You are on page 1of 54

See discussions, stats, and author profiles for this publication at: https://www.researchgate.

net/publication/371500084

Green Hydrogen Production via Bioconversion and Water Electrolysis – Process


Modeling and Techno-Economic Assessment (TEA) using SuperPro Designer.

Preprint · June 2023


DOI: 10.13140/RG.2.2.21811.40487

CITATIONS READS

0 632

5 authors, including:

Yannis Stavropoulos Amir Mustafa


Intelligen Inc. Intelligen, Inc.
9 PUBLICATIONS 18 CITATIONS 7 PUBLICATIONS 0 CITATIONS

SEE PROFILE SEE PROFILE

Nikiforos Misailidis Ben Huffer

42 PUBLICATIONS 240 CITATIONS


Isomer Project Group
2 PUBLICATIONS 0 CITATIONS
SEE PROFILE
SEE PROFILE

Some of the authors of this publication are also working on these related projects:

Inorganic Materials View project

Biosurfactants Manufacturing - Process Simulation and Cost Analysis View project

All content following this page was uploaded by Demetri Petrides on 12 June 2023.

The user has requested enhancement of the downloaded file.


Green Hydrogen Production via
Bioconversion and Water Electrolysis
Process Modeling and Cost Analysis
using SuperPro Designer®
by
Yannis Stavropoulos, Amir Mustafa, Nikiforos Misailidis, Ben Huffer, and Demetri Petrides
June 2023

This is the ReadMe file of a SuperPro Designer example that analyzes the production of green hydrogen
via bioconversion pathways and water electrolysis. The bioconversion process utilizes 18 metric tons
(MT) per hour of wheat straw as feedstock which is converted into fermentable sugars via
thermochemical and enzymatic hydrolysis. Biohydrogen is produced using a combination of sequential
dark and photo fermentation and the product is purified using pressure swing adsorption. The process
generates 1.1 MT/hour of purified biohydrogen. The electrolytic process utilizes 10 MT/hour of municipal
water, which is converted into ultrapure water and then electrolyzed to produce 1.1 MT/hour of purified
hydrogen. This example also includes a model for hydrogen liquefaction. The analysis results for all three
processes include material and energy balances, equipment sizing, capital, and operating cost
estimation. The results indicate that significant government subsidies are necessary for the financial
viability of such investments considering the current technologies and market dynamics.

The flowsheets of the three processes are appended to the bottom of this document. You may test-drive
this model by downloading the functional evaluation edition of SuperPro Designer from the downloads
page of our website (www.intelligen.com). This example’s files can be found in the Examples \ Bio-Fuels
\ Hydrogen folder. You may download a ZIP file that includes all the example files of SuperPro Designer
from the following page of our website (top paragraph of the page).
https://www.intelligen.com/products/superpro-examples/ . If you have any questions regarding this
example or SuperPro Designer in general, please send an email message to dpetrides@intelligen.com

INTELLIGEN, INC.
Simulation, Design, and Scheduling Tools
For the Process Manufacturing Industries
www.intelligen.com
Introduction

The increasing disparity between the energy demands of the industrialized world and limited energy
sources has resulted in a significant increase in the use of fossil fuels. The increasing energy demand
worldwide has severely strained declining fossil fuels, which have increased pollution levels worldwide.
Burning fossil fuels results in ever-increasing levels of greenhouse gases (GHGs), which exacerbate the
dangers of global warming. Burning fossil fuels adds around 36.8 Giga tons of carbon to the atmosphere
each year in the form of carbon dioxide (Figure 1). This number has been increasing steadily, with a global
average annual increase of around 1.5% per year over the last decade [1]–[3]. Fossil fuel combustion is
also responsible for the emission of pollutants such as COx, NOx, SOx, CxHy, ash, particulate matter, and
other organic compounds [4], [5]. These pollutants are the main reason behind ozone layer depletion, global
warming, climate change, and acid rain [6], [7]. Over 86% of the world’s energy consumption and almost
100% of the energy consumed in the transportation sector is currently supplied by fossil fuels [8].

Figure 1: Global CO2 emissions from energy combustion and industrial processes, 1900-2022 [9].

The limited availability of global oil reserves and concerns about climate change from greenhouse gas
emissions has sparked a clear interest in developing clean and renewable energy alternatives to meet
growing energy needs [1], [8], [10]. Hydrogen gas (H2) is a promising alternative fuel that could significantly
reduce greenhouse gas emissions. Hydrogen produces water as a byproduct when combusted and is
therefore considered an environmentally benign and carbon-neutral energy source. With a high energy

P a g e 2 | 50
yield of 142 MJ/kg (three times that of gasoline), hydrogen is the ideal source of energy (Figure 2). Hydrogen
gas is an attractive future fuel due to potentially higher efficiencies in conversion to usable energy, low to
non-existent generation of pollutants, and high calorific value [1], [6], [11]–[15].

Figure 2: Calorific value of different fuel types, including hydrogen [14].

Around 38 million tons (5,000 petajoules) of hydrogen are produced annually globally, a market valued at
around US $60 billion. The demand for hydrogen is not limited to its use as an energy source. Hydrogen
gas is a widely used feedstock for the manufacture of chemicals, the hydrogenation of fats and oils in the
food industry, the manufacture of electronic equipment, steel processing, and for desulfurization and
gasoline reformulation in refineries. Demand for hydrogen is expected to grow exponentially in the near
future as hydrogen is needed to meet the demand for refining increasingly heavier, higher-sulfur crude oils
and oil sands and to meet tighter regulations on sulfur levels in gasoline and diesel fuel. Demand for
hydrogen is also expected to increase as it enters the transport sector as a fuel; about 40 million metric
tons (MT) of hydrogen per year would be needed to fuel around 100 million fuel cell-powered cars after full
market penetration [15], [16].

