts (MFM, F-201 CL, Bronkhorst, High-Tech, Netherlands) after the flow
passed the gas–liquid separator and silica adsorbents (Silica gel blue,
For Pd/Ta/Pd composite membrane fabrication, dense Ta tubes medium 5–10 mesh, Samchun Chemical Co.). The H2 permeation rate
(>99.95%) with an outer diameter of 6.35 mm and a wall thickness of was determined according to the total flowrate of the permeate side
0.25 mm were purchased from Koralco Co., (Republic of Korea). The measured via MFMs, excluding the N2 or steam sweep gas flowrate. All
lengths of the Ta tubes tested were 100 and 50 mm for higher NH3 feed the MFCs and MFMs used in this study were calibrated using film flow
and lower NH3 feed, respectively. To deposit the thin Pd layer on the meters (SF–1U/2U, Horiba, Japan).
outside and inside surfaces of the Ta tubes, electroless plating (ELP) was In the experimental system, the temperature was controlled by an
employed. The detailed procedure of the ELP method adopted in this electrical tube furnace, and three thermocouples were installed at the
study was presented in previous reports [38,39,41]. The thickness of the feed inlet, the permeate, and the outer wall of a double pipe reactor to
Pd layer on the Ta tube was controlled by changing the plating time, and ensure that the experimental temperature was in the desired range. The
the thickness of the coating layer was determined via scanning electron pressures of both the feed (catalytic reaction bed) and permeate sides
microscopy (SEM, Inspect F, FEI). After the Pd plating on the Ta tube, were measured using a pressure transducer (PSAN, Autonics; range:
heat treatment was performed under a N2 atmosphere at 773 K for 2 h to 0–10 bar; accuracy within 0.1% of full scale). Pneumatic valves were
form a dense Pd layer on the Ta surface and to reduce the voids between installed and automatically activated according to the operating con­
Pd and Ta layers for better adhesion. The crystallinity analysis was ditions and designated flow directions. Check valves and ball valves
conducted using X-ray diffraction spectroscopy. were installed at the proper positions to prevent reverse flow and
For the catalyst preparation, ruthenium chloride hydrate physical impacts on the measuring instruments. To analyze the con­
(RuCl3∙xH2O, 40.2 wt% Ru, Sigma–Aldrich), lanthanum nitrate hexa­ centration of the unreacted NH3 from the reactor, retentate gases were
hydrate (La(NO3)3⋅6H2O, Crown Guaranteed Reagents), and γ-Al2O3 1/ introduced to the NH3 analyzer (IR type (0–6000 ppm), Airwell +7,
8′′ pellet (Alfa Aesar) were purchased and used as received, unless stated KINSCO technology, Republic of Korea) with 7 L min− 1 of dry air as a
otherwise. The γ-Al2O3 1/8′′ pellet was pulverized and sieved to diluent.
appropriate pellet sizes (1–2 mm) considering the inner diameter of the
Pd/Ta/Pd tubular membrane. La (target concentration of 10 mol%) was 2.3. Performance evaluation
doped onto the γ-Al2O3 as a promoter of the catalyst via immersion in a
lanthanum nitrate solution followed by a synthesis condition introduced The performance of dense metallic composite or alloy membrane
in a previous study [16]. Subsequently, Ru was deposited on La-γ-Al2O3 separation processes is generally evaluated with regard to the ability to
supports via the wet impregnation method. The supports were immersed transport H2 through the membrane, e.g., the flux and permeability, and

3
Y. Park et al. Journal of Membrane Science 614 (2020) 118483

Fig. 2. Process flow diagram (PFD) of the PBR and CMR.

reactor is assembled, a minor leak may occur, depending on the fittings


Table 1
or gaskets used to seal the membrane feed chamber from the permeate
Characteristics of the CMR and operating conditions.
chamber or the outside of the reactor. In our study, Swagelok ferrules
Configuration 1 (for Configuration 2 (for were adopted (with the optimized fitting torque between the membrane
higher NH3 feed) lower NH3 feed)
and the ferrules) to avoid failure due to H2 embrittlement while ensuring
Membrane Materials leakage-free operation of the reactor system. Swagelok fittings were
Membrane type Pd/Ta/Pd composite membrane (tubular)
tightened using a torque wrench under constant torque and then tested
Ta tube outer diameter [mm] 6.35
Ta tube thickness [μm] 250
for leakage, to find the minimum torque assuring leakage-free operation.
Ta tube Length/Effective 100/76.0 50/26.0 The results of impurity analyses of the as-produced H2 are presented in
permeation length [mm] Section 3.4. The NH3 conversion, H2 recovery yield, and H2 productivity
Effective permeation area 1396.0 477.6 were calculated via the following equations.
[mm2]
Pd deposition time [min] 40 FRetentate, NH3 FFeed, NH3 (1 − x)
Pd layer thickness [μm] Approximately 1–2 CNH3 = = (2)
FRetentate FFeed, NH3 (1 + x) + FAir − FPermeate, H2
Catalytic Materials ( )
Catalyst Ru(0.65 wt%)/La(10 mol%)-Al2O3 FFeed, NH3 − CNH3 FFeed, NH3 +FAir − FPermeate, H2
NH3 conversion, 100x[%]= ( ) ×100
Particle pellet size [mm] 1–2 FFeed, NH3 1+CNH3
Packed catalyst amount [g] 1.0 0.38
(3)
Operating conditions
FPermeate, H2
Pressure [barabsolute] 1–5 H2 recovery yield[%] = × 100 (4)
Temperature [K] 673–773 1.5FFeed, NH3 x
Feed flowrate [cm3 min− 1] 20–100
Sweep gas flowrate [cm3 0–100 [ ] 1.5FFeed, NH3 x
min− 1]
Productivity mol hr− 1 g− 1 = (5)
Catalyst (or metal) amount

where, CNH3 , FRetentate,NH3 , FRetentate , FFeed,NH3 , FAir , FPermeate,H2 , and x


by the quality (e.g., purity) of the produced H2. The H2 permeation rate,
represent the concentration of ammonia in the retentate gas measured
permeability, and H2 purity, along with a flow diagram including the
by the NH3 analyzer, the unreacted NH3 flowrate in the retentate, the
mass balance of the CMR used in this study, are presented in Fig. S1. The
total flowrate of the retentate gas, the NH3 feed flowrate, the diluent
permeability, which is defined as a combination of the diffusivity (D)
(air) flowrate, the H2 permeation rate, and the ammonia conversion
and solubility (S), was calculated, assuming the conventional solution-
rate, respectively. NH3 conversion was calculated based on the
diffusion mechanism in bulk Ta. The H2 flux, which is an important
measured NH3 concentration in the retentate stream (detailed deriva­
performance metric for the metallic membrane, was determined by the
tion shown in Appendix 1 in Supplementary Information).
H2 permeation rate divided by the effective permeation area shown
Permeated H2 and sweep gas were directly introduced to the com­
below
mercial PEMFC (H-12, Horizon Fuel Cell Technologies, Japan). The
Fpermeate, H2 concentration of NH3 on the permeate side was determined using an
H2 flux = (1)
Effective permeation area ultraviolet–visible (UV–vis) spectrophotometer (Cary 100, Agilent Tech,
USA) after the permeated gas was dissolved for 2 h in 10 mL of a 10 mM
The H2 produced from dense metallic membranes with a thickness
sulfuric acid solution with Nessler’s reagent.
greater than a certain level is usually ultrapure. However, when the

