Sie sind auf Seite 1von 13

Applied Energy 240 (2019) 359–371

Contents lists available at ScienceDirect

Applied Energy
journal homepage: www.elsevier.com/locate/apenergy

Demonstrating direct methanation of real biogas in a fluidised bed reactor T


Julia Witte, Adelaide Calbry-Muzyka, Tanja Wieseler, Peter Hottinger, Serge M.A. Biollaz,

Tilman J. Schildhauer
Energy & Environment Research Division, Paul Scherrer Institut, 5232 Villigen PSI, Switzerland

H I GH L IG H T S

• Catalytic direct methanation of biogas was demonstrated with real biogas.


• Stable operation of the fluidised bed methanation for over 1100 h.
• Biomethane with an average methane yield of 96% was injected into the gas grid.
• Slow deactivation by organic sulphur compounds was identified and solved.
• Experimental results were in accordance with results from a rate-based model.

A R T I C LE I N FO A B S T R A C T

Keywords: The catalytic direct methanation of biogas to produce biomethane was conducted in a pilot plant with real biogas
Catalytic methanation from a biogas plant in Zurich. Stable operation of the methanation system including a bubbling fluidised bed
Long duration experiment could be demonstrated for over 1100 h of regular operation with subsequent injection into the gas grid. An
Deactivation average methane yield of 96% was reached. During the long-duration experiment, the slow deactivation process
Power-to-gas
was monitored and found to be only moderate. Organic sulphur compounds could be identified as the main
Biogas upgrading
Rate based modelling
source of deactivation. However, deactivation from coking could not be fully excluded. With appropriate
measures in the gas cleaning unit, so that no sulphur compounds were measured subsequently (limit of detec-
tion: < 0.05 ppm of relevant components), the yield reduction went close to zero. Additionally, experimental
results were compared to simulation results from a rate-based model presented elsewhere. Model predictions and
experimental results are in accordance. The kinetics used in the rate-based model allowed a correct prediction of
the optimum reaction temperature found by experimental results.

1. Introduction with knowledge transfer from research to industry.


While methanation of pure CO2 separated from biogas or other
The storage of excess energy from renewable sources is becoming sources has been shown in the past [6], only few attempts were done to
more important with an increasing share of renewable energies in the test methanation of biogas without prior separation of the CO2. Stan-
energy system, which is targeted by countries like Germany or geland et al. [7] developed a nickel based catalyst with ruthenium
Switzerland [1,2]. A widespread concept for the storage of fluctuating dopants and showed its activity and stability in methanation of simu-
electric energy at large scale is the Power-to-Gas (PtG) method [3] lated biogas. Based on the research of the Centre for Solar Energy and
which can supplement other storage technologies such as pumped Hydrogen Research (ZSW), the company Solarfuel GmbH (later known
hydro power [4]. For this, core technologies like electrolysers [5] and as etogas and now a subsidiary of Hitachi-Zosen Inova) built a cooled
methanation reactors [6] connect the electricity grid with the gas grid, fixed bed reactor and tested it at 25 kW scale at different biogas plants
so that electricity is converted into storable gases like methane to a [8]. Catalytic methanation of biogas in adiabatic fixed bed reactors has
large extent. However, those core technologies are often just at the also been tested at lab scale in Austria [9] and at pilot scale by Haldor-
beginning of being applied in industry due to their low maturity. For Topsøe A/S in Denmark [10]. Jürgensen [11] investigated the effect of
further application in industry, it must be shown that these technologies ammonia, a typical constituent of biogas, and found that it has no ne-
are feasible under real conditions for long-term operation, combined gative effect on methanation catalysts.


Corresponding author at: OVGA 110, CH-5232 Villigen PSI, Switzerland.
E-mail address: tilman.schildhauer@psi.ch (T.J. Schildhauer).

https://doi.org/10.1016/j.apenergy.2019.01.230
Received 13 September 2018; Received in revised form 24 January 2019; Accepted 28 January 2019
0306-2619/ © 2019 Elsevier Ltd. All rights reserved.
J. Witte, et al. Applied Energy 240 (2019) 359–371

Nomenclature p pressure [barg or bara]


T temperature [°C]
i index for components [–] t time [h]
j index for time step [–] Vi̇ standard volume flow [Nl min−1]
mcat catalyst mass [kg] Y yield [%]
ṅ i mole flow of component i [mol s−1] Yred yield reduction [%/100 h]

Besides catalytic methanation at > 300 °C, biological methanation biogas plant in Zurich, at Werdhölzli. Biogas contains, besides trace
of biogas is also being investigated by several groups [12,13]. Two components, mainly methane (∼60 vol%) and carbon dioxide (∼40 vol
companies, Microbenergy/Viessmann and electrochaea, have built pilot %). For an unrestricted injection into the gas grid, the carbon dioxide
scale plants that also convert real biogas [14] by applying pressurised must be removed to a minimum methane concentration of 96 vol% to
stirred bubble columns. Rachbauer et al. [15] report a long duration obtain biomethane [25,26]. Conventionally, the carbon dioxide is se-
test converting real biogas with hydrogen by mesophilic biological parated from the methane via CO2 scrubbing or a gas separation
methanation at < 40 °C. They successfully reached > 98% methane in membrane. After this, the carbon dioxide is released into atmosphere as
the product stream using a counter-current trickle bed reactor at am- waste. A new way of biogas upgrading includes the application of a PtG
bient pressure with space velocities orders of magnitude smaller than system. Here, the biogas enters a methanation reactor where carbon
typical for catalytic methanation reactors. dioxide is converted to further methane together with hydrogen from
Further studies discuss the system aspects of biogas upgrading for an electrolyser fuelled by excess renewable electricity. The following
specific regions, e.g. [16] and compare CO2 separation by amine chemical equations are considered:
scrubbing with methanation of separated CO2 and direct methanation
0 kJ
of the biogas without prior separation of CO2 and methane [17]. While CO2 + H2 ↔ CO + H2 O ΔHreac = +41.16
mol (1)
separation of CO2 from biogas by amine scrubbing and its subsequent
release to the atmosphere is the cheapest way to produce bio-methane kJ
0
due to the high costs of electrolysers, direct methanation of biogas 3H2 + CO ↔ CH4 + H2 O ΔHreac = −206.28
mol (2)
without prior separation of CO2 outperforms methanation of separated
CO2 by efficiency and costs [17–20]. Studies on the environmental With this method, the biomethane production can be increased by
impact of biogas methanation show that significant greenhouse gas 60% and CO2 emissions are avoided. A set-up for the methanation of
savings can be achieved by the technology, if renewable electricity is biogas, named COSYMA (COntainer-based SYstem for MethAnation),
used [21–24]. was built with a pilot-scale size of 10–20 kW chemical output.
This work focuses on direct methanation of biogas in a bubbling The scope of this work includes the demonstration of stable opera-
fluidised bed methanation reactor for Power-to-Gas applications, which tion of direct methanation with real biogas in a long-duration experi-
has significant advantages with respect to heat removal compared to ment of over 1000 h. For the first time, all scientific and technical de-
fixed bed reactors. tails of a long duration test of catalytic methanation are presented
A cooperation was started with the natural gas supplier energie360° including thorough discussion of deactivation phenomena, suitable
in Switzerland so that the feasibility of a fluidised bed methanation unit solutions and optimisation of operation conditions based on experi-
could be demonstrated under real conditions with raw biogas from their ments and model predictions. For further upscaling and application in
industry, demonstrating feasibility of this Power-to-Gas technology

Fig. 1. Simplified flowsheet of the COSYMA set-up.