It has been reported that 50 million MT per annum of hydrogen is traded globally, with a current growth rate
of almost 10% annually. According to the US National Hydrogen Program, hydrogen will account for 8-10%

P a g e 3 | 50
of the total energy market by 2025. The US Department of Energy (US-DOE) reported that H2 power and
transport systems will be available in all regions of the United States by 2040. Due to the increasing demand
for hydrogen energy, the development of inexpensive and efficient hydrogen generation technologies has
received significant attention in recent years [16]–[19].

Although Hydrogen is the most common element on Earth, it does not exist in elemental form and is
produced from a variety of pathways and feedstocks. Almost 98% of industrial hydrogen is produced from
fossil fuels since this is the most cost-effective feedstock for hydrogen generation. Globally, 40% of
hydrogen is produced from natural gas or steam reforming of hydrocarbons, 30% from oil (consumed mainly
within refineries), 18% from coal, and the remaining 4% via water electrolysis [6], [8], [1]. From an energy
source perspective, hydrogen can be produced from biomass (agricultural waste, municipal waste, algae,
etc.), renewable energy (solar, wind, etc.), fossil fuels (coal, oil, natural gas, etc.), and nuclear energy
(Figure 3) [20].

Figure 3: Sources and Methods of Hydrogen Production [20].

An essential stage of the hydrogen value chain is the storage and transport of produced hydrogen gas.
Figure 4 illustrates the different pathways of hydrogen storage and transport. Different modes of hydrogen
storage and transportation may be selected depending on the intended end-use of hydrogen, distance, and
available infrastructure from both technical and cost perspectives. To achieve a viable energy density, the
produced hydrogen gas can be compressed or liquefied or converted to other molecules such as ammonia

P a g e 4 | 50
[21] or adsorbed onto solid carriers such as metal hydrides [22], [23]. Compressed hydrogen gas can be
transported via gas pipelines, whereas liquid hydrogen and hydrogen carriers can be transported by road,
rail, or sea [22].

Figure 4: Hydrogen storage and transportation pathways [22].

Liquid hydrogen is widely regarded as a potential key player in the global decarbonization effort as a clean
fuel source, particularly for fuel cell vehicles, due to its high purity, ease of gasification through simple heat
exchange with the environment, and higher density compared to compressed hydrogen. However, liquid
hydrogen also has many disadvantages that could impede its wider use. One major drawback is the high
cost of hydrogen liquefaction, which ranges between $2.5-3/kg. Another issue is the high specific energy
consumption (SEC) of hydrogen liquefaction, which is 12-15 kWh/kg, or approximately 30-40% of the
energy content of hydrogen. In contrast, less than 10% of the energy content of natural gas is consumed
in the production of liquefied natural gas (LNG). Additionally, liquid hydrogen has a low volumetric energy
density, which is only 40% of that of LNG. Due to the lower volumetric energy density and lower liquefaction
temperature of liquid hydrogen compared to LNG, the shipping cost per unit of energy of liquid hydrogen is
estimated to be 4-6 times higher than that of LNG. There are also other challenges associated with the
handling of liquid hydrogen, such as higher boil-off losses during storage and transportation compared to
LNG, lack of safety standards and regulations, technical difficulties in scaling-up of necessary equipment
(compressors, turbines, and cold boxes), and equipment damage due to hydrogen embrittlement [22],
[24]–[26].

This work focuses on the following green hydrogen production processes:

• Biohydrogen production through dark and photo fermentation.


• Electrolytic hydrogen production through water electrolysis.
• Hydrogen liquefaction using renewable energy.

These processes are described in detail in the following sections.


P a g e 5 | 50
Biohydrogen

Theoretical Background

Biohydrogen is defined as hydrogen produced biologically (usually by bacteria) from organic materials.
Biological generation of hydrogen technologies provides a wide range of approaches to generate hydrogen,
including direct bio-photolysis, indirect bio-photolysis, photo fermentation, dark fermentation, and by a
combination of the above. Figure 5 illustrates the various pathways of biohydrogen production.

Figure 5: Various routes of Biohydrogen Production [1].

In Summary, biohydrogen can be broadly referred to as hydrogen produced chemically, thermochemically,


biologically, biochemically, and bio-photolytically from any biomass [27]. The production of hydrogen from
biomass is associated with significant challenges. The hydrogen yield from biomass is low because the
hydrogen content in biomass is inherently low (about 6% vs. 25% for methane), and the energy content is
low due to biomass's 40% oxygen content [27].

Although biohydrogen production using bacteria and microalgae began about 35 years ago, most of these
biological means of hydrogen production are mainly from laboratory-scale experimentation. Dark and photo
fermentation of biomass or organic waste offer a route to biological hydrogen production that can be scaled
up to an industrial scale [27]. This work aims to analyze the feasibility of such a process, which is a
combination of dark and photo fermentation routes for hybrid mode hydrogen production.

P a g e 6 | 50
Dark Fermentation
Anaerobic spore-forming bacteria are used in dark fermentation, which takes place in the absence of
oxygen (anoxic) on substrates high in carbohydrates. The reaction below shows a typical dark fermentation
stoichiometry with glucose as the model substrate [27], [28].