4
Y. Park et al. Journal of Membrane Science 614 (2020) 118483

The CMR was operated at a pressure up to 5 bar and in the tem­ 1.0 for mass transfer control. As the bulk diffusion dominant hydrogen
perature range of 673–773 K, with a gas hourly space velocity (GHSV) transport is widely accepted for Pd coated body-centered cubic (V, Nb,
ranging from 1200 to 6000 cm3 g−cat1 h− 1. Prior to each operation, the Ru/ Ta) composite membranes [43] and experimentally verified in previous
La–Al2O3 catalyst was reduced at 773 K under H2:N2 (50:50 cm3 min− 1) studies by our group [38,39], 0.5 for exponential factor was adopted in
conditions in reactors for >2 h to ensure removal of Cl ions which might this study as well.
affect the catalytic activity. The experimental reproducibility of the The flux of the membrane and recovery yield reached 10.34 cm3
CMR operation was confirmed to be within 5% by performing experi­ min− 1 cm− 2 and 96.3% at 5 bar of feed pressure, respectively, and both
ments in triplicate under the same conditions. Finally, to evaluate the of them significantly increased at a higher operating pressure (Fig. 3b).
theoretical energy and mass balance of the developed CMR process, They tended to converge above a feed pressure of 4 bar, as the feed
Aspen Plus (Aspen Technology Inc., USA) was used as a process simu­ flowrate was fixed at 200 cm3 min− 1, but more H2 can permeate with
lation tool. For calculations of the thermodynamic properties, the Pen­ higher feed rates. H2 recovery yield did not increase further, owing to
g–Robinson equation of state was applied [42]. mass transfer limit incurred by increasing partial pressure of N2 with
increasing H2 recovery yield above 95%. With pure H2, the calculated
3. Results and discussion permeability and permeance values were 6.03 × 10− 8 mol m− 1 s− 1
Pa− 0.5 and 3.05 × 10− 7 mol m− 2 s− 1 Pa− 1 at 773 K. The permeability was
3.1. Characterization and performance analyses of composite membrane much higher than that of the Pd–Ag membrane (1.6–2.4 × 10− 8 mol
and catalyst m− 1 s− 1 Pa− 0.5) and comparable to those of other Pd/Ta/Pd membranes
reported in previous studies (6–20 × 10− 8 mol m− 1 s− 1 Pa− 0.5) at similar
The Pd/Ta/Pd composite membrane used in the membrane reactor temperature range [40,43]. Additionally, no embrittlement issues were
was fabricated via ELP, and the amount of Pd deposited on the Ta was observed in the operating range tested, owing to the optimized torque
estimated according to the weight change after the ELP. After ELP on fitting around the membrane tube as described in the experimental
100-mm-long tubular Ta, approximately 0.38 g of total weight increase section and previous method by our group [39,40].
was observed, including various chemical compounds from the ELP
2NH3 → N2 + 3H2 (6)
solutions. The actual thickness of the Pd layer and the surface mor­
phologies were determined via SEM, as shown in Fig. 3a. Approximately Ammonia can decompose to hydrogen (Eq. (6)), and equilibrium
1.3–2.3 μm of Pd was uniformly deposited on the both feed and conversion of ammonia to hydrogen can be achieved to >98% at 623 K
permeate surface of the Ta tube, with good adhesion after the heat [12]. However, a relevant catalyst is necessary to lower the activation
treatment. After repeated ELP, a repeatable thickness and weight gain energy, promoting the dehydrogenation reaction from ammonia. While
from the Pd was confirmed under the given plating conditions. Ru is known to be the most active metal for NH3 decomposition, catalyst
Accordingly, it was considered that a Pd/Ta/Pd composite membrane supports were also known to play an important role in promoting cat­
having an average Pd thickness of approximately 1.5 μm was obtained. alytic activity. Previous studies indicated that La-doped Al2O3 catalysts
XRD peaks of the bare Ta and as-fabricated Pd/Ta/Pd membrane increase the electron density of Ru via strong metal to support interac­
confirmed deposition of metallic Pd on Ta (Fig. S2). The differences in tion and form Lewis acid sites on the catalyst surface [16,40]. Promoting
the thicknesses of the plated Pd between the inside and outside of the Ta effects of alkali metals (e.g. La, Mg, K) on catalysts performance
tube were possibly due to concentration gradients formed in the plating improvement, in terms of activity and durability, have been previously
bath, requiring further optimization of the plating methods. reported [44,45]. In this study, a Ru(0.65 wt%)/La(10 mol%)-Al2O3
In this study, the H2 permeation of the fabricated Pd/Ta/Pd com­ pellet was synthesized using the wet impregnation method. Detailed
posite membrane (Config. 1, Table 1) was tested using a H2(75%)/ characterizations of the synthesized catalyst were reported in the liter­
N2(25%) binary gas mixture, simulating the gas compositions when ature [16]. Fig. 4 shows the catalytic activity of the synthesized pellet
ammonia feed is fully converted into H2 and N2. The H2 permeation rate catalyst for ammonia decomposition under different GHSVs (NH3
through the membrane significantly depends on the difference in H2 flowrate divided by the packed catalyst amount) and temperatures. The
partial pressures on the feed and permeate sides, as well as permeability activity of the catalyst increased with the temperature and decreased
determining bulk diffusion of H atoms across the bulk Ta layer by with the increasing GHSV. The ammonia conversion was approximately
solution-diffusion mechanism [43] (detailed mechanism described in 99% at >798 K and 1 bar under all the GHSV conditions. Based on the
Fig. S3). According to Sieverts’ law, the H2 permeation rate increases results obtained, the kinetically limited region of the synthesized cata­
linearly with a constant exponential factor of the pressure difference. lysts was <773 K, where the membrane reactor can be beneficial for
The numerical value of the exponential factor in Sieverts’ law depends increasing the NH3 conversion.
on the transport mechanism in the membrane; 0.5 for bulk diffusion and

Fig. 3. (a) SEM image of cross section and surface of the composite membrane with digital picture of the membrane assembly and (b) H2 permeation performance of
the fabricated Pd/Ta/Pd composite membrane under 200 cm3 min− 1 flow of a H2(75%)/N2(25%) binary gas mixture with a feed pressure of up to 5 bar at 773 K with
a sweep N2 flowrate of 100 cm3 min− 1.

5
Y. Park et al. Journal of Membrane Science 614 (2020) 118483

Fig. 4. Catalytic activity of the ammonia decomposition reaction on Ru(0.65


wt%)/La(10 mol%)-Al2O3. The reaction was conducted in the temperature
range of 623–823 K (increments of 25 K) with a GHSV of 1200–3000 (in­
crements of 600 cm3 g−cat1 h− 1).