360
J. Witte, et al. Applied Energy 240 (2019) 359–371

under real conditions for long-term operation is a necessary and im- The reacted wet gas mixture leaves the reactor and enters the particle
portant step. filter unit, where particles are removed from the gas and are analysed.
In a next step, the wet gas mixture is cooled, so that the water in the
reacted gas condenses. The condensate is directed to a tank, which is
2. Methods weighed by scale II. With this procedure, it is possible to determine the
water flow entering and leaving the reactor. The difference of these two
2.1. Experimental set-up streams is equal to the water produced during reaction, which contains
information about the present conversion. The dry gas leaving the
A simplified version of the pilot-scale COSYMA set-up is illustrated condenser is analysed again continuously via mGC and a non-dispersive
in Fig. 1. The set-up can be operated either by synthetic bottled gas or infrared (NDIR, Siemens Ultramat) sensor. If the NDIR sensor displays
by real biogas coming from the digesters of a biogas plant. Between 1.4 permissible concentrations of methane (CH 5%), the gas, which is now
and 2.3 N m3/h of real or synthetic biogas entered the set-up. Starting called ‘biomethane’, flows through a gas meter (standard gas meter for
from real biogas, the stream first is compressed to the desired system natural gas) into the natural gas grid as a restricted injection. If the
pressure and then directed through an orifice in order to determine the methane content was below the mentioned limit, the gas left the set-up
volumetric flow. After this, the biogas enters the gas cleaning unit, as exhaust and the plant switched into stand-by mode.
where harmful trace components like organic sulphur are removed. The COSYMA set-up was transported to a biogas plant of energie360°
These contaminants lead to deactivation of the catalyst even in small in Zurich at Werdhölzli where it was connected to the existing biogas
amounts, therefore it is essential to remove them for stable operation of flow. An overview of the biogas plant with the COSYMA set-up can be
the methanation reactor. It was possible to measure the relevant con- seen in Fig. 2. The biogas is produced from two different sources: bio-
taminants until a limit of detection below 0.1 ppm. A detailed de- waste and sewage sludge. Then, the gases are directed into one gas tank
scription of the gas cleaning unit can be found in [27]. Subsequently (gasometer). About 1400 standard cubic meters per hour of biogas
biogas is mixed with hydrogen and helium from the bottled gas section. enters the conventional upgrading unit, where CO2 is separated by
The mass flows of bottled gases are measured and controlled by mass amine scrubbing and released. The obtained biomethane (about
flow controllers (MFC). The addition of a known amount of helium 860 N m3/h, depending on the CO2 content in the biogas) is injected
enables the determination of the flows of all components via con- into the gas grid. After the biogas tank, a slip stream (about 1–2 N m3/
centration measurements. Next, a small volume flow of the gas mixture h) is directed to the COSYMA setup where CO2 is converted to methane
is continuously directed to a micro gas chromatograph (mGC, Varian in combination with hydrogen. Then, the product gas stream is mixed
CP-4900) to measure bulk gas concentrations including helium. Ap- again with the main biomethane stream and is injected into the gas
propriate and repeated calibration of the microGC for for helium en- grid.
sured that the error for the flow rate calculations were kept o a
minimum. Then the stream is preheated to the reactor inlet temperature
and mixed with steam. For this, water from a closed tank is heated to 2.2. Measurement system and data evaluation
360 °C.
The decreasing mass of water in the tank is measured by scale I, Operational conditions, which can be measured directly, are re-
which allows the determination of the inlet mass flow of water. Water corded and visualised by an Intermodulation Analysis System (IAS).
addition is intended to prevent coking of the catalyst. Now the wet gas This includes temperatures, pressures, weights, concentrations from the
mixture enters the bubbling fluidised bed reactor, where the metha- NDIR sensor etc. (155 measurement points in total). Calculations of
nation reaction takes place with a nickel catalyst of Geldart particle operational parameters with data from IAS or mGC are executed se-
type B. The reactor diameter is 5.2 cm and the catalyst mass was 800 g parately. During the whole operation,
at the beginning of operation. The corresponding non-fluidised bed a small flow of helium (0.4 Nl/min) was added to the biogas flow,
height is 58 cm. Inside the reactor, two lances are present. One lance which made it possible to derive volumetric flow rates from the con-
measures temperatures at different heights. The other lance dis- centration measurements. A redundant measurement system was es-
continuously takes gas samples at 7 cm bed height. These samples are tablished, such that the same relevant operational condition could be
directed to a mGC to measure bulk gas concentration including helium. measured or calculated in multiple ways independently from each

Fig. 2. Overview of the biogas plant in Zurich-Werdhölzli (Biogas Zürich – energie360°) including the demonstration plant COSYMA.

361
J. Witte, et al. Applied Energy 240 (2019) 359–371

Table 1 was reduced stepwise to zero. Here, the deactivation of the catalyst
Operational conditions during reference experiments with bottled gas. caused by coking was investigated in order to evaluate which additional
H2, Nl/ CO2, Nl/ N2, Nl/ He, Nl/ H2/ H2O/ p, barg Treac, °C amount of water at the inlet is necessary to protect the catalyst.
min min min min CO2, – CO2, –

53.4 12.8 22.1 0.4 4.17 0.86 6.1 320; 350


2.3.2. Monitoring of deactivation
A set of reference experiments were performed under kinetically
limited conditions at repeated intervals over the course of the experi-
other. This allows a direct verification of the redundantly obtained mental campaign to monitor catalyst deactivation. Since the con-
values for one operational condition. For this, a measurement data centrations in the biogas were subjected to fluctuations, only bottled
evaluation tool (MeDEa) programmed in Matlab© was established. gas was used for the reference experiments in order to achieve stable
First, the MeDEa tool merges the data from the different sources. For conditions. The reference experiments were distributed over the 1130 h
this, the same time stamp must be applied to every data source. In order of regular operation and were repeated about every 100 operational
to obtain a value within two measured data points, linear interpolation hours (starting after 400 h) or when there was a specific incident like
between those two data points was done. Second, non-directly mea- catalyst addition etc. An average reference casegas composition of
surable operational conditions (like inlet volume flow of biogas, flui- about 37 vol% CO2 and 63 vol% nitrogen with bottled gas as feed was
disation in the methanation reactor, conversion of CO2 etc.) were cal- chosen to realise conversions which are at least partly limited by in-
culated via IAS and mGC data. Third, mass balances for C, H and O were trinsic kinetics. The reference experiments were carried out at two
executed for the whole set-up. temperature levels. The predominant conditions of the reference ex-
The data evaluation procedure allowed an identification of reliable periments are listed in Table 1.
measurement values by which mass balances could be closed with an In addition, eight catalyst samples were taken over time from the
average error of ± 5%. Redundantly obtained values are in agreement bed material of the reactor in order to analyse the depositions on the
with each other. A prompt tracking and detailed evaluation of the catalyst surface. This analysis was done by temperature programmed
course of operation over 1130 h of operation became possible with the oxidation (TPO) in a thermo-gravimetric analyser (TGA; STA 449 F3
help of the automated MeDEa tool. Here, early insights could be ob- from NETZSCH) at temperature levels between 150 °C and 900 °C. The
tained during operation which allowed the optimisation of operation gases formed during heating under diluted air were measured with a
parameters and the identification and rectification of irregularities mass spectrometer (MS; Omnistar GSD 301 01 from Pfeiffer). With this
during operation at an early stage. method, it is possible to distinguish between carbon species and sulphur
species which block the catalyst surface and are responsible for the
catalyst deactivation.
2.3. Experiments