C6H12O6 + 2H2O → 2CH3COOH + 4H2 + 2CO2

Figure 6: Flow diagram of the dark fermentation process [28].

In the dark fermentation process (Figure 6), a variety of waste materials with various chemical compositions
are employed as a substrate to create hydrogen. Agriculture wastes (such as rice, wheat and corn straw,
and animal manure), different types of wastewater (such as distillery wastewater, cheese whey effluent,
and palm oil mill effluent), food waste, municipal sewage waste, and sewage sludge are among those that
are most used. The waste substrates rich in carbohydrates tend to produce more hydrogen in comparison
to substrates rich in lipids or proteins.

Several microorganisms that can metabolize a wide variety of organic waste substrates are used in dark
fermentation to produce biohydrogen. These microorganisms are split into three groups based on their
preferred living temperatures: thermophiles (45–65 ⁰C), mesophiles (25–45 ⁰C), and psychrophiles (0 - 25
⁰C). While Thermoanaerobium is the most commonly reported thermophilic culture, Clostridium and
Enterobacter (Clostridium beijerinckii, Clostridium butyricum, Enterobacter aerogenes, and Enterobacter
asburiae) are the most commonly utilized mesophilic cultures for hydrogen generation
(Thermoanaerobacterium thermosaccharolyticum) [13], [29]–[31].

Hydrogenases are the most significant enzymes that control hydrogen metabolism and are also the most
preferred enzymes due to their high turnover rate and lower metabolic requirement. [FeFe] hydrogenase
and [NiFe] hydrogenase are the two fundamental hydrogenases that are phylogenetically distinct and have
different active sites. The following reversible reaction is catalyzed by these enzymes:

2H+ + 2e- ↔ H2

Compared to [NiFe] hydrogenases, which primarily catalyze the oxidation of molecular hydrogen, [FeFe]
hydrogenases are more active in the synthesis of molecular hydrogen. It should be noted that [FeFe]
hydrogenases are generally oxygen-sensitive [32], [33].

P a g e 7 | 50
Photo Fermentation
Photo fermentation is a biochemical process that is carried out in deficient nitrogen conditions using light
energy and organic acids. Gest and Kaman first identified biohydrogen production through photo
fermentation utilizing photosynthetic bacteria in 1949. Due to the presence of nitrogenase, some
photosynthetic bacteria from the family of Purple Non-Sulphur (PNS) bacteria (which are facultative
anoxygenic phototrophs belonging to the class of Alphaproteobacteria including several genera within
order., Rhodobium, Rhodobacter, Rhodospirillum, and Rhodopseudomonas) are capable of converting
organic acids (acetic, lactic and butyric) into hydrogen and carbon dioxide [28], [34], [35]. With acetic acid
as the starting organic acid, the conversion into hydrogen is shown below.

CH3COOH + 2H2O + light energy → 4H2 + 2CO2

During the photo fermentation biohydrogen production process, PNS bacteria can utilize both the reducing
power from the oxidation reaction of organic compounds and the light energy to reduce H + to H2.
Nitrogenase, whose primary function is to turn molecular nitrogen into ammonia so that the nitrogen source
can be fixed and used for cell growth, drives the generation of hydrogen. That is why photo fermentation
requires a nitrogen-deficient media [36].

Challenges related to hydrogen production via photo fermentation include the need for elaborate anaerobic
photo-bioreactors covering large areas and the limited availability of organic acids. Figure 7 illustrates a
schematic diagram depicting hydrogen production by photosynthetic bacteria.

Figure 7: Flow diagram of the photo fermentation process.

Hybrid systems can result in increased hydrogen production yields and a decrease in the need for light
energy. Both anaerobic and photosynthetic microorganisms can be found in such a setup. As depicted in
Figure 8, a variety of carbohydrates can be metabolized by anaerobic bacteria to produce hydrogen in the

P a g e 8 | 50
absence of light, and the resulting organic acids can serve as sources of hydrogen for photosynthetic
bacteria.

Figure 8: Flow diagram of the sequential dark and photo fermentation process.

Sequential dark/photo-fermentation is another name for the process that combines dark and light
fermentation. Such a two-stage process can be represented as follows:

Stage I: Dark fermentation

C6H12O6 + 2H2O → 2CH3COOH + 2CO2 + 4H2

Stage II: Photo fermentation

2CH3COOH + 4H2O → 8H2 + 4CO2

The theoretical hydrogen yield is 12 moles of hydrogen produced per mole of glucose, while in practice, the
maximum reported is 7.1 mol H2/mol glucose [12], [28], [37]–[40].

Process Description

A conceptual process for producing hydrogen using bioconversion was modeled and economically
evaluated in SuperPro Designer to estimate the expected raw material requirements, process equipment
capacities, utility requirements, capital investment, and production costs. The “BioHydrogen_v#.spf”
SuperPro file is a simplified example of a continuous biohydrogen production plant that utilizes wheat straw
as biomass feedstock and combines dark and photo fermentation in sequence. The development of the
model was based on data available in the technical and patent literature supported by our engineering
judgment and experience with related processes.

P a g e 9 | 50
For reporting and analysis purposes, the process flowsheet has been divided into four sections:

• Pretreatment and Hydrolysis (dark red icons)


• Dark Fermentation (black icons)
• Photo Fermentation (blue icons)
• Hydrogen Purification (green icons)

Flowsheet sections in SuperPro are simply sets of related unit procedures (processing steps). For
information on how to specify flowsheet sections and edit their properties, please use the Help facility, or
refer to Chapter 8.1 of the SuperPro manual provided in PDF format with the SuperPro installation. The
contents of each process section are described in greater detail next.