3.2. Performance of composite membrane reactor: optimal feed flowrate


and pressure

The performance of the CMR for ammonia decomposition was


evaluated with respect to the inlet ammonia flowrate (or GHSV), tem­
perature, and NH3 feed pressure. To increase the pressure difference
between the retentate and permeate sides, a 100 cm3 min− 1 N2 carrier
flow was introduced to permeate side in some cases. Critical perfor­
mance metrics, i.e., the H2 permeation rate and NH3 conversion of each
CMR operation (Config. 1, Table 1) at a pressure of 5 bar, are presented
in Fig. 5a and b. As shown in Fig. 5a, the H2 permeation rate increased
rapidly with the increasing temperature and GHSV, and the H2 perme­
ation rate gradually converged above a certain GHSV, except at 773 K.
The performance of the H2 permeation at 773 K had the potential to
increase with the GHSV. In binary H2/N2 separation, as shown in Fig. 3,
the composite membrane produced >144 cm3/min of H2 at a H2 partial
pressure of 3.75 bar on the retentate side. Therefore, when the GHSV
and partial pressure of H2 on the retentate side are increased at 773 K,
the H2 permeation rate of the CMR can reach a value similar to that for Fig. 5. (a) H2 permeation rate, (b) NH3 conversion, and (c) H2 recovery yield of
H2/N2 separation. the composite membrane reactor (Config. 1, Table 1) at temperatures ranging
As shown in Fig. 5b, the NH3 conversion gradually decreased with from 673 to 773 K (increments of 25 K) under a feed pressure of 5 bar with
GHSVs ranging from 1200 to 6000 (increments of 600 cm3 g−cat1 h− 1), corre­
the increasing GHSV, and the conversion was >99% even at a GHSV of
sponding to an ammonia flowrate of 20–100 cm3 min− 1, and with a sweep N2
4800 cm3 g−cat1 h− 1 at 773 K. Additionally, up to a GHSV of 3000 cm3 g−cat1
flowrate of 100 cm3 min− 1.
h− 1, the NH3 conversion was >99% at both 748 and 773 K. Furthermore,
>99% conversion was achieved at a GHSV of 1200 cm3 g−cat1 h− 1 at all the
of H2 was recovered at 773 K. However, the recovery yields rapidly
temperatures except for 673 K. Compared with previous results,
decreased with the increasing GHSV, despite increasing the H2 pro­
considering the even lower Ru content on the catalyst and the same
duction amount, as shown in Fig. 5a. Fig. S4b presents a comparison of
composite membranes, the CMR geometry with a shorter distance from
the H2 flowrates on the permeate and retentate sides according to the H2
the catalysts to the surface of the membrane used in the present study
production amount at 723 K. As the NH3 feed flowrate increased, both
enhanced the mass transfer of H2 to the membrane side and reduced the
the amount of H2 produced and the amount of H2 permeated increased.
concentration gradient on the catalytic side (shown in Fig. S1). CMR
However, the residence time in the catalytic bed decreased owing to the
operations of 673 K were not possible at a GHSV of >3600 cm3 g−cat1 h− 1,
high GHSV, causing a majority of the produced H2 to be discharged
owing to the measurement limits of the ammonia analyzer employed.
through the retentate stream rather than being separated. According to
These results suggest higher NH3 conversion at lower operating tem­
the foregoing results, the GHSV range of 2400–3000 cm3 g−cat1 h− 1 was the
peratures and higher GHSVs, indicating significant improvements
optimal condition in the developed CMR configuration.
compared to the previous study of our group [40] and other previous
Furthermore, compared to PBR results in Fig. 4, the H2 productivity
literatures. This was further confirmed by evaluating the partial pres­
was significantly improved with CMR at the identical temperature and
sures in the feed and permeate sides (Fig. S4a).
GHSV despite at higher pressures (Fig. 5a). H2 productivities of PBR and
Fig. 5c shows the H2 recovery yields of the CMR operations. For all
CMR are 40 and 70 mmol-H2 g−cat1 h− 1, respectively, at 663 K and at GHSV
the CMR operations, the H2 recovery yield reached its maximum value
of 3000 cm3 g−cat1 h− 1, presenting more than 63% improvements
and then decreased with the increasing GHSV. The GHSV range of
benefitting from the in situ H2 extraction. However, the improvement of
2400–3000 cm3 g−cat1 h− 1 exhibited the highest recovery values in all the
H2 productivity of CMR decreased to 7% at 773 K, implying that the
temperature ranges tested. All the recovery yields were >85%, and 91%

6
Y. Park et al. Journal of Membrane Science 614 (2020) 118483

effectiveness of CMR is more prominent at lower temperature. Then, H2 function of the pressure in the temperature range of 673–773 K and at
productivity was compared to results of previous studies including our pressures up to 5 bar. The conversion decreased significantly with an
group’s. H2 productivity was measured to be 350 mmol-H2 g−cat1 h− 1 at increase in the pressure. The conversion at 773 K and 5 bar was similar
773 K under 5 bar. H2 productivity was under 200 mmol-H2 g−cat1 h− 1 for to that at 748 K and 1 bar, and the other results also exhibited a decrease
literatures using CMR with Ru or Ni catalysts at optimum operating in the conversion from 6.7% to 17.7% with the increasing pressure.
conditions [27,28,46]. Considering specific Ru or Ni metal loading, H2 However, the reaction kinetics can be enhanced with in situ removal of
1
productivity values are 10,700 and 54,800 mmol-H2 g−catmetal h− 1 at 673 the H2 by reducing the average H2 concentration in the catalyst bed. The
K and 773 K under 5 bar, respectively, about 30 times higher than that of performance of the CMR was compared with that of the PBR under the
CMR using Ni catalyst and more than 2 times higher than that of CMR identical conditions (Fig. 6b), and a conversion improvement was
using Ru catalyst. apparent even under a comparatively high GHSV. It has to be noted that
Fig. 6a and b show the effect of the operating pressure on the per­ PBR eventually needs to be operated at pressurized conditions with PSA.
formance of CMR with respect to that of PBR. Because the GHSV of 3000 With in situ H2 extraction using the membrane reactor, the conversion
cm3 g−cat1 h− 1 was the optimal condition for CMR operation, the perfor­ increased with the pressure under the identical conditions used for the
mance was evaluated with respect to the operating pressure of the CMR PBR. Increasing the partial pressure of NH3 through in situ H2 extraction
under this condition. As mentioned previously, a pressurized feed is enhanced the NH3 adsorption on the catalyst, improving the reaction
preferred for both H2 extraction and recovery yield improvement owing kinetics. This implies that an increase in the H2 permeation/extraction
to the increased driving force for H2 permeation. H2 recovery yield from the CMR not only increased the H2 recovery but also improved the
increased with the increasing operating pressure, as shown in Fig. 6a, NH3 conversion. At a reaction temperature of 773 K, the conversion was
and they tended to increase marginally above a feed pressure of 3 bar, as >99% regardless of the pressure. The NH3 conversion was improved by
explained by H2 flux trends shown in Fig. S5a. Furthermore, considering >36.7% compared with that of the PBR at temperature of 723 K and
vapor pressure of ammonia (6.2 bar at 283 K) charged for trans­ pressure of 5 bar.
portation, the operation pressure of 5 bar for the CMR is deemed In addition to the operating GHSV, pressure, and temperature of the
appropriate. In the CMR operation at 723 K, the H2 recovery yield was CMR, the performance of the membrane reactor may vary according to
improved from 77.5% to 88.7% when the pressure increased from 1 to 5 the effective area of the composite membrane and the amount of cata­
bar, and under other temperature conditions, the recovery yields lysts packed in the membrane reactor. In Fig. 7a and b, the H2 flux and
increased by 8.1%–12.9% with the increasing pressure. NH3 conversion for two different CMR configurations (Config. 1 and
While a pressurized feed is beneficial for H2 production in a mem­ Config. 2, Table 1) are compared over different GHSVs. The two con­
brane reactor, it negatively impacts the ammonia decomposition in the figurations exhibited similar trends of the flux and conversion with the
catalytic reaction. Fig. 6b shows the NH3 conversion in the PBR as a increasing GHSV, but Config. 2 exhibited a significantly higher H2 flux
than Config. 1. Considering theoretical permeability of Ta membranes