2.3.1. Regular operation 2.4. Assessment of rigorous methanation model


Regular operation hours are defined as the time during which real
biogas from the plant in Werdhölzli was directed to the COSYMA set-up One additional target is comparing the predictions of the existing
and biomethane produced by the reactor was injected into the gas grid. rate-based model of a bubbling fluidised bed for methanation with the
More than 1130 regular operation hours could be completed. The experimental data obtained during regular operation. The applied
regular operation time can be divided into three phases. model is taken from [28] with the adaptions made in [19]. In general,
In the first phase (0 h–429 h), operational conditions were changed the model considers thermodynamic and kinetically driven effects for
in order to find optimal conditions for maximum methane content in the reaction as well as the interplay of reaction and hydrodynamic in-
the reactor outlet gas. For this, the reactor temperature and the hy- fluences caused by bubble formation and interphase mass transfer from
drogen feed were varied. The amount of hydrogen added to the system the bubbles to the catalyst phase. The model is a classical two zone
is expressed via the hydrogen-to-CO2 ratio, where the hydrogen feed is model where the bubble phase is considered as one plug flow, and the
placed in relation to the carbon dioxide molar flow in the raw biogas. gas moving through the catalyst particles is considered as a second
With this ratio, variations in the biogas feed are reflected although the pseudo-homogeneous phase, also with plug flow for the gas. Hydro-
hydrogen feed was constant. The pressure was set to 5.7 barg for the dynamic correlations describe the size and rise velocities of the bubbles,
whole regular operation time. Steam was added to the reactor in order while Langmuir-Hinshelwood type kinetics are used to predict the
to prevent deactivation of the catalyst from coke formation. Here, an chemical conversion on the catalyst particles within the pseudo-
overall molar ratio of water-to-CO2 of 0.5 was chosen as a safe measure homogeneous dense phase. A detailed description of the bubbling
to avoid carbon deposition based on experience from previous long fluidised bed model can be found in [19].
duration campaigns. This way, the feed is also just outside the graphite For the assessment, the model is tested via variations in temperature
formation region according to thermodynamic equilibrium. and hydrogen addition to the biogas. With this, the kinetics of the
In the second phase (430 h–1035 h), operational conditions were model could be evaluated as well as the influence of the hydrodynamics
held constant within the range of the optimal conditions for tempera- on the reacting system. The experimental and rate based model results
ture and hydrogen-to-CO2 ratio identified in the first phase. Pressure are compared to an equilibrium model for methanation based on
and water-to-CO2 ratio remained the same as in the first phase. The minimising Gibbs free energy for the species CO, CO2, CH4, H2, H2O
compressor for the biogas feed was not temperature-controlled. Hence, (ASPEN Plus®).
the temperature of the compressor was subject to ambient temperature
variations from day to night, which resulted in changing output stan- 3. Results and discussion
dard volume flows. In order to maintain a constant hydrogen-to-CO2
ratio within a tolerated range, the hydrogen flowrate had to be re-ad- The pilot-scale methanation system COSYMA ran in total 1131
justed over time. However, fluctuations of the hydrogen-to-CO2 ratio regular operating hours. In addition, eleven reference experiments were
could not be fully avoided. Changes in the compressor flow affect carried out and eight catalyst samples were taken from the reactor over
various other operational conditions e.g. the fluidisation in the reactor time. The corresponding large amount of data was processed with the
and the water-to-CO2 ratio. automated MeDEa tool with which inconsistent data could be identified
During the last 96 operating hours, which represent the third phase, immediately so that the causes of the inconsistencies could be removed
a stress test for the catalyst was executed. For this, the water addition directly.

362
J. Witte, et al. Applied Energy 240 (2019) 359–371

3.1. Regular operation first signs of deactivation occured with the start of a slow decrease in
methane concentration at the outlet. However, the exact starting time
The operation of the COSYMA set-up was stable over the entire of deactivation cannot be identified since the first measurements of
operating time with average concentrations in the outlet gas of 87 vol% non-H2S sulphur compounds breaking through the gas cleaning unit
methane, 10.6 vol% hydrogen, 1.4 vol% carbon dioxide and 1 vol% were executed at 300 h, and the first reference experiment with used
helium. Helium was only added for advanced measurements which are catalyst was conducted at 400 h operation time. Deactivation processes
not necessary for an industrial operation. The average yield of methane are described in more detail in Section 3.2.
was 95.8%. In Figs. 3 and 4, operational conditions and the perfor-
mance of the reactor are illustrated. The different ratios and the yield of
methane used in the figures and elsewhere are defined as follows: 3.1.2. Second experimental phase (operating at optimal conditions)
ṅ H2, in During the second phase (430 h–1035 h), operational parameters
H2
= were mainly set constant at around optimum conditions. Temperature
CO2 ṅCO2, biogas (3) and pressure were set to 355 °C and 5.7 barg, respectively. The op-
H2 O ṅ H2 O, in timum temperature increased due to deactivation issues. The average
= catalyst stress and the average H2O/CO2 ratio were 12 NlCO2,biogas/
CO2 ̇ 2, biogas
nCO (4)
(min kgcat) and 0.55, respectively. Since it was not possible to hold the
̇ out4
nCH − ̇ in 4
nCH H2/CO2 ratio constant due to the changes of the compressor flow, the
YCH4 = ratio varies between 3.85 and 4.15 (Fig. 4). In the second phase, me-
̇ in 2
nCO (5)
thane yields reached 93 vol% to 98 vol% with methane outlet gas
̇ std , biogas concentrations between 84 vol% and 88 vol%. At 400 operating hours,
VCO
catalyst stress = 2
an abrupt decrease of methane content by 2.2 vol% – points in the
mcat (6) biomethane stream can be observed (Fig. 3). At this time, the operation
In Fig. 3, the dry molar fractions of the bulk components methane, was paused for about two weeks, and the system was purged with ni-
hydrogen and carbon dioxide in the outlet gas are illustrated as a trogen during the shut-down. The nitrogen passed through the gas
function of the regular operation time. Here, important incidents during cleaning unit and then entered the methanation reactor, so that it is
operation are marked. In Fig. 4, the corresponding ratio of hydrogen-to- very likely that sulphur present on the adsorber in the gas cleaning unit
carbon dioxide at the inlet and the yield of methane are shown as a desorbed to some extent and then entered the reactor together with the
function of the regular operating hours. The values illustrated in both nitrogen stream. The sulphur adsorbed on the catalyst, which led to the
figures are averaged over one operating day. observed abrupt deactivation. From 400 to 600 h, stable operation of
the COSYMA set-up was obtained with an average level of 85.4 vol%
3.1.1. First experimental phase (optimisation) methane concentration in the biomethane flow. However, small de-
During the first phase (0 h–429 h), the reactor temperature was creases of the methane concentrations in the biomethane flow were
changed between 320 °C and 360 °C, and for a short time up to 380 °C. observed over time. The reference experiments also showed signs of
The system pressure was set constant. Especially during the first 200 h, deactivation in the reactor (see Section 3.2). Therefore, after 641 op-
the temperature and the hydrogen-to-CO2 ratio (Fig. 4) in the inlet gas erating hours, 150 g of fresh catalyst were added to the used catalyst
was varied significantly to identify optimal operation settings. A de- (800 g) in the reactor, which improved the yield of methane so that
tailed description of the optimisation process can be found in Section methane concentrations of 86.5 vol% were reached again. Over the next
3.3. Due to the pronounced changes of these parameters, the dry con- few 100 operation hours, measures were taken in the gas cleaning unit
centration of the bulk components in the outlet gas (Fig. 3) and the which improved the gas cleaning performance. More specifically, the
methane yield (Fig. 4) strongly vary during the first 200 operational chosen sorbent materials were regenerated and the entering biogas flow
hours. In the first phase, yields reached values between 94 vol% and was actively cooled to minimise competing adsorption between water
99 vol% with corresponding methane concentrations in the outlet and sulphur molecules [27]. These measures were successful, so that
stream between 85 vol% and 90 vol%. Almost the entire amount of the methane concentration in the biomethane flow remained constant
reacted carbon dioxide was converted to methane. Hence, a selectivity at approximately 86 vol% until the end of this second phase.
of nearly 100% can be assumed. After 200 regular operating hours, the

Fig. 3. Molar fractions (dry) of bulk components after the methanation reactor as a function of regular operation hours.

363
J. Witte, et al. Applied Energy 240 (2019) 359–371

Fig. 4. Inlet ratio of hydrogen to carbon dioxide from biogas and the yield of methane as a function of regular operation hours.