Pretreatment and Hydrolysis Section


Different first and second-generation renewable biomass feedstocks can be used as substrates for the
production of biohydrogen, such as agricultural waste, food industry waste, sewage sludge from wastewater
treatment plants, lignocellulosic waste, etc. [1], [11]. Additionally, as municipal and agricultural wastes can
be disposed of simultaneously, the production of biohydrogen from biomass and organic waste can provide
both the benefits of clean energy generation and waste management [20].

Biomass is an abundant renewable resource to produce hydrogen gas and occurs in various forms in
nature. However, in most such forms, biomass cannot be used directly as feedstock for fermentative
hydrogen production due to the presence of poorly biodegradable compounds and unsuitable nitrogen
levels. Hydrogen-producing microorganisms prefer fermentable sugar monomers (like pentoses and
hexoses) with low nitrogen in the substrate. Therefore, biomass is usually subjected to an appropriate pre-
treatment to provide the desired microbial environment [41]. This process utilizes wheat straw as the
biomass feedstock, whose composition is given in Table 1.

Table 1: Chemical composition of the wheat straw feedstock [42].

Component Mass (%)


Ash 7
Cellulose 34
Hemicellulose 26
Lignin 18
Solubles 10
Water 5

Pretreatment techniques are necessary to either remove or solubilize the lignin present in cellulosic
biomass to increase the digestibility of materials such as wheat straw. Additionally, pretreatment improves

P a g e 10 | 50
the biodegradability of cellulose-based biomass by decreasing the crystallinity of cellulose, increasing the
surface area that is accessible, and decreasing the amount of surface area that is used [20].

Procedure P-1 represents the transportation of wheat straw from fields to the plant. It was assumed that
wheat straw could be acquired from farmers for $30 per metric ton (MT), and the transportation cost is
$40/MT. The feedstock inlet flow rate is 18 MT/h, which translates to 432 MT/day. The wheat straw is then
sent to a shredder (P-4/SR-101) for size reduction through a hopper (P-3/HP-101). The shredded material
is stored in a silo (P-6/SL-102) via a belt conveyor (P-5/BC-101). The shredded material is then diluted with
water and acidified with 96% H2SO4 to achieve a final water concentration of 65% (in P-7/MX-101) and
H2SO4 concentration of 0.5% (in P-8/MX-102) [20], [43], [44]. The acidified slurry is stored in a blending
tank (P-9/V-101) and then sent for thermal hydrolysis to a plug flow reactor (P-15/PFR-101). The feed
stream of the thermal hydrolyzer is preheated to ~178 ⁰C 1 using high-pressure live steam (30 bar), which
is generated in P-14 (GBX-101). A custom mixing procedure (P-13 / MX-103) is used to set the desired
temperature of 178 ⁰C. It is essential to choose an appropriate temperature for thermal treatment. Low
temperatures may lead to insufficient disintegration, while too high temperatures can result in excessive
degradation of organic matter in biomass, thus reducing the value of biomass as an organic source for
fermentation. Too high treatment temperatures may also cause the formation of refractory compounds,
which are inhibitory for fermentation [44].

The heat treatment aims to convert the cellulose and hemicellulose present in wheat straw to sugar
monomers such as pentoses and hexoses. Apart from these sugars, some inhibitory materials such as
furfural and 5-HMF(5-(hydroxymethyl)furfural) are also formed through the Maillard reaction. The thermal
hydrolysis is carried out in a plug flow reactor (P-15/PFR-101) with a residence time of one hour. Table 2
displays the stoichiometry and assumed conversion of the thermal hydrolysis reactions.

Table 2: Thermal Hydrolysis Reactions.

Thermal Hydrolysis Conversion (%)


162 Hemicellulose + 18 Water → 2 Acetic Acid + 27 Hexoses + 151 Pentoses 80
Cellulose + Water → Hexoses 10
Pentoses → Furfural + 3 Water 1
Hexoses → 5-HMF + 3 Water 1

The hydrolyzed slurry (stream S-116), which is under high pressure (~12 bar) and temperature (~176 ⁰C),
is adiabatically flashed to 2 bar pressure (in P-16 / V-102). The vapors formed (stream S-118) are used to

1 For the Cellulose-based biomass, the treatment temperature ranges from 50 oC to 220 oC and duration
time from 90 s to 90 min. [38]
P a g e 11 | 50
preheat the incoming stream (in P-12 / HX-102) by introducing the hot vapors (120 ⁰C) in the shell side of
the heat exchanger. The temperature of the flashed stream is lowered to ~46 ⁰C by exchanging heat with
the incoming slurry (in P-11 / HX-101). The cooled slurry is then cooled down further to 35 ⁰C (in P-21/HX-
104), and then it enters a neutralizer (P-22/V-103) where the excess H2SO4 is neutralized using lime
(Ca(OH)2) according to the molar stoichiometry of Table 3.

Table 3: Neutralization Reaction.