Fig. 6. (a) H2 recovery yield of CMR at temperatures ranging from 673 to 773 K
(increments of 25 K) and (b) NH3 conversion of the PBR and CMR (Config. 1, Fig. 7. (a) H2 flux and (b) NH3 conversion for two different CMR configurations
Table 1) at temperatures ranging from 673 to 773 K (increments of 50 K) under (Config. 1 and Config. 2, Table 1) at temperatures ranging from 723 to 773 K
a feed pressure of 5 bar with a GHSV of 3000 cm3 g−cat1 h− 1 (corresponding to an (increments of 25 K) under a feed pressure of 5 bar with GHSVs ranging from
ammonia flowrate of 50 cm3 min− 1), and with a sweep N2 flowrate of 100 cm3 1200 to 6000 (increments of 600 cm3 g−cat1 h− 1 for Config. 1 and increments of
min− 1. A complete set of data comparing PBR and CMR at temperature from 1580 cm3 g−cat1 h− 1 for Config. 2), corresponding to an ammonia flowrate of
673 to 773 K with increments of 25 K is shown in Figs. S5b–c. 20–100 cm3 min− 1 with a sweep N2 flowrate of 100 cm3 min− 1.

7
Y. Park et al. Journal of Membrane Science 614 (2020) 118483

reported in the literature, the flux of the membrane did not reach its
maximum value, and there was room to increase the H2 flux if the partial
pressure of H2 was increased at a higher GHSV with higher catalyst
loading. The difference in the H2 flux between the two configurations
also affected the ammonia conversion: the higher NH3 conversion of
Config. 2 compared with that of Config. 1 was due to the higher H2 flux
for Config. 2 under the same GHSV, which resulted in faster catalytic
reaction kinetics and more effective removal of H2.
The performance of the two CMR configurations according to the
operating pressure was tested at a fixed NH3 flowrate (Fig. 8a). Despite
the approximately threefold difference in the effective permeation area
between the two configurations, the difference in the H2 permeation rate
was only approximately 80% at 773 K and 5 bar. The H2 permeation rate
of the CMR can be increased at higher GHSV, and the area of the
membrane was used more efficiently in Config. 2. The ammonia con­
version for Config. 2 was compared with that under the same GHSV
condition of the PBR in Fig. 8b, and the tendency was similar to the
comparison in Fig. 6b. However, in contrast to the case of Fig. 6b, where
the conversion increased with the pressure, the conversion remained
almost constant with the increasing pressure. As mentioned previously,
although removing the H2 can improve the reaction kinetics, in the high-
GHSV region, the ammonia conversion is not significantly increased,
owing to the low concentration of ammonia.
Fig. 9 shows a comparison of the H2 recovery yields for both the
tested CMR configurations. The H2 recovery yield with respect to the
GHSV exhibited similar trends for both configurations, but for Config. 2,
it decreased significantly after the GHSV reached 4700 cm3 g−cat1 h− 1. In
contrast, for Config. 2, the H2 recovery yield increased with the pressure,

Fig. 9. H2 recovery yield for two different CMR configurations (Config. 1 and
Config. 2) at temperatures ranging from 723 to 773 K (increments of 25 K) with
a sweep N2 flowrate of 100 cm3 min− 1 (a) under a feed pressure of 5 bar with
GHSVs ranging from 1200 to 6000 (increments of 600 cm3 g−cat1 h− 1 for Config. 1
and increments of 1580 cm3 g−cat1 h− 1 for Config. 2), corresponding to ammonia
flowrate of 20–100 cm3 min− 1, and (b) under feed pressures ranging from 1 to
5 bar at an NH3 feed rate of 50 cm3 min− 1 (GHSV: 3000 cm3 g−cat1 h− 1 for Config.
1 and 7900 cm3 g−cat1 h− 1 for Config. 2).

similar to the tendency of Fig. 8a. Increasing H2 permeation rate with


higher pressure strongly affected the H2 recovery yield, as shown in
Fig. 8a. The study on different CMR configurations clearly shows the
tradeoff relations between H2 production rate, H2 recovery, and NH3
conversion, as described above. These need to be considered in a large
scale system for process optimization, with reaction kinetic and mass
transfer simulations as conducted by previous studies [31,33,37].

3.3. Effects and benefits of steam as sweep gas in permeate stream

A sweep gas such as N2 or Ar on the permeate side facilitates the


permeation of H2 by increasing the driving force of the H2 transport
across the membrane, improving the H2 recovery yield as well as the
ammonia conversion. Figs. S6a–b shows the performance variation ac­
Fig. 8. (a) H2 permeation rates for the two different CMR configurations
(Config. 1 and Config. 2) and (b) NH3 conversion for PBR and CMR (Config. 2) cording to the N2 sweep gas flowrate in CMR Config. 1 at a pressure of 5
at temperatures ranging from 723 to 773 K (increments of 25 K) under feed bar and a GHSV of 3000 cm3 g−cat1 h− 1. H2 flux, ammonia conversion, and
pressures ranging from 1 to 5 bar with an ammonia feed rate of 50 cm3 min− 1 H2 recovery yield at all temperatures significantly increased with the N2
(GHSV: 3000 cm3 g−cat1 h− 1 for Config. 1 and 7900 cm3 g−cat1 h− 1 for Config. 2), sweep gas flowrates, and the increment of each performance metric was
and with a sweep N2 flowrate of 100 cm3 min− 1. marginal with a further increase in the sweep gas flowrates. Thus, the