3.1.3. Third experimental phase (decreasing steam addition) deactivation effects.


In the third phase (1036 h–1131 h), the steam addition to the re- Still, operational conditions were continuously changing over the
actor was stepwise reduced to observe the effect of steam on the re- time, which impedes a clear interpretation of the results over the op-
action yield. On the one hand, steam influences the thermodynamic eration hours regarding deactivation. In order to obtain reliable in-
equilibrium towards less methane production. On the other hand, steam formation about the effects of deactivation, reference experiments were
should protect the catalyst against coking, since a higher ratio of ele- conducted.
mental hydrogen in comparison to carbon is achieved in the reactor.
Reducing the steam addition can demonstrate whether steam addition
is necessary for a stable operation. In Fig. 3, the effect of steam re- 3.2. Monitoring of deactivation
duction can be seen in the bulk concentrations. For about 50 h, the
steam was reduced by half, and for the last 50 h no steam was added to For the reference experiments, dry molar concentrations of the
the reactor. Due to the improved thermodynamic equilibrium, the components methane, hydrogen and carbon dioxide after the reactor
methane concentrations increased with the decreasing steam addition. are illustrated in Fig. 5 at a reactor temperature of 350°C. The con-
Hence, no severe deactivation occurred during this phase, so that the centrations of the mentioned components are normalised to 100% so
more beneficial thermodynamic equilibrium has its intended effect on that nitrogen and helium are not illustrated in the graph. The yield of
the conversion. methane is also shown for the corresponding experiment. At hour zero,
In the following sections, however, it will be discussed that the a reference experiment was conducted before real biogas entered the
methane content would have been higher, if no deactivation had oc- COSYMA set-up. This point can be seen as the reference state for a non-
curred. In general, changes in catalyst stress within the given ranges deactivated catalyst with a yield of 96.1 vol% and a normalised me-
have only a minor influence on the reactor performance whereas var- thane concentration after the reactor of 82.5 vol% at corresponding
iations of hydrogen addition and of temperature clearly impact the dry conditions. The normalised concentration of methane is not directly
outlet concentration of the reactor. The yield of methane correlates comparable with the results of Section 3.1 since only the produced
directly with the hydrogen-to-CO2 ratio (Fig. 4), aside from methane is shown in Fig. 5 and not the sum of methane from biogas and
the produced methane as in Fig. 3. In Fig. 6, the corresponding yield

Fig. 5. Normalised dry molar fraction of bulk components after the reactor and the yield of methane as a function of regular operation hours with a reference gas
mixture at a reactor temperature of T = 350 °C.

364
J. Witte, et al. Applied Energy 240 (2019) 359–371

Fig. 6. Yield reduction per hundred operation hours as a function of regular hours of operation at 320 °C and 350 °C measured during reference experiments.

reduction per 100 h observed during the reference experiments is is the limit for a sufficient conversion rate in the reactor.
shown as a function of regular operating hours at reactor temperatures In the last phase the steam addition was reduced stepwise to zero.
of 320°C and 350°C. The yield reduction per time step j is defined as Although the methane content during regular operating hours incresed
follows: during this last phase (Fig. 3), a significant deactivation can be ob-
served (Figs. 5 and 6). The positive effect of a more beneficial ther-
Yj ⎞ Yred, j Yred, j 100h
Yred, j = ⎜⎛1 − ⎟ ·100% = · modynamic equilibrium could not completely be overruled by the de-
⎝ Y j−1 ⎠ 100h t j − t j − 1 100h activation process during regular operation. However, whether the
deactivation was caused by coking due to the lack of steam in the last
This definition is equivalent to a deactivation gradient or deacti-
phase or again by sulphur compounds cannot by identified only by
vation rate over time. At about 400 h of regular operation, the first
comparing deactivation rates in full-steam and no-steam modes.
reference experiment was conducted with used catalyst. A decreased
yield can be observed together with a decreased normalised con-
centration of methane after the reactor, see Fig. 5. Correspondingly, the
3.2.1. Concentration profile measurements
hydrogen and carbon dioxide content increases due to the lower con-
Concentrations inside the bed height were measured during several
version rates. After further regular operation hours, deactivation pro-
reference experiments (Fig. 7). Here, the dry molar fractions of the bulk
gresses until reaching a yield of 91.3 vol% at about 630 h. Several
components methane, hydrogen and carbon dioxide are illustrated.
breakthroughs of sulphur compounds in the gas cleaning unit could be
After 408 operational hours, the first lance measurement was con-
identified as source of the deactivation as shown in this section. The
ducted when the reactor had already suffered a performance loss due to
gradient of deactivation increased during this period and forms a
deactivation as discussed earlier. Then, before (630 h) and after the
maximum at 630 h with a value of 1.8%/100 h at 350 °C and 2.4%/
catalyst addition (640 h), the corresponding profiles are shown. In
100 h at 320 °C. The gradient of deactivation at 320 °C is always higher
general, approximately two thirds of the total methane production is
than at 350 °C, which is explained in the next section.
achieved by 7 cm of bed height, which is about 12% of the total height.
As a countermeasure to the growing deactivation gradient, 150 g of
In the following 88% of bed height, the remaining one third of total
new catalyst were added to the 800 g of catalyst already inside the
methane production is achieved which most probably is slowed by mass
reactor at about 640 operating hours. As a result, yield and methane
transport limitations. Normalised concentration profiles changed over
concentration increased again, achieving a level between the initial
time due to deactivation. For the first three measurements in time
state and the reference experiment at 400 h. Less than one fifth of the
(408 h, 540 h and 630 h), the normalised concentration of methane
original catalyst amount was added to the reactor, but this resulted in
decreases, hence hydrogen and carbon dioxide concentrations increase
almost in the same performance of the reactor as for the non-deacti-
both at the lance height and at the end of bed. After the catalyst ad-
vated catalyst. Since the methanation reaction in a fluidised bed is not
dition at 640 h, the methane content increases. Due to the previously
mainly restricted by kinetics, but among others by heat transfer at the
discussed deactivation, the methane concentration over the bed height
given operational conditions, the main part of the catalyst bed is used to
decreases again with time.
dissipate heat via heat exchangers, so that approximately isothermal
It was expected that the difference of concentrations for one com-
conditions are reached. The 150 g of new catalyst, a relatively small
ponent at different points in time would be more pronounced at the
amount, is still kinetically active enough to almost compensate for the
lance height than at the total bed height. Therefore, this method was
partial deactivation of the used catalyst. Nevertheless, the catalyst
intended to give an early warning of deactivation. However, the mea-
material inside the reactor slowly deactivated again as time progressed.
surements show that this is hardly the case. Mass transfer limitations
As mentioned before, adaptations in the gas cleaning unit were made to
from bubble to dense phase may influence the reaction in a way that the
improve the gas cleaning performance. In the period between 914 h and
reaction course is dampened; therefore deactivation of the catalyst
1035 h of regular operation, the reference experiments show that al-
cannot be seen so clearly at the lance. The catalyst particle movement
most no further deactivation occurred.
in the reactor might also contribute to the similar concentration pro-
The yield reduction for both temperature levels almost reached zero
files. Deactivated particles are evenly distributed in the reactor volume;
with 0.25%/100 h. With this deactivation rate, it would be possible to
therefore, the reaction rates are evenly decreased over the entire bed
run the methanation with a yearly replacement of the catalyst assuming
height.
8000 h of operation per year. After one year, 80% of the total catalyst
material will be deactivated. Simulations have shown that this amount