Neutralization Reaction – Molar Stoichiometry Conversion

Ca(OH)2 + H2SO4 → CaSO4 + 2H2O 100%

The neutralized slurry is mixed with a recycled solution (P-23/V-104), which is analyzed later. A custom
mixer (P-24 / MX-105) manages the addition of 5% cellulase [44], and the slurry is fed to a set of 6 stirred
tank reactors (P-25 / R-101) operating in cyclical (batch) and staggered mode. The icon of the three small
blue lines on the lower-left corner of the procedure symbolizes the use of multiple equipment units operating
in staggered mode (alternating). The availability of staggered units is specified by right-clicking on the
procedure icon and selecting “Equipment Data”, then, checking the “Stagger Mode” box on the lower left
corner of the window and specifying the extra units. The icon of the clock next to the small blue lines
indicates the cyclical (batch) operation of each reactor. More specifically, each reactor receives material for
6 hours (the duration of its PULL-IN operation), hydrolyzes the received material for 24 hours (the duration
of its REACT operation), and then pumps material out for six hours (the duration of the TRANSFER-OUT
operation). The cycle duration of each reactor is 36 hours (6 + 24 + 6). The availability of 6 batch reactors
operating in staggered mode ensures the handling of a continuous inlet stream (S-128) and the generation
of a continuous output stream (S-129). Table 4 below displays the stoichiometry and assumed conversion
of the enzymatic hydrolysis reactions.

Table 4: Enzymatic Hydrolysis Reactions.

Lignocellulosic Enzymatic Hydrolysis Reactions Conversion


Cellulose Partial Hydrolysis (Mass Stoichiometry)
162 Cellulose + 18 Water → 180 Hexoses 90%
Hemicellulose Hydrolysis (Mass Stoichiometry)
162 Hemicellulose + 18 Water → 27 Hexoses + 153 Pentoses 70%
Cellulase Deactivation (Mass Stoichiometry)
1 Cellulase → 1 Solubles 100%

The resulting hydrolysate slurry contains concentrated sugars and unhydrolyzed solids such as lignin. The
slurry is first stored in a blending tank (P-26/V-108) and then pressed in a screw press (P-27/SP-102). The
resulting filtrate liquid syrup has about 75% water content and contains the fermentable sugars that are

P a g e 12 | 50
sent to the dark fermentation section. The filter cake (stream S-132) contains around 45% moisture and
significant amounts of fermentable sugars. For this reason, the filter cake is mixed with water (in P-19/MX-
104) which comes from the condensed vapors (P-17/HX-103) and combined with extra water supplied by
a flow adjusting mixer (P-18/FAD-101). It is then re-pressed using another screw press (P-20 / SP-101).
The recovered filtrate is recycled back to the P-23 / V-104 blending tank, and the cake is directed to a
wastewater treatment storage unit.

Dark Fermentation Section


The filtered hydrolysate from the pretreatment and hydrolysis section is mixed with an extra amount of water
(~14 MT/h) using a mixer (P-28/MX-106) to create a medium that is suitable for dark fermentation
(containing around 80% water). The diluted media is then mixed with Ammonium Bicarbonate (NH4HCO3)
as the source of ammonia for dark fermentation using a custom mixer (P-29/MX-107) to achieve a target
concentration 2 g/L of NH4HCO3 [45]. The mixed media is sterilized using heat (140 ⁰C) in a pasteurization
procedure (P-30/PZ-101), and the sterilized media is stored in a blending tank (P-31/V-106) which supplies
media to six fermentors operating continuously. The fermentors (P-32/FR-101) operate at 37 ⁰C with a
residence time of 24 hours. Table 5 below displays the stoichiometry and assumed conversion of the dark
fermentation reactions.

Table 5: Dark Fermentation Reactions.

Dark Fermentation Conversion


Ammonium Bicarbonate Dissociation (Molar Stoichiometry)
1 NH4HCO3 → 1 CO2 + 1 NH3 + 1 H2O 100%
Hexoses to Biomass (Mass Stoichiometry)
2 Ash + 100 Hexoses + 2 NH3 + 7 Solubles → 40 Biomass + 50 CO 2 + 21 H2O 4%
Pentoses to Biomass (Mass Stoichiometry)
2 Ash + 100 Pentoses + 2 NH3 + 7 Solubles → 40 Biomass + 50 CO 2 + 21 H2O 4%
Hexoses Decomposition (Molar Stoichiometry)
1 Hexoses + 2 H2O → 2 Acetic Acid + 2 CO2 + 4 H2 [28] 99%
Pentoses Decomposition (Molar Stoichiometry)
1 Pentoses + 1.67 H2O → 1.67 Acetic Acid + 1.67 CO 2 + 3.2 H2 [46] 99%

The hydrogen gas generated along with carbon dioxide and ammonia are vented. The vent stream (S-138),
which contains around 8% hydrogen gas, is sent to the hydrogen purification section which will be discussed
later in this document.

Several other organic acids besides acetic acid (e.g., butyric, propionic, lactic acid, and ethanol) are formed
during dark fermentation, which can be utilized as substrates for photo fermentation (next section). For the

P a g e 13 | 50
sake of modeling simplicity, we only considered the formation of acetic acid with elevated conversion than
what is reported in the literature [11], [47]. This is also reflected in the final concentration of acetic acid
(~113 g/L) in the outlet stream (S-139), which is equivalent to the total concentration of all organic acids as
reported in the literature [4], [16], [48].

Stream S-139 is then sent for centrifugation in a decanter centrifuge (P-34/DC-101) [33], [49] to remove the
biomass, 70% of which is then recycled back to the fermentors (stream S-141). The supernatant of the
centrifuge (stream S-143) containing acetic acid becomes the substrate of photo fermentation.