8
Y. Park et al. Journal of Membrane Science 614 (2020) 118483

performance of the CMR can be improved through the use of N2 sweep major application of the produced H2. Using the Nessler method,
gas, but the sweep gas must eventually be separated via an additional quantitative titration of NH3 was conducted along with UV–vis spec­
unit process, such as pressure swing adsorption, to obtain high-purity troscopy. In Scheme 1, it was confirmed that the produced H2 contained
H2. approximately 4.2 ppm of ammonia with steam sweep gas. However, the
To solve this problem, steam was applied as an alternative sweep gas, amount of ammonia in the permeated product decreased to approxi­
and high-purity H2 was produced via relatively simple gas–liquid phase mately 1.8 ppm after the material passed through a gas–liquid separator
separation. There are few studies which adopted steam as a sweep gas in containing 300 cm3 of DI water with simple bubbling to mimic the effect
H2 extraction reactions [47,48], but its application to NH3 decomposi­ of a scrubber. Thus, a gas–liquid separator can act as a reservoir for
tion membrane reactor has not been previously reported. Fig. 10a shows steam sweep gas recirculation and prevent the poisoning of the PEMFC
the performance variation of the CMR with different steam sweep gas by removing trace amounts of ammonia from the product to ensure that
amounts at a pressure of 5 bar and GHSV of 6000 cm3 g−cat1 h− 1, compared it is fuel-cell-grade. Additionally, CMR operation under prolonged time
with the case where N2 sweep gas was used under the same conditions. of 65 h with steam sweep gas showed no significant decrease in the
At a high GHSV, the performance of the CMR increased with the sweep performance (Scheme 1). For further practical applications of the sug­
gas flowrates, similar to the trends presented in Fig. S6, and the per­ gested system, long term durability test of the system would be useful, as
formance difference between the two sweep gases was insignificant. stability of the Pd/Ta/Pd membrane can be problematic under high
Again, the benefits of using a sweep gas were apparent (increased H2 temperature operations [49–51]. Thermal stability of Pd-based mem­
recovery yield, NH3 conversion, and H2 permeation rate). In particular, branes is known to be affected by inter-diffusion [52] and surface
when the sweep gas was used, the H2 recovery rate was improved by morphology change, so strategies for stability improvement needs to be
approximately 5.5% compared with the case where the sweep gas was studied further by adopting diffusion barrier [53,54] or by alloying [55].
not used. Additionally, experiments were conducted with the steam To generate power using the as-produced H2, a commercial
sweep gas of 100 cm3 min− 1 at different feed pressures (Fig. 10b), and laboratory-scale PEMFC (12-W-class) was fed by the product H2 gas of
the trends of the performance increase were similar to those in Fig. S6. the CMR. As shown in Fig. 11a, the I–V curves were measured and
When the CMR was operated at a high GHSV (>5000 cm3 g−cat1 h− 1), as in compared with those for pure H2 and H2 produced by the CMR with N2
the case of Config. 2, the H2 recovery yield was highly dependent on the or steam sweep gases. Even though the same amount of H2 was intro­
operating pressure. duced into PEMFC, significant differences in the I–V curves were
For a membrane reactor, leak-free scalability is important, and there observed. When N2 was used as the sweep gas in the CMR, the voltage of
must be a way to cope with unexpected failures of the system. Leakage is the PEMFC was approximately 2.5% lower than that when pure H2 was
particularly critical for the NH3 decomposition membrane, as the pres­ used, owing to the lower H2 partial pressure on the anode side. However,
ence of NH3 in H2 gas is detrimental to PEMFC operation, which is the the voltage increased slightly when the H2 produced with steam sweep
gas was used. Because the H2 produced after passing through the
gas–liquid separator was humid, the PEMFC performance was increased.
The effects of anode humidification on proton conductivity of mem­
brane utilized in the current PEMFC are well established in previous
studies [56,57]. To validate the effect of humidification, pure H2 from
the gas cylinder was provided to PEMFC with and without bubbling as
schematic shown in Fig. 11b, and the performance increase with hu­
midification was similar as shown in Fig. 11a. Therefore, using steam as
the sweep gas is beneficial for onsite power-generation applications of
the composite membrane reactor.

3.4. Energy and mass balance evaluation of composite membrane reactor

The CMR process developed in this study requires the heat source to
replenish thermal energy owing to the endothermic reactions of NH3
decomposition and the latent heat of vaporization for the generation of
steam (as a sweep gas). Furthermore, utilities such as air and water are
necessary to complete the process. A process flow diagram (PFD) of the
CMR for ammonia decomposition and H2 production is presented in
Fig. 12 based on the experimental results at 773 K and 5 bar with GHSV
of 6000 cm3 g−cat1 h− 1 using and sweep steam flowrate of 100 cm3 min− 1.
To design the CMR, the model was divided into two units: a reformer for
ammonia decomposition and a membrane separator, and both of them
was assumed as steady-state, isothermal, and isobaric conditions. In the
PFD, the product gas stream (S1) of NH3 decomposition in a reformer is
linked to the membrane to purify the H2. The permeated H2 with steam
sweep gas from the membrane (S2) is introduced to gas-liquid (G-L)
separator after cooling down by a heat exchanger (S3). The retentate gas
after pass through the heat exchanger (S4) is mixed with air and supply
to the burner (S5).
The energy consumption and production of the process were evalu­
ated with respect to the process units. i) The heat loss due to the
Fig. 10. H2 permeation rate, H2 recovery yield, and NH3 conversion (a) for endothermic reaction of the ammonia decomposition reaction was
sweep gas flowrates ranging from 0 to 100 cm3 min− 1 with increments of 25 referred to as Q1. ii) To approach the reaction temperature of the sweep
cm3 min− 1 at 773 K and 5 bar with a GHSV of 6000 cm3 g−cat1 h− 1 and (b) for feed gas after the heat exchanged by retentate and permeate flows, an
pressures ranging from 1 to 5 bar at 773 K with a GHSV of 6000 cm3 g−cat1 h− 1 additional heat exchanger to replenish the energy (Q2) was applied. iii)
with a sweep steam flowrate of 100 cm3 min− 1. The retentate gas after the heat exchanger was combusted with air to

9
Y. Park et al. Journal of Membrane Science 614 (2020) 118483

Scheme 1. NH3 concentration in the permeate gas before and after passing through the vapor–liquid separator with long term testing of the CMR at conditions listed
in the scheme. Cross-sectional morphology of the fresh and spent membrane, after >60 h of testing, from previous study of our group [40].