365
J. Witte, et al. Applied Energy 240 (2019) 359–371

Therefore, coking is most likely not the main cause for deactivation of
the catalyst used for biogas conversion even for the operating time
where reduced or no steam was added to the reactor.
Between 500 °C and 800 °C, a second peak of the MS signal for
carbon dioxide can be observed (Fig. 8(a)). Here, graphite (CC ) from the
catalyst structure itself is combusted. The graphite content of the cur-
rent and the former catalyst (Güssing 2007) is not exactly equal,
therefore different peak intensities are obtained. The last mass change
in Fig. 8(a) can be observed between 800 °C and 900 °C
(300 min–400 min), which is due to the release of sulphur dioxide.
The corresponding normalised MS signal for sulphur dioxide can be
seen in Fig. 9. For the currently used catalyst samples, only one peak is
observed between 800 °C and 900 °C which is directly related to sulphur
on the catalyst surface. It can be seen that with progressing time, the
formed peaks become higher with the same width. Hence, more sulphur
was adsorbed at the catalyst surface over the operating time, which
corresponds to the previous deactivation monitoring results. For the
unused catalyst, no sulphur signal was obtained. The catalyst used in
2007 showed a higher deposition of sulphur than the samples from the
biogas methanation. Here, also a second peak between 500 °C and
800 °C can be observed, where most likely the sulphur situated on the
graphite was burned together with the graphite from the catalyst ma-
terial [32] according to the carbon dioxide signal from Fig. 8(a) at the
mentioned temperature level. The concentration of sulphur con-
taminants in the gasifier gas from wood was significantly higher than in
the biogas [32], which explains the larger integral of the MS signals for
SO2 of the catalyst sample from 2007.
Fig. 7. Normalised (dry) concentration profiles of bulk components over bed
height during reference experiments at 350 °C reactor temperature;; measure-
ment positions: inlet, lance height (7 cm), outlet; the lines are for orientation 3.3.1. Synopsis of sulphur related measurements
only. In Fig. 10, results from different measurements regarding sulphur
poisoning of the catalyst during biogas methanation are illustrated.
3.3. Analysis of catalyst samples Here, the breakthrough concentrations of total sulphur after the gas
cleaning unit [27] as a function of the operating hours are compared to
Catalyst samples and breakthrough concentrations after the gas the relative mass loss of the used catalyst samples during the TPO
cleaning unit were analysed to identify the source of deactivation. The analysis at 800 °C–900 °C, which is directly related to sulphur deposi-
catalyst samples taken from the reactor were analysed by Temperature tion on the catalyst surface as explained in the previous section. Fur-
Programmed Oxidation (TPO) to identify which compounds are bound ther, the yield of methane during the reference experiments is shown
to the catalyst surface t. In Figs. 8 and 9, the results of the TPO analysis again as a function of the operating hours. The gas phase total sulphur
are presented for the investigation of carbon and sulphur depositions. plotted in Fig. 10 is a minimum value based on the sum of quantified
Catalyst samples were frequently taken over the operation time. Ad-
ditionally, the fresh catalyst used in the COSYMA set-up and a sample
from a former methanation experiment converting wood gasifier gas in
Güssing (Austria) in 2007 [29,30] were investigated.
In Fig. 8(b), the temperature profile applied for all TPO samples is
illustrated. In Fig. 8(a), the corresponding relative mass change of the
investigated catalyst samples are shown together with the MS signal for
carbon dioxide normalised by the catalyst sample mass. From 20 °C to
150 °C, the mass decreases for every sample mainly due to a drying of
previously adsorbed humidity. A first peak in the CO2 signal can also be
seen, which might correspond to the combustion of adsorbed carbon
(Cα) or desorption of CO2. At higher temperatures, the masses of most
samples increase until a temperature of 360 °C, which can be explained
by oxidation of the bulk nickel in the catalyst sample. At the same time,
small CO2 signals indicate combustion of polymeric carbon [31]. The
mass of the Güssing 2007 sample, however, further decreases with
rising temperatures, and a significantly higher amount of carbon di-
oxide is released. Hence, increasing mass due to oxidation of the bulk
nickel and decreasing mass due to polymeric carbon combustion are in
competition, and the mass difference cannot give a precise result for the
amount of carbon on the catalyst surface. However, the signals for
carbon dioxide between 150 °C and 500 °C, which corresponds to Fig. 8. (a) Relative mass change of catalyst samples and MS measurement
polymeric carbon (Cβ ), are not significant for the current samples in signal of CO2 during TPO analysis for used catalyst samples
comparison to the Güssing 2007 sample. This 2007 catalyst sample was (03.05.2017–06.07.2017), fresh catalyst and for a catalyst sample taken from a
still active for methanation, although it had been subjected to a more methanation project using wood gasifier gas in 2007 [29]; (b) corresponding
challenging feed gas composition containing wood gasifier tars. temperature profile during TPO.

366
J. Witte, et al. Applied Energy 240 (2019) 359–371

experiments is correspondingly reduced again. The addition of steam


was reduced at this time, but significant deactivation due to carbon
deposition can be excluded. The TPO analysis showed only minor ef-
fects of carbon deposition and the yield reduction can be explained
directly by the sulphur poisoning. However, the yield reduction at the
end of operation was globally at a maximum (Fig. 6) although only a
moderate amount of 0.5 ppmv of total sulphur was entering the reactor.
This is an indication that sulphur was not the only cause of the deac-
tivation, but carbon depositions were also present due to the lack of
steam.
In general, the total sulphur concentration in the biogas feed should
be significantly lower than 0.5 ppmv over a year so that a maximum
deactivation rate of 0.25%/100 h allows a yearly exchange of the cat-
alyst material with the chosen space velocity. The space velocity, and
thus the sulphur concentration limit, is based on a reactor size which is
designed for a sufficient heat transfer. As a result, approximately 20%
of the catalyst material is required for the conversion assuming no re-
levant mass transfer limitations, and the other 80% are necessary for
Fig. 9. MS measurement signal of sulphur dioxide during TPO analysis for used the heat dissipation, creating a catalyst buffer to alleviate deactivation
catalyst samples (03.05.2017–06.07.2017), fresh catalyst and for a catalyst effects.
sample taken from a methanation project from wood gasifier gas in 2007 [29].
3.4. Model evaluation

The model evaluation was conducted with data from the operation
optimisation experiments. For this, the ratio of hydrogen and carbon
dioxide in the feed and the reactor temperature were varied in the
model and during the experiments (regular operation – first phase). The
other parameters which influence the methanation reaction (H2O/CO2,
catalyst stress and reactor pressure) were kept constant within a narrow
range. The water-to-CO2 ratio and the catalyst stress varied in a range
from 0.51 to 0.67 and from 13.3 Nl/(min kg) to 17.1 Nl/(min kg), re-
spectively, due to different ambient conditions, whereas the reactor
pressure was constant at 5.7 barg. The experimental data for the ana-
lysis were taken during the regular operation hours and are compared
to the results of the equilibrium and rate-based methanation models.
Dry molar fractions of the bulk components methane, hydrogen and
carbon dioxide downstream of the reactor are illustrated as a function
of the inlet ratio H2/CO2 ratio in Fig. 11. After the reactor, a particle
filter system was placed (Fig. 2) running at temperatures between
Fig. 10. Breakthrough concentrations of total sulphur after the gas cleaning 215 °C and 243 °C. Here, a small amount of catalyst material was
unit, relative mass loss of catalyst samples related to sulphur during TPO
analysis and the methane yield at reference experiments as a function of regular
operation hours, adapted from [27].

sulphur compounds detected by the sampling and analytical devices in


use during the long-duration experiment [27].
All three indicators correlate with each other. During a sulphur
breakthrough, the concentrations of total sulphur in the biogas were
between 0.5 and 3 ppmv. With increased gas phase sulphur con-
centrations, the relative mass loss of the corresponding catalyst sample
increased. Simultaneously, the yield of methane decreased as a result of
catalyst deactivation due to sulphur. At about 600 operating hours, new
catalyst was added. The used catalyst was diluted with fresh catalyst so
that the specific sulphur deposition per catalyst mass decreased, which
improved the yield temporarily. After that, a third breakthrough of
sulphur can be observed, which is reflected again in the mass loss
during TPO and in the methane yield. During the operation hours be-
tween 750 h and 900 h, only small amounts of sulphur entered the re-
actor, so that no change of mass loss can be seen. But these small
amounts of sulphur seemed to be sufficient to cause deactivation, since
the yield decreased further. Only at the operating hours between 900 h Fig. 11. Dry molar fractions of bulk components after the reactor over different
and 1050 h, where no sulphur was detected in the feed gas, the yield hydrogen-to-CO2 inlet ratios derived via experiment and equilibrium model for
stayed approximately stable. During the last 100 operating hours, the the methanation at T = 340 °C and for the filter at 243 °C; average H2O/
CO2 = 0.55, average catalyst stress = 16.4 Nl/(min kg), biogas concentrations
concentration of sulphur components increased again and with that the
of 63.8% methane and 36.2% CO2 and p = 5.7 barg, experimental results taken
relative mass loss in the TPO analysis. The yield of the reference
from the first 300 regular operating hours.