Photo Fermentation Section


The supernatant of the P-34/DC-101 centrifuge is then mixed with some inorganic salts and trace elements
to prepare the medium for photo fermentation using a mixture preparation procedure (P-36/MX-108) and
then stored in a blending tank (P-37/V-107) with a residence time of 4 hours after which it is sterilized (P-
38/PZ-102). The sterilized media is stored in another blending tank (P-39/V-108) which feeds an array of
ten parallel photo bioreactors (P-40/PBR-101) operating in continuous mode to prolong logarithmic growth
of photosynthetic bacteria and ensure maximum hydrogen production [50]. Table 6 below displays the
stoichiometry and assumed conversion of the photo fermentation reactions.

Table 6: Photo Fermentation Reactions.

Photo Fermentation Conversion


Biomass Formation (Mass Stoichiometry)
100 Acetic Acid + 2 Inorganic Salts + 1 Trace Elements → 50 CO2 + 20 PNS 50%
(Biomass) + 33 H2O
Acetic Acid Decomposition (Molar Stoichiometry)
2 Acetic Acid + 4 H2O → 4 CO2 + 8 H2 [28] 98%

Photo fermentation is operated at 30 ⁰C anaerobically utilizing Purple Non-Sulfur (PNS) bacteria like
Rhodobacter sphaeroides which are mesophiles and perform best in the temperature range from 25 to
35°C in the absence of oxygen and nitrogen deficient environment [33], [49], [51], [52].

The photo bioreactor is a tubular reactor made of transparent plastic material [53]. It is illuminated using
sunlight in day time and using an array of LED lights during nighttime with an illumination intensity of 10
kilolux (Klux) [54]. The power requirement for the nighttime illumination is calculated using the Rapid Tables
utility [55] based on the horizontal cross-sectional area of the photobioreactors (Length x Diameter)
calculated by SuperPro Designer. We estimated a power of 160 KW per reactor. The nighttime power
consumption is represented by a generic box (P-41/BGBX-101) where two hold operations are scheduled
for 12 hours each to represent the day and nighttime illumination schedule. The power requirement of 160
KW was specified on the Labor tab (as Auxiliary Utility) of the nighttime illumination operation.

P a g e 14 | 50
The outlet from the photo bioreactors (stream S-148) is then flashed adiabatically (P-42/V-109) to
atmospheric pressure. The vapor stream (S-149), containing hydrogen and carbon dioxide, is sent to the
hydrogen purification section.

A portion of the liquid stream of the flash procedure (stream S-152) containing the PNS (Purple Non-Sulfur)
bacteria is recycled back to the sterile media storage tank (P-39/V-108), while another portion (stream S-
151) is concentrated using a decanter centrifuge (P-53/DC-102). A portion of the concentrate of the
centrifuge (stream S-155) is also recycled back to the sterile media tank (P-39/V-108).

Hydrogen Purification Section


The hydrogen gas generated along with carbon dioxide from both the dark and photo fermentation sections
is first compressed to 5 bar pressure using centrifugal gas compressors (P-33 / VP-101) and (P-43 / VP-
102). The two compressed gas streams are combined (P-44/MX-109) and directed for purification using
pressure swing adsorption (PSA) in a series of PSA columns (P-45 / GAC-101) and (P-46 / GAC-102).
Since an explicit PSA procedure is not available in the current version of SuperPro Designer, Granular
Activated Carbon (GAC) Adsorption (for gaseous streams) columns were used to represent the PSA
columns. PSA generally involves 5-6 cycles of adsorption and desorption (namely pressurization, high-
pressure adsorption, blowdown, desorption at low pressure, pressure equalization, and rise) [56]. The
duration for all these cycles is accounted for in the breakthrough and regeneration time of the GAC
procedure and is set to be 400 and 200 seconds in both columns [57]. Both GAC columns operate in batch
(cyclical) mode with a cycle duration of 600 seconds (equal to the sum of breakthrough and regeneration
durations). A portion of the purified gas stream is used for the regeneration of the columns via (P-48 / FSP-
104) and (P-49 / FDIS-101). The CO2 streams of the two columns are combined (P-47 / MX-110) and
released into the atmosphere. Alternatively, the generated CO2 can be purified (not modeled) and sold to
beverage companies for making carbonated drinks. The plant generates 1.1 MT/h of purified hydrogen,
which translates to a daily production of 26.4 MT and annual production of around 9000 MT.

Material Requirements

Table 7 displays the raw material requirements in MT/year, kg/hour, and kg/kg of MP for this process (“MP”
stands for the main product, purified Hydrogen in this case). This table was extracted from the RTF version
of the Materials & Streams Report, which can be generated by selecting Reports Materials & Streams
from the main menu bar of SuperPro Designer. The format of the report (e.g., PDF, RTF, XLS, etc.) can be
specified through the dialog that is displayed when you select Reports Options. As expected, wheat
straw and water are the main materials utilized in this process.

P a g e 15 | 50
Table 7: Material Requirements for the Entire Process.