amounts of water were considered. The amount of H2 produced by the


CMR (M1) was divided by the splitter to be used for the power gener­
ation by the PEMFC (M2) or the heat source of the process (M3). The
water consumed for generating the steam sweep gas (M4) included the
water recovered from the gas–liquid separator (M5) and the byproduct
of the electrochemical reaction in PEMFC (M6).
Table S1 presents the values of the mass and heat stream according to
the foregoing descriptions. Analyzing the energy balance using the
simulation tool revealed that a total energy of 5.285 W was required for
ammonia decomposition and steam generation through the heat
exchanger. However, when the retentate gas was used as a fuel for the
combustor, approximately 3.2 W of heat energy could be replenished for
the process. To balance the total caloric value in the process, some of the
H2 produced by the gas–liquid separator was utilized as a heat source.
When the retentate gas and approximately 16.7% of the generated H2
were used as the heat source of the process, they covered the entire
amount of energy required for the process. Additionally, if the generated
H2 is used for power generation for the PEMFC, it can produce
approximately 6.7 W of electricity.
The PEMFC generates water as a byproduct, which can be utilized for
the sweep gas of the CMR. As shown in the stream information of
M4–M6 in Table S1, the amount of water produced in the PEMFC is
similar to the amount required by the CMR, and most of the water can be
recovered through the gas–liquid separator. This is another benefit of
utilizing steam as a sweep gas in the proposed process. The present
simulation was performed to evaluate the feasibility of the process by
estimating the energy and mass balance; further process simulations and
experimental studies are needed to estimate the system efficiency at the
full scale.
Fig. 11. (a) I–V curve of the commercial PEMFC operated using pure H2 (68
cm3 min− 1) from a gas cylinder, as-produced H2 (68 cm3 min− 1) from the CMR
4. Conclusions
operated at 773 K, 5 bar, GHSV of 3000 cm3 g−cat1 h− 1 with N2 and steam sweep
gas (b) Normalized I–V curve with respect to open circuit voltage (OCV) using
the dry H2 and humidified H2 streams.
A Pd/BCC composite membrane was fabricated to be adopted for an
NH3 decomposition CMR. The as-fabricated membrane had a perme­
ability of 6.03 × 10− 8 mol m− 1 s− 1 Pa− 0.5, with a H2 flux of 10.3 cm3
supply the heat energy (Q3) required in the process. iv) The H2 produced
min− 1 cm− 2. These values were comparable to previously reported
by the process was partially fed to the combustor to replenish the energy
values for dense metallic membranes using BCC metals, and the mem­
(Q4). v) To generate electrical power, the produced H2 was used as a
brane was operated with high stability and without embrittlement is­
source for the conjugated PEMFC (Q5). For the material balance, the
sues. Then, Ru catalysts were synthesized, and their catalytic activity
amount of H2 produced by the CMR and the consumed/recovered
was evaluated at different temperatures, pressures, and GHSVs. Using

10
Y. Park et al. Journal of Membrane Science 614 (2020) 118483

Fig. 12. PFD of the composite membrane reactor for ammonia decomposition and H2 production.

the as-fabricated membranes and catalysts, different configurations of produced by the PEMFC. Finally, a process simulation was conducted,
membrane reactors were developed and operated in a region where confirming that waste energy from the integrated process can provide
further conversion increase could not be achieved solely by changing the the energy required to generate the steam and heat needed for endo­
operating conditions of a PBR. thermic NH3 decomposition.
Comprehensive experiments were performed at temperatures A higher H2 productivity, a lower operating temperature, faster ki­
ranging from 673 to 773 K, under feed pressures ranging from 1 to 5 bar, netics, and improved H2 purification for NH3 decomposition are
and with GHSVs ranging from 1200 to 6000 cm3 g−cat1 h− 1. For the as- demonstrated based on the developed CMR system in this study. How­
developed CMR, the NH3 conversion was >99% with a GHSV of 4800 ever, it has to be noted that the benefits could be marginal in some
cm3 g−cat1 h− 1 at 773 K, and >99% conversion was maintained up to a operating conditions; the adoption of CMR concept is meaningful in
GHSV of 3000 cm3 g−cat1 h− 1 at temperatures as low as 748 K. H2 pro­ specific operating conditions. Therefore, further analyses would be
ductivity improved more than 63% with CMR, with respect to PBR, required to assess potential benefits of CMR in terms of process inten­
benefitting from the in situ H2 extraction. Calculated H2 productivity per sification at higher scales. The proposed system utilizing steam sweep
catalyst loading indicated more than 2 times higher values than that of gas can widely be applied to the other membrane reactor configurations
previous studies of CMR using Ru catalyst, showing a significant with different membrane materials as well as to hydrogen extraction
improvement in the H2 productivity compared with previous results. A from other liquid hydrogen carriers including methanol, ethanol, formic
H2 recovery yield of >85% was observed over a wide range of GHSVs, acid, liquid organic hydrogen carriers, and so forth.
and the maximum H2 recovery yield of 91% was achieved at 773 K. With
an increase in the feed pressure from 1 to 5 bar at 723 K, the H2 recovery Credit author statement
yield improved from 77.5% to 88.7%, accompanied by a significant
increase in the H2 flux. YP: Conceptualization, Investigation, Writing – Original Draft,
While the NH3 conversion decreased with increasing pressure in the Formal analysis.
case of the PBR, it increased with pressure under the same operating JC: Conceptualization, Investigation.
conditions for the CMR. A significant increase in the NH3 conversion was HO: Investigation.
achieved under pressurized conditions (even when compared to the NH3 TL: Methodology.
conversion using the PBR at the atmospheric pressure) at all the tem­ SL: Resources.
perature ranges tested. Different CMR configurations with different MP: Resources.
membrane lengths and catalyst loadings were further compared with HJ: Data curation.
regard to the GHSV, pressure, temperature, and sweep gas effects. The YK: Visualization.
two configurations exhibited similar performance trends, but the HS: Validation.
configuration with a higher H2 flux resulted in higher NH3 conversion. SN: Project administration.
Considering that the membrane flux can be increased further, the H2 JH: Resources.
productivity can be improved with better catalyst design or higher CY: Funding Acquisition, Project administration, Writing – Review­
catalyst loading per unit membrane area achieved by changing the ge­ ing & Editing.
ometry of the membrane reactor. YJ: Conceptualization, Supervision, Writing – Original Draft, Writing
Then, steam was adopted as a sweep gas, yielding comparable per­ – Reviewing & Editing.
formance to cases using conventional sweep gases such as N2. However,
use of steam was beneficial for the following reasons. (i) The steam Declaration of competing interest
recirculation reservoir can act as an NH3 scrubber, reducing the NH3
content in the H2 stream. (ii) The reservoir can act as an NH3 buffer with The authors declare that they have no known competing financial
an in situ concentration measurement device in the case of membrane interests or personal relationships that could have appeared to influence
system failure. (iii) Wet H2 can be produced, eliminating the need for a the work reported in this paper.
humidifier, and the DI water reservoir can be replenished using water