367
J. Witte, et al. Applied Energy 240 (2019) 359–371

present which could cause some reaction even at these low tempera- outlet are shown, as experimentally obtained for different H2/CO2 inlet
tures [33]. In order to reflect this situation, the equilibrium con- ratios. Model results are illustrated with a full line and a dashed line as
centrations at filter temperature are also illustrated. a maximum and minimum case, because during the experiment, the
For high H2/CO2 inlet ratios, larger amounts of hydrogen remain in water-to-CO2 inlet ratio and the catalyst stress varied in a certain range.
the outlet stream of the reactor; vice versa, the fraction of carbon di- The experimental results should be situated within the range bound by
oxide decreases with the H2/CO2 ratio. For high hydrogen addition, the the maximum and minimum case of the rate-based model. The input
outlet carbon dioxide fraction is at a minimum, because larger amounts parameters of the two cases are listed in Table 2. For the model case
of hydrogen enhance the reaction and therefore the conversion of CO2. with maximum methane concentration at the outlet, a low H2/CO2 inlet
With decreasing hydrogen addition, the reaction becomes more re- ratio was chosen together with a low steam addition and a low catalyst
stricted by the lack of hydrogen and less carbon dioxide converts, so stress of 0.7 N m3 h−1 kg−1 cat . At the beginning of this section, it was
more CO2 remains in the outlet stream of the reactor. The methane already shown that sub-stoichiometric H2/CO2 inlet ratios until 3.9
fraction forms a slight maximum. For increased hydrogen addition, cause higher methane contents at the outlet, than stoichiometric or
high conversion rates to methane are obtained, but more hydrogen also hyper-stoichiometric hydrogen addition. Lower water addition also
remains in the outlet flow. For decreased hydrogen addition, less hy- influences the thermodynamic equilibrium beneficially towards higher
drogen remains in the outlet flow, but the conversion to methane is methane concentration. A lower catalyst stress and in general a smaller
inhibited. The interplay of these factors results in an optimum value of inlet volume flow results in higher reaction performance and less mass
H2/CO2 ratio where the conversion rates to methane and low hydrogen transfer limitations. For the case with minimum methane concentra-
addition are balanced. This optimum can be found at an H2/CO2-ratio tions at the outlet, by contrast a hyper-stoichiometric hydrogen addi-
of 3.9 in the model as well as from experimental results. However, sub- tion was chosen, high steam addition and a high catalyst stress value
stoichiometric hydrogen addition may increase the risk of catalyst de- (1.1 N m3 h−1 kg−1 cat ). All parameters in Table 2 were chosen from the
activation due to coking. range of actual experimental conditions corresponding to the experi-
For lower H2/CO2–ratios, the conversion rates in the experiments mental results shown in Fig. 14. The experimental results illustrated in
are even higher than the equilibrium model of the reactor predicts. Fig. 14 form a characteristic curve with a maximum for the methane
Further conversion of carbon dioxide might have taken place in the concentration at a certain temperature. At lower temperatures the
subsequent filter, but did not reach equilibrium (compare experimental catalyst is less active and low methane concentrations are obtained. The
data points with the dashed line in Fig. 11), most probably due to the reaction does not reach the thermodynamic limit due to kinetic lim-
low temperatures and small catalyst amounts in the filter. Second, a itations.
temperature profile was established in the reactor during experiments At higher temperatures, the thermodynamic limit for the methane
as illustrated in Fig. 12 that is not considered in the (isothermal) concentration is reached but the limit also decreases with the tem-
equilibrium and the rate-based models. perature due to the exothermic character of the reaction. Hence, a
Higher temperatures at the beginning of the reactor bed result in a maximum of methane concentration is formed, where the reaction
higher activity of the catalyst, which leads to faster conversion. If then course limited by the kinetics reaches the thermodynamic limit. In
downstream, the reactor temperature decreases, thermodynamically general, the kinetics of the rate-based model can predict the activity of
more beneficial conditions are obtained, which results in further but the catalyst. The experimental results are mostly in between the two
also slower conversion of the remaining carbon dioxide due to the de- simulated cases. There is a small tendency of higher experimental va-
creased activity at lower temperatures. lues than predicted values due to conversion in the filter as discussed
In Fig. 13, the same experimentally obtained methane concentra- before. The influence of the H2/CO2 inlet ratio on the experimentally
tions are shown as in Fig. 11. In addition, aside from the results of the obtained methane outlet concentration is not as clear as expected. Here,
equilibrium model the results of the rate-based model are also illu- changing conditions in the reactor due to fluctuating catalyst stress and
strated as a function of the H2/CO2 inlet ratio. A lower filter tempera- thus changing H2O/CO2 inlet ratios, temperature profiles with a spread
ture (215 °C) results in slightly higher conversion rates, so that the of 5 K in average, as well as partial deactivation of the catalyst made it
experimentally found methane concentrations are slightly increased in difficult to obtain comparable results. However, trends for the position
comparison to the methane concentrations at a filter temperature of of the optimal temperature can be derived. Like in the model, the op-
243 °C. This indicates not only that more conversion than predicted by timal temperature increases with the H2/CO2 ratio. As expected, the
the thermodynamic equilibrium was reached in the reactor due to the data points from experiments have the tendency to be closer to the
temperature profiles in the reactor, but also that further reaction oc- model results of the maximum case for low H2/CO2 ratios, while with
curred in the filter. The rate-based model and the equilibrium model increasing H2/CO2 ratio, experimental data points evolve more towards
show differences in their results for the reactor. For higher H2/CO2 inlet
ratios, the equilibrium model predicts higher methane concentrations
than the rate-based model. The rate-based model considers hydro-
dynamic effects. With increasing hydrogen addition, the fluidisation of
the reactor bed increases and with that the average bubble size. Larger
bubbles worsen the interphase mass transfer from the bubbles to the
catalyst phase so that the interphase mass transport of the reactants
takes more time, which reduces the reaction rates. This results in a
steeper decrease of the predicted methane concentration, which also
can be seen in the experimentally obtained methane concentrations.
In this case, the rate-based model provides more precise results than
the equilibrium model because hydrodynamic effects are considered.

3.4.1. Assessment of the applied kinetics


To assess the kinetics applied in the rate-based model, the tem-
perature in the reactor during the regular operation was varied, see
Fig. 14. Experimental results are compared with results of the rate-
based model for bubbling fluidised bed methanation as a function of the Fig. 12. Temperature profile over the COSYMA reactor height for different
reactor temperature. Dry molar fractions of methane at the reactor control temperatures.

368
J. Witte, et al. Applied Energy 240 (2019) 359–371

Fig. 15. Theoretical dry molar fraction of methane after the reactor as a
Fig. 13. Dry molar fractions of methane after the reactor over different hy-
function of the temperature for a standard inlet gas composition with simulated
drogen-to-CO2 inlet ratios derived via experiment, rate-based model and equi-
fresh catalyst and deactivated catalyst and without water addition and fresh
librium model for the methanation at T = 340 °C and for the filter at 243 °C or
catalyst; the results are obtained by the rate-based model and the equilibrium
215 °C; average H2O/CO2 = 0.6, average cat stress = 15.5 Nl/(min kg) and
model.
p = 5.7 barg, experimental results taken from the first 300 regular operating
hours.
simulation results.