BULK MATERIALS (Entire Process)


Material MT/yr kg/h kg/kg MP
Acetic-Acid 54 6.5 0.01
Ammonium Bicarb 1,046 127.0 0.12
Ca(OH)2 1,527 185.5 0.17
Cellulase 453 55.1 0.05
Furfural 7 0.9 0.00
H2SO4 2,022 245.6 0.22
Inorganic Salts 2,515 305.5 0.28
Steam HP 38,842 4,718.4 4.28
TraceElem Solution 7,545 916.6 0.83
Water 453,663 55,109.7 49.99
Wheat Straw 148,176 18,000.0 16.33
TOTAL 655,850 79,670.8 72.26

Cost Analysis

SuperPro Designer performs thorough cost analysis, estimating capital (CAPEX) as well as operating
(OPEX) costs, and generates the following three pertinent reports (through the Reports menu): the
Economic Evaluation Report (EER), the Cash Flow Analysis Report (CFR), and the Itemized Cost Report
(ICR). Table 8 displays the Executive Summary of the Economic Evaluation Report. The Executive
Summary shows that the BioHydrogen plant analyzed in this example requires a total capital investment of
approximately $97 million, and its annual operating cost (AOC) is $61 million. This leads to a unit production
cost of $6.7/kg, which is very close to the cost of BioHydrogen reported in the literature [17], [58] but very
high in comparison to the traditional routes of hydrogen production from fossil fuels. However, according to
the US Department of Energy, with the advancement of technology, government incentives and regulations,
the unit cost of BioHydrogen is expected to drop to around $3.50/kg by 2025 [17], [59].

Assuming a selling price of $8/kg of hydrogen (considering subsidies from the government), it leads to a
gross margin of 16%, return on investment (ROI) of 18%, and a payback time of 5.5 years. These metrics
suggest that with sufficient government subsidies, such an investment would be financially feasible.

P a g e 16 | 50
Table 8: Executive Summary.

EXECUTIVE SUMMARY (2023 prices)


Total Capital Investment 96,526,000 $
Capital Investment Charged to This Project 96,526,000 $
Operating Cost 61,033,000 $/yr
Revenues 72,607,000 $/yr
Cost Basis Annual Rate 9,075,842 kg MP/yr
Unit Production Cost 6.72 $/kg MP
Unit Production Revenue 8.00 $/kg MP
Gross Margin 15.94 %
Return On Investment 18.08 %
Payback Time 5.53 years
IRR (After Taxes) 10.93 %
NPV (at 7.0% Interest) 25,098,000 $
MP = Total Flow of Stream 'Bio-Hydrogen'

Figure 9 displays a breakdown of the annual operating cost, which is also part of the EER. This type of
chart can be included in the report by selecting Reports Options and activating the Include Charts
option in the lower right corner of the dialog. The chart indicates that the cost of raw materials is the most
important, accounting for 35% of the overall operating cost. The facility-dependent cost is in the second
position, accounting for 31% of the overall operating cost. It is followed by the cost of labor (13%),
transportation (10%), and utilities (10%).

Figure 9: Annual Operating Cost Breakdown Chart.

P a g e 17 | 50
Electrolytic Hydrogen

Theoretical Background

Electrolysis is the splitting of a molecule into simpler components using energy from an electric current. In
the case of water electrolysis, the feedstock is water, which is split into hydrogen and oxygen. The following
equation describes the basic electrochemical reaction [60], [61]:

H2O ⇆ H2 + O2, ΔΗ°r = 285.83 kJ/mol

The reaction is endothermic, with the standard heat of reaction at 25°C being 285.83 kJ/mol when the
feedstock is liquid water, thus corresponding to the higher heating value of hydrogen. When hydrogen is
produced using low-carbon grid electricity or from renewable energy sources such as wind and solar, the
resulting hydrogen gas is commonly called “green” hydrogen. The hydrogen gas can be either stored as
compressed gas or liquefied and stored as liquid hydrogen. The oxygen produced is either released into
the atmosphere or stored to supply other industrial processes [60]–[62].

Water electrolysis is accomplished in a device called an electrolyzer. The “building block” of an electrolyzer
unit is the electrolytic cell, which consists of two electrodes – a cathode and an anode – connected by wires
to a source of direct current and a membrane that separates the cathode chamber from the anode chamber.
This is illustrated in Figure 10 [61], [63]–[66]:

Figure 10: Schematic illustration of a PEM electrolysis cell.

Two main types of electrolytic cells are currently used for industrial hydrogen production: the alkaline

P a g e 18 | 50
electrolysis cell, which uses a liquid alkaline solution (usually KOH) as the electrolyte, and the proton
exchange membrane (PEM) electrolysis cell, which uses a proton-conducting membrane (usually Nafion®)
that also serves as the electrolyte. From these two electrolytic cell technologies, the PEM electrolysis cell
is considered more suitable for hydrogen production, mainly due to its higher energy density and
compactness. Currently, PEM electrolysis cells have current densities of 1.5-2 A/cm2 and alkaline
electrolysis cells have current densities of 0.2-0.5 A/cm2. Another advantage of the PEM electrolysis cell is
that it produces hydrogen that is already compressed electrochemically to almost 50-60 bar, which means
that little-to-no mechanical compression may be necessary at a later stage for storage, transport, and other
uses (e.g., liquefaction). The main disadvantage of the PEM electrolysis cell design is the high cost of
materials used, which include precious metal catalysts such as iridium and platinum [60].

To achieve high production capacity, an electrolyzer is usually a modular system of several interconnected
electrolysis stacks and other peripheral systems (compressors, pumps, rectifiers, gas-liquid separators,
and demisters). A PEM electrolysis stack can have approximately 30-250 cells electrically connected in
series, with the power consumption ranging from less than a few hundred kW to more than 1 MW [60], [64]–
[68].