11
Y. Park et al. Journal of Membrane Science 614 (2020) 118483

Acknowledgements [21] G. Bernardo, T. Araújo, T. da Silva Lopes, J. Sousa, A. Mendes, Recent advances in
membrane technologies for hydrogen purification, Int. J. Hydrogen Energy 45 (12)
(2019) 7313–7338, 4.
This work was supported financially by a National Research Foun­ [22] M. Temkin, V. Pyzhev, Kinetics of ammonia synthesis on promoted iron catalysts,
dation (NRF) grant funded by the Korean Government (Ministry of Acta Physicochim. URSS 12 (1940) 217–222.
Science, ICT & Future Planning) [grant number NRF- [23] J.P. Collins, J.D. Way, Catalytic decomposition of ammonia in a membrane reactor,
J. Membr. Sci. 96 (1994) 259–274.
2019M3E6A1064611], by a Korea Institute of Energy Technology [24] E.N. Gobina, J.S. Oklany, R. Hughes, Elimination of ammonia from coal
Evaluation and Planning (KETEP) grant funded by the Korean Govern­ gasification streams by using a catalytic membrane reactor, Ind. Eng. Chem. Res.
ment (Ministry of Trade, Industry & Energy) [grant number 34 (1995) 3777–3783.
[25] M.E.E. Abashar, Y.S. Al-Sughair, I.S. Al-Mutaz, Investigation of low temperature
20183010042020], and by the KIST Institutional Program of the Korea decomposition of ammonia using spatially patterned catalytic membrane reactors,
Institute of Science and Technology [grant numbers 2E30520, Appl. Catal. Gen. 236 (2002) 35–53.
2E30202]. [26] J. Zhang, H. Xu, W. Li, High-purity COx-free H2 generation from NH3 via the ultra
permeable and highly selective Pd membranes, J. Membr. Sci. 277 (2006) 85–93.
[27] F.R. García-García, Y.H. Ma, I. Rodríguez-Ramos, A. Guerrero-Ruiz, High purity
Appendix A. Supplementary data hydrogen production by low temperature catalytic ammonia decomposition in a
multifunctional membrane reactor, Catal. Commun. 9 (2008) 482–486.
[28] S.H. Israni, B.K.R. Nair, M.P. Harold, Hydrogen generation and purification in a
Supplementary data to this article can be found online at https://doi. composite Pd hollow fiber membrane reactor: experiments and modeling, Catal.
org/10.1016/j.memsci.2020.118483. Today 139 (2009) 299–311.
[29] G. Li, M. Kanezashi, T. Tsuru, Highly enhanced ammonia decomposition in a
bimodal catalytic membrane reactor for COx-free hydrogen production, Catal.
References
Commun. 15 (2011) 60–63.
[30] G. Li, M. Kanezashi, H.R. Lee, M. Maeda, T. Yoshioka, T. Tsuru, Preparation of a
[1] A. Klerke, C.H. Christensen, J.K. Nørskov, T. Vegge, Ammonia for hydrogen novel bimodal catalytic membrane reactor and its application to ammonia
storage: challenges and opportunities, J. Mater. Chem. 18 (2008). decomposition for COx-free hydrogen production, Int. J. Hydrogen Energy 37
[2] T. He, P. Pachfule, H. Wu, Q. Xu, P. Chen, Hydrogen carriers, Nat. Rev. Mater. 1 (2012) 12105–12113.
(2016). [31] N. Itoh, A. Oshima, E. Suga, T. Sato, Kinetic enhancement of ammonia
[3] R.F. Service, Liquid sunshine, Science 361 (2018) 120–123. decomposition as a chemical hydrogen carrier in palladium membrane reactor,
[4] F. Zhou, L.M. Azofra, M. Ali, M. Kar, A.N. Simonov, C. McDonnell-Worth, C. Sun, Catal. Today 236 (2014) 70–76.
X. Zhang, D.R. MacFarlane, Electro-synthesis of ammonia from nitrogen at ambient [32] E. Rizzuto, P. Palange, Z. Del Prete, Characterization of an ammonia decomposition
temperature and pressure in ionic liquids, Energy Environ. Sci. 10 (2017) process by means of a multifunctional catalytic membrane reactor, Int. J. Hydrogen
2516–2520. Energy 39 (2014) 11403–11410.
[5] T.E. Bell, L. Torrente-Murciano, H2 production via ammonia decomposition using [33] Z. Zhang, S. Liguori, T.F. Fuerst, J.D. Way, C.A. Wolden, Efficient ammonia
non-noble metal catalysts: a review, Top. Catal. 59 (2016) 1438–1457. decomposition in a catalytic membrane reactor to enable hydrogen storage and
[6] A.M. Karim, V. Prasad, G. Mpourmpakis, W.W. Lonergan, A.I. Frenkel, J.G. Chen, utilization, ACS Sustain. Chem. Eng. 7 (2019) 5975–5985.
D.G. Vlachos, Correlating particle size and shape of supported Ru/gamma-Al2O3 [34] M.R. Rahimpour, A. Asgari, Production of hydrogen from purge gases of ammonia
catalysts with NH3 decomposition activity, J. Am. Chem. Soc. 131 (2009) plants in a catalytic hydrogen-permselective membrane reactor, Int. J. Hydrogen
12230–12239. Energy 34 (2009) 5795–5802.
[7] Y. Li, L. Yao, Y. Song, S. Liu, J. Zhao, W. Ji, C.T. Au, Core-shell structured [35] M.Á. Gómez-García, I. Dobrosz-Gómez, J. Fontalvo, J.M. Rynkowski, Membrane
microcapsular-like Ru@SiO2 reactor for efficient generation of CO(x)-free reactor design guidelines for ammonia decomposition, Catal. Today 191 (2012)
hydrogen through ammonia decomposition, Chem. Commun. (Camb) 46 (2010) 165–168.
5298–5300. [36] M.E.E. Abashar, Ultra-clean hydrogen production by ammonia decomposition,
[8] X. Li, W. Ji, J. Zhao, S. Wang, C. Au, Ammonia decomposition over Ru and Ni J. King Saud Univ.- Eng. Sci. (2016).
catalysts supported on fumed SiO2, MCM-41, and SBA-15, J. Catal. 236 (2005) [37] M.E.E. Abashar, The impact of ammonia feed distribution on the performance of a
181–189. fixed bed membrane reactor for ammonia decomposition to ultra-pure hydrogen,
[9] L. Li, Z.H. Zhu, G.Q. Lu, Z.F. Yan, S.Z. Qiao, Catalytic ammonia decomposition over Int. J. Hydrogen Energy 44 (2019) 82–90.
CMK-3 supported Ru catalysts: effects of surface treatments of supports, Carbon 45 [38] Y.S. Jo, C.H. Lee, S.Y. Kong, K.-Y. Lee, C.W. Yoon, S.W. Nam, J. Han,
(2007) 11–20. Characterization of a Pd/Ta composite membrane and its application to a large
[10] X. Ju, L. Liu, P. Yu, J. Guo, X. Zhang, T. He, G. Wu, P. Chen, Mesoporous Ru/MgO scale high-purity hydrogen separation from mixed gas, Separ. Purif. Technol. 200
prepared by a deposition-precipitation method as highly active catalyst for (2018) 221–229.
producing COx-free hydrogen from ammonia decomposition, Appl. Catal. B [39] C.H. Lee, Y.S. Jo, Y. Park, H. Jeong, Y. Kim, H. Sohn, C.W. Yoon, S.W. Nam, H.
Environ. 211 (2017) 167–175. C. Ham, J. Han, Unconventional hydrogen permeation behavior of Pd/BCC
[11] S.F. Yin, B.Q. Xu, W.X. Zhu, C.F. Ng, X.P. Zhou, C.T. Au, Carbon nanotubes- composite membranes and significance of surface reaction kinetics, J. Membr. Sci.
supported Ru catalyst for the generation of COx-free hydrogen from ammonia, 595 (2020).
Catal. Today 93–95 (2004) 27–38. [40] Y.S. Jo, J. Cha, C.H. Lee, H. Jeong, C.W. Yoon, S.W. Nam, J. Han, A viable
[12] S.F. Yin, B.Q. Xu, X.P. Zhou, C.T. Au, A mini-review on ammonia decomposition membrane reactor option for sustainable hydrogen production from ammonia,
catalysts for on-site generation of hydrogen for fuel cell applications, Appl. Catal. J. Power Sources 400 (2018) 518–526.
Gen. 277 (2004) 1–9. [41] P.P. Mardilovich, Y. She, Y.H. Ma, M.-H. Rei, Defect-free palladium membranes on
[13] S.-F. Yin, Q.-H. Zhang, B.-Q. Xu, W.-X. Zhu, C.-F. Ng, C.-T. Au, Investigation on the porous stainless-steel support, AIChE J. 44 (1998) 310–322.
catalysis of COx-free hydrogen generation from ammonia, J. Catal. 224 (2004) [42] B. Lee, J. Park, H. Lee, M. Byun, C.W. Yoon, H. Lim, Assessment of the economic
384–396. potential: CO -free hydrogen production from renewables via ammonia
[14] P. Xie, Y. Yao, Z. Huang, Z. Liu, J. Zhang, T. Li, G. Wang, R. Shahbazian-Yassar, decomposition for small-sized H2 refueling stations, Renew. Sustain. Energy Rev.
L. Hu, C. Wang, Highly efficient decomposition of ammonia using high-entropy 113 (2019).
alloy catalysts, Nat. Commun. 10 (2019) 4011. [43] M.D. Dolan, Non-Pd BCC alloy membranes for industrial hydrogen separation,
[15] L. Huo, B. Liu, H. Li, B. Cao, X.-c. Hu, X.-p. Fu, C. Jia, J. Zhang, Component synergy J. Membr. Sci. 362 (2010) 12–28.
and armor protection induced superior catalytic activity and stability of ultrathin [44] J. Ni, R. Wang, F. Kong, T. Zhang, J. Lin, b. Lin, K. Wei, Highly efficient Ru-Ba/AC
Co-Fe spinel nanosheets confined in mesoporous silica shells for ammonia catalyst promoted by magnesium for ammonia synthesis, Chin. J. Catal. 32 (2011)
decomposition reaction, Appl. Catal. B Environ. 253 (2019) 121–130. 436–439.
[16] J. Cha, Y.S. Jo, H. Jeong, J. Han, S.W. Nam, K.H. Song, C.W. Yoon, Ammonia as an [45] C.F. Huo, B.S. Wu, P. Gao, Y. Yang, Y.W. Li, H. Jiao, The mechanism of potassium
efficient CO X -free hydrogen carrier: fundamentals and feasibility analyses for fuel promoter: enhancing the stability of active surfaces, Angew Chem. Int. Ed. Engl. 50
cell applications, Appl. Energy 224 (2018) 194–204. (2011) 7403–7406.
[17] F.A. Uribe, S. Gottesfeld, T.A. Zawodzinski, Effect of ammonia as potential fuel [46] J. Liu, X. Ju, C. Tang, L. Liu, H. Li, P. Chen, High performance stainless-steel
impurity on proton exchange membrane fuel cell performance, J. Electrochem. supported Pd membranes with a finger-like and gap structure and its application in
Soc. 149 (2002). NH3 decomposition membrane reactor, Chem. Eng. J. 388 (2020).
[18] Y.A. Gomez, A. Oyarce, G. Lindbergh, C. Lagergren, Ammonia contamination of a [47] H. Wang, P. Kölsch, T. Schiestel, C. Tablet, S. Werth, J. Caro, Production of high-
proton exchange membrane fuel cell, J. Electrochem. Soc. 165 (2018) F189–F197. purity oxygen by perovskite hollow fiber membranes swept with steam, J. Membr.
[19] H. Miyaoka, H. Miyaoka, T. Ichikawa, T. Ichikawa, Y. Kojima, Highly purified Sci. 284 (2006) 5–8.
hydrogen production from ammonia for PEM fuel cell, Int. J. Hydrogen Energy 43 [48] Y. Zhang, Q. Sun, X. Gu, Pure H2production through hollow fiber hydrogen-
(2018) 14486–14492. selective MFI zeolite membranes using steam as sweep gas, AIChE J. 61 (2015)
[20] K.E. Lamb, D.M. Viano, M.J. Langley, S.S. Hla, M.D. Dolan, High-purity H2 3459–3469.
produced from NH3 via a ruthenium-based decomposition catalyst and vanadium- [49] R.E. Buxbaum, T.L. Marker, Hydrogen transport through non-porous membranes of
based membrane, Ind. Eng. Chem. Res. 57 (2018) 7811–7816. palladium-coated niobium, tantalum and vanadium, J. Membr. Sci. 85 (1993)
29–38.