3.4.2. Influence of deactivation and steam addition on optimal operation


temperature
After 430 regular operating hours, the optimal temperature was
readjusted from 340 °C to 355 °C in order to obtain again a higher
conversion rate. Here, the effect of deactivation on the optimal tem-
perature could be observed. This effect is described in Fig. 15 via si-
mulation results from the rate based model and the equilibrium model.
Here, dry concentrations of methane after the reactor are shown as a
function of the temperature for a standard case (H2/CO2 = 4.0, H2O/
CO2 = 0.5), a deactivated standard case, and a case with no steam
addition (H2/CO2 = 4.0, H2O/CO2 = 0).
The case with partly deactivated catalyst shows a reduction in ac-
tivity in comparison to the standard case, so thatthe kinetics reach the
optimum at the thermodynamic limit only at higher temperatures. It
also shows why the yield reduction is more pronounced at 320 °C than
at 350 °C. The concentration difference between the standard case and
Fig. 14. Dry molar fractions of methane after the reactor as a function of the
the deactivated case becomes less with increasing temperatures until
temperature, experimentally obtained at different H2/CO2 ratios and simulated
both cases reach the thermodynamic equilibrium. For the case of no
with the rate based model for a minimum and maximum case.
steam addition, a new equilibrium is established above the standard
case. As a result, a lower optimal temperature is achieved and a higher
Table 2 maximum methane concentration.
Input parameters corresponding to experimental conditions of the cases max- During the last phase of regular operation, the steam addition was
imum and minimum methane output concentrations for the rated-based me-
stepwise reduced from H2O/CO2 = 0.5 to zero. In Fig. 16, the dry bulk
thanation model.
concentrations after the reactor are shown as a function of the H2O/CO2
Case H2/CO2 H2O/CO2 VCH4,in, Nl/ VCO2,in, Nl/ Pressure, barg ratio at the inlet via experimental data points and from the rate based
(inlet) (inlet) min min and the equilibrium model. The lack of steam addition results in higher
Max 3.95 0.47 17.36 9.85 5.71
methane contents as a consequence of the more beneficial equilibrium.
Min 4.10 0.51 24.94 14.15 5.71 The equilibrium model shows the same result as the rate-based
model. This means that the reaction was limited neither by mass
transfer nor by insufficient catalyst activity due to low temperatures
the minimum case. The kinetics used in the model were derived from and could reach the maximum conversion rate, which is then only re-
experiments where CO methanation was investigated and at opera- stricted by thermodynamics. The experimental results are consequently
tional conditions far away from this work (H2/CO = 5–6, p = 1 barg) below the model results, which can be explained by the deactivation of
[34]. Nevertheless, experimental data are in good agreement with the the catalyst over time. However, the gradient of the methane con-
rate-based model indicating that combining water-gas-shift reaction centrations are approximately the same for the model and experimental
and CO – methanation allows a description of CO2 – methanation. The results.
optimal temperature was identified to be between 340 °C and 350 °C for
the given range of operating conditions both from experimental and

369
J. Witte, et al. Applied Energy 240 (2019) 359–371

accuracy of the experiment. In addition, the rate based model could


reflect the consequences of changing fluidisation in the reactor.
In summary, the demonstration of the direct methanation of real
biogas in a long-term experiment was successful. Product concentra-
tions close to equilibrium were reached. With an appropriately adjusted
gas cleaning unit, the deactivation of the catalyst was minor, which
would allow an exchange of the whole catalyst material only once per
year. The attrition rates of the catalyst particles due to fluidisation were
small, so that in total only 70 g of the catalyst material was lost in the
particle filter. All three aspects are beneficial not only for operational
but also for economic reasons. A successful long duration experiment
with a real feedstock at a Technical Readiness Level of 5 is an important
and necessary step towards up-scaling and implementation of this in-
novative technology. The simulation results showed that important
aspects of optimum operation conditions can be properly predicted,
limiting the risks during upscaling. Pilot and demonstration units can
be based on these results with respect to gas cleaning technology, re-
actor sizing and optimum operation conditions.

Fig. 16. Dry molar fractions of bulk components after the reactor as a function Acknowledgments
of water addition, experimentally obtained (last 200 regular operating hours)
and simulated with the rate-based model and the equilibrium model. This research project is financially supported by the Swiss
Innovation Agency lnnosuisse and is part of the Swiss Competence
4. Conclusions Center for Energy Research SCCER BIOSWEET. The authors would like
to acknowledge financial and other support from the Swiss Federal
The direct methanation of biogas was tested with a set-up operating Office for Energy (SFOE), Energie360°, FOGA and the ESI Platform.
on-site with real biogas from a biogas plant in Zurich. The methanation Biogas Zürich and ERZ are gratefully acknowledged for providing the
reactor used was a bubbling fluidised bed, which is expected to be less biogas source and access to the site location. Alwin Frei and Sergio
sensitive towards deactivation due to carbon depositions than a fixed Rodriguez Duran are sincerely thanked for their contribution to oper-
bed reactor. Stable operation over 1100 h could be demonstrated with ating the methanation plant; Hansjörg Wagenbach, Daniel Meyer,
an average yield of 96% and average bulk concentrations of 88 vol% Marcel Hottiger are acknowledged for their contribution to build the
CH4, 11 vol% H2 and 1 vol% CO2 in the product gas, which is close to plant.
equilibrium. The obtained product gas quality is in accordance with
model predictions. However, for an unrestricted injection into the gas References
grid, the remaining hydrogen in the product gas must be separated in a
next step, which can be done via a gas separation membrane, or con- [1] Feredal Ministry for Economic Affairs and Energy of Germany, Gesetz zur
Einführung von Ausschreibungen für Strom aus Erneuerbaren Energien und zu
verted in a second methanation reactor after intermediate water con-
weiteren Änderungen des Rechts der Erneuerbaren Energien. Bundesgesetzblatt.
densation to shift equilibrium [19]. Optimised operating conditions for 2016; 1: 2258–360.
the COSYMA set-up could be found supported by modelling/simula- [2] Piot M. Energiestrategie 2050 der Schweiz. 13. Symp. Energieinnovation. 2014. p.
21–3.
tions. During the last 100 operating hours, a stress test was carried out
[3] Schiebahn S, Grube T, Robinius M, Tietze V, Kumar B, Stolten D. Power to gas:
with a stepwise reduction of steam addition. Additional steam should to technological overview, systems analysis and economic assessment for a case study
protect the catalyst from coking in order to inhibit deactivation pro- in Germany. Int J Hydrogen Energy 2015;40:4285–94. https://doi.org/10.1016/j.
cesses. Even without any steam addition during 50 h, no significant ijhydene.2015.01.123.
[4] de Boer HS, Grond L, Moll H, Benders R. The application of power-to-gas, pumped
carbon depositions were found on the catalyst surface. hydro storage and compressed air energy storage in an electricity system at different
Deactivation processes could be monitored via reference experi- wind power penetration levels. Energy 2014;72:360–70. https://doi.org/10.1016/j.
ments and catalyst samples taken frequently over the operating time. energy.2014.05.047.
[5] Gahleitner G. Hydrogen from renewable electricity: an international review of
The main source for the catalyst deactivation were sulphur compounds, power-to-gas pilot plants for stationary applications. Int J Hydrogen Energy
which broke through the gas cleaning unit and entered the reactor with 2013;38:2039–61. https://doi.org/10.1016/j.ijhydene.2012.12.010.
minimum concentrations between 0.5 and 3 ppmv. A direct relation [6] Götz M, Lefebvre J, Mörs F, Mc Daniel Koch A, Graf F, Bajohr S, et al. Renewable
power-to-gas: a technological and economic review. Renew Energy
could be shown between breakthrough concentrations of total sulphur, 2016;85:1371–90. https://doi.org/10.1016/j.renene.2015.07.066.
the relative mass of sulphur from the catalyst samples (TPO method) [7] Stangeland K, Kalai DY, Li H, Yu Z. Active and stable Ni based catalysts and pro-
and the progression of the decreasing yield during reference experi- cesses for biogas upgrading: the effect of temperature and initial methane con-
centration on CO2 methanation. Appl Energy 2018;227:206–12.
ments. Significantly less than 0.5 ppmv of sulphur in the biogas feed
[8] Specht M, Brellochs J, Frick V, Stürmer B, Zuberbühler U. The power-to-gas (P2G®)
stream to the reactor is needed for a yield reduction of 0.25% per 100 process: storage of renewable energy in the natural gas grid via fixed bed metha-
operation hours, which would allow a yearly exchange of the catalyst nation of CO2/H2. In: Schildhauer TJ, Biollaz SMA, editors. Synthetic natural gas
from coal, dry biomass and power-to-gas applications. John Wiley & Sons; 2016.
material for a reactor sizing adapted to heat transfer management.
[9] Kirchbacher F, Biegger P, Miltner M, Lehner M, Harasek M. A new methanation and
Carbon depositions contributed to the deactivation processes only membrane based power-to-gas process for the direct integration of raw biogas –
marginally. feasability and comparison. Energy 2018;146:34–46.
Two different methanation models were compared with experi- [10] Hansen John Bøgild. SOEC enabled biogas upgrading. Presentation at the ETIP
SNET’s northern region workshop, Riga, December 7. 2017.
mental data from the COSYMA set-up. Here, the main target was the [11] Jürgensen L, Ehimen EA, Born J, Bo Holm-Nielsen J, Rooney D. Influence of trace
assessment of the applied kinetics in the rate based model. A compar- substances on methanation catalysts used in dynamic biogas upgrading. Bioresour
ison with results from a simple equilibrium model was consistent re- Technol 2015;178:319–22.
[12] Lecker B, Illi L, Lemmer A, Oechsner H. Biological hydrogen methanation – a re-
garding conditions in the reactor where equilibrium can be reached. view. Bioresour Technol 2017;245(Part A):1220–8.
Predictions from the rate-based model in regions which are kinetically [13] Bassani I, Kougias PG, Treu L, Porté H, Campanaro S, Angelidaki I. Optimization of
limited were also in accordance with experimental data within the hydrogen dispersion in thermophilic up-flow reactors for ex situ biogas upgrading.
Bioresour Technol 2017;234:310–9.

370
J. Witte, et al. Applied Energy 240 (2019) 359–371

[14] Aryal N, Kvist T, Ammam F, Pant D, Ottosen LDM. An overview of microbial biogas [25] Schweizerischer Verein des Gas- und Wasserfaches (SVGW), Richtlinie für die
enrichment. Bioresour Technol 2018;264:359–69. Einspeisung von erneuerbaren Gasen – G13/G18; 2016.
[15] Rachbauer L, Voitl G, Bochmann G, Fuchs W. Biological biogas upgrading capacity [26] Deutscher Verein des Gas und Wasserfaches, Technische Regel Arbeitsblatt G262;
of a hydrogenotrophic community in a trickle-bed reactor. Appl Energy 2011.
2016;180:483–90. [27] Calbry-Muzyka AS, Gantenbein A, Schneebeli J, Frei A, Schildhauer TJ, Biollaz
[16] Jürgensen L, Ehimen EA, Born J, Bo Holm-Nielsen J. Utilization of surplus elec- SMA. Deep removal of sulfur and trace organic compounds from biogas to protect a
tricity from wind power for dynamic biogas upgrading: Northern Germany case catalytic methanation reactor. Chem Eng J 2019;360:577–90.
study. Biomass Bioenergy 2014;66:126–32. [28] Kopyscinski J, Schildhauer TJ, Biollaz SMA. Methanation in a fluidized bed reactor
[17] Vo TTQ, Wall DM, Ring D, Rajendran K, Murphy JD. Techno-economic analysis of with high initial CO partial pressure: Part II – modeling and sensitivity study. Chem
biogas upgrading via amine scrubber, carbon capture and ex-situ methanation. Appl Eng Sci 2011;66:1612–21.
Energy 2018;212:1191–202. [29] Schildhauer TJ, Biollaz SMA. Fluidised bed methanation for SNG production –
[18] Uebbing J, Rihko-Struckmann LK, Sundmacher K. Exergetic assessment of CO2 process development at the Paul-Scherrer Institut. In: Schildhauer TJ, Biollaz SMA,
methanation processes for the chemical storage of renewable energies. Appl Energy editors. Synthetic natural gas from coal, dry biomass, and power-to-gas applica-
2019;233–234:271–82. tions. New York: Wiley & Sons; 2016. p. 221–9.
[19] Witte J, Settino J, Biollaz SMA, Schildhauer TJ. Direct catalytic methanation of [30] Seemann MC, Schildhauer TJ, Biollaz SMA. Fluidized bed methanation of wood-
biogas – Part I: new insights into biomethane production using rate-based modelling derived producer gas for the production of synthetic natural gas. Ind Eng Chem Res
and detailed process analysis. Energy Convers Manage 2018;171:750–68. 2010;49:7034–8. https://doi.org/10.1021/ie100510m.
[20] Witte J, Kunz A, Biollaz SMA, Schildhauer TJ. Direct catalytic methanation of [31] Schildhauer TJ. Methanation for SNG production – chemical reaction engineering
biogas – Part II: techno-economic process assessment and feasibility reflections. aspects. In: Schildhauer TJ, Biollaz SMA, editors. Methanation for SNG production –
Energy Convers Manage 2018;178:26–43. chemical reaction engineering aspects, in synthetic natural gas from coal, dry
[21] Castellani B, Rinaldi S, Bonamente E, Nicolini A, Rossi F, Cotana F. Carbon and biomass, and power-to-gas applications. New York: Wiley & Sons; 2016. p. 77–159.
energy footprint of the hydrate-based biogas upgrading process integrated with [32] Struis RPWJ, Schildhauer TJ, Czekaj I, Janousch M, Biollaz SMA, Ludwig C. Sulphur
CO2 valorization. Sci Total Environ 2018;615:404–11. poisoning of Ni catalysts in the SNG production from biomass: a TPO/XPS/XAS
[22] Castellani B, Morini E, Bonamente E, Rossi F. Experimental investigation and en- study. Appl Catal A Gen 2009;362:121–8. https://doi.org/10.1016/j.apcata.2009.
ergy considerations on hydrate-based biogas upgrading with CO2 valorization. 04.030.
Biomass Bioenergy 2017;105:364–72. [33] Zarfl J, Ferri D, Schildhauer TJ, Wambach J, Wokaun A. DRIFTS study of a com-
[23] Vo TTQ, Rajendran K, Murphy JD. Can power to methane systems be sustainable mercial Ni/γ-Al2O3 CO methanation catalyst. Appl Catal A Gen 2015;495:104–14.
and can they improve the carbon intensity of renewable methane when used to https://doi.org/10.1016/j.apcata.2015.02.005.
upgrade biogas produced from grass and slurry? Appl Energy 2018;228:1046–56. [34] Kopyscinski J, Schildhauer TJ, Vogel F, Biollaz SMA, Wokaun A. Applying spatially
[24] Collet P, Flottes E, Favre A, Raynal L, Pierre H, Capela S, et al. Techno-economic resolved concentration and temperature measurements in a catalytic plate reactor
and life cycle assessment of methane production via biogas upgrading and power to for the kinetic study of CO methanation. J Catal 2010;271:262–79. https://doi.org/
gas technology. Appl Energy 2017;192:282–95. 10.1016/j.jcat.2010.02.008.

371

Das könnte Ihnen auch gefallen