PEM electrolyzers require water of high purity, with ASTM Type II water being the minimum requirement
and Type I water being preferred for the process. Type II water is defined by a resistivity of >1 MΩ-cm and
<50 ppb of total organic carbon (TOC). Type I water is ultrapure water that has been purified to the highest
standards, with a resistivity of >18 MΩ-cm and the same maximum TOC content [69]. Therefore, if tap water
is used as feed, it must be treated to achieve ultrapure water quality before it can be used in a PEM
electrolyzer. Usually, a combination of different purification methods is used, including membrane
processes such as reverse osmosis (RO) and electrodialysis (ED), ion exchange processes such as
deionization (DI) and electrodeionization (EDI), and filtration processes such as carbon filtration and
ultrafiltration [70]. RO is a core process that is widely used for water purification since it can remove up to
99% of most impurities (ions, particles, colloids, organics, bacteria, and pyrogens). However, if the objective
is ultrapure water, it cannot stand alone. Generally, the production of ultrapure water from raw water is done
in three stages: 1) pretreatment stage, where carbon filtration is included to remove suspended solids, 2)
deionization stage, where RO is used to remove most impurities, and 3) polishing stage, where EDI can be
included to achieve ultrapure water quality [71].

Process Description

The “ElectrolyticHydrogen_v#.spf” SuperPro file represents a simplified model of a continuous plant that
produces hydrogen via water electrolysis. It processes 10,000 kg/h of tap water and produces 1,100 kg/h
or 26.4 metric tons (MT) per day of hydrogen. The process can be divided into three distinct stages: (1)
water purification, (2) electrolysis, and (3) hydrogen compression and storage. In the first stage, the city

P a g e 19 | 50
water is deionized and polished to produce ultrapure water. In the second stage, the ultrapure water is
electrolyzed to produce hydrogen. In the third stage, the produced hydrogen is compressed and stored.

For reporting and analysis purposes, the process flowsheet has been divided into three flowsheet sections:

• Water Purification (blue icons)


• Water Electrolysis (black icons)
• Hydrogen Compression & Storage (green icons)

Flowsheet sections in SuperPro are groups of related unit procedures (processing steps). They can be
used to facilitate reporting and analysis by specifying section-wide parameters associated with capital and
operating expenses. For information on how to specify flowsheet sections and edit their properties, please
consult the Help facility of the tool. The three flowsheet sections of this example are described in detail
below.

Water Purification
The process starts with tap water, assumed to have a total solids content of 300 ppm. Table 9 displays the
assumed composition of tap water.

Table 9: Assumed composition of tap water used as feed.

Pure Component Composition (%)


Anions 0.0150
Cations 0.0145
TOC 0.0005
Water 99.9700

The user-defined pure components “Cations”, “Anions”, and “TOC” represent the respective cations (e.g.,
Ca2+, Mg2+, Na+, K+), anions (e.g., Cl-, HCO3-, SO4-2, NO3-), and organic compounds and chemicals (e.g.,
chlorinated disinfection byproducts, free chlorine, hydrogen sulfide) contained in the tap water used.

First, the tap water (10 MT/h) is combined with the concentrate streams of the RO and EDI units (6.15 MT/h
in total) and fed to a Granular Activated Carbon (GAC) column (P-3 / GAG-101). This unit facilitates
dichlorination of the water, which is necessary to avoid damage to the RO membrane. The GAC column is
assumed to remove 95% of the TOC contained in the inlet stream.

Next, the water stream passes through ion exchange columns used for deionization to avoid scaling of the
RO membrane. First, it passes through a cation exchange column (P-4 / INX-101), which is assumed to
remove 95% of the cations (substituted with H+ ions from the resin), and then it passes through an anion
exchange column (P-5 / INX-102), which is assumed to remove 95% of the anions (substituted with OH -

P a g e 20 | 50
ions from the resin). To enable continuous operation, the equipment resource behind the cation exchange
procedure P-4 is set to represent two cation exchange columns that alternate between sorption and
desorption, as does the equipment resource behind the anion exchange procedure P-5. The cation resin is
regenerated using an aqueous solution containing 5.5% wt/wt HCl (prepared by mixing HCl 37% and
deionized water in appropriate quantities). The anion resin is regenerated using an aqueous solution
containing 5% wt/wt NaOH (prepared by mixing NaOH 50% and deionized water in appropriate quantities).

Next, a two-pass RO system (P-7 and P-8) is used to remove most of the remaining ions and organics
species to reach sufficiently low concentrations of ions and TOC. According to the ASTM D1193 standard
for water, ultrapure water of Type I should have a maximum TOC content of 50 ppm and a maximum
conductivity of 0.056 μS/cm [69]. Each RO unit is assumed to remove 99% of cations, anions, and TOC
and recover 95% of the water.

The RO system is followed by a membrane degasser (P-9/DG-101), which is used to remove dissolved
gases (e.g., CO2) as well as additional CO2 produced in the product stream of the cation exchange column.
It is produced by the reaction of the H+ ions that are transferred from the resin with existing CO3-2 and HCO3-
anions to produce carbonic acid, which is then decomposed into CO 2 and H2O.

Finally, an EDI unit (P-11 / ED-101) is used to reach the very low conductivity required. This unit is assumed
to remove 90% of the remaining ions and recover 90% of the water. The concentrate streams from the RO
and EDI units are recycled back to the GAC column. Note that due to the lack of an EDI unit procedure in
SuperPro, the electrodeionization process was simulated using an Electrodialysis (ED) unit procedure.
Considering that an electrodeionization process is basically an electrodialysis process done with the help
of ion exchange resins, this is considered sufficient for the purposes of this example.

In addition, three buffer tanks are used to handle temporary supply shortages: one before the GAC filter
(P-2 / V-