12
Y. Park et al. Journal of Membrane Science 614 (2020) 118483

[50] K.S. Rothenberger, B.H. Howard, R.P. Killmeyer, A.V. Cugini, R.M. Enick, [54] Y. Huang, R. Dittmeyer, Preparation of thin palladium membranes on a porous
F. Bustamante, M.V. Ciocco, B.D. Morreale, R.E. Buxbaum, Evaluation of tantalum- support with rough surface, J. Membr. Sci. 302 (2007) 160–170.
based materials for hydrogen separation at elevated temperatures and pressures☆, [55] E. Yan, R.N. Min, P. Zhao, R.D.K. Misra, P.R. Huang, Y.J. Zou, H.L. Chu, H.
J. Membr. Sci. 218 (2003) 19–37. Z. Zhang, F. Xu, L.X. Sun, Design of Nb-based multi-phase alloy membranes for
[51] V.N. Alimov, A.O. Busnyuk, M.E. Notkin, E.Y. Peredistov, A.I. Livshits, Hydrogen high hydrogen permeability and suppressed hydrogen embrittlement, J. Membr.
transport through V–Pd alloy membranes: hydrogen solution, permeation and Sci. 595 (2020).
diffusion, J. Membr. Sci. 481 (2015) 54–62. [56] H. Sun, G. Zhang, L.-J. Guo, S. Dehua, H. Liu, Effects of humidification
[52] D.J. Edlund, J. McCarthy, The relationship between intermetallic diffusion and flux temperatures on local current characteristics in a PEM fuel cell, J. Power Sources
decline in composite-metal membranes: implications for achieving long membrane 168 (2007) 400–407.
lifetime, J. Membr. Sci. 107 (1995) 147–153. [57] S. Kim, I. Hong, Effects of humidity and temperature on a proton exchange
[53] Y. Huang, R. Dittmeyer, Preparation and characterization of composite palladium membrane fuel cell (PEMFC) stack, J. Ind. Eng. Chem. 14 (2008) 357–364.
membranes on sinter-metal supports with a ceramic barrier against intermetallic
diffusion, J. Membr. Sci. 282 (2006) 296–310.

13

You might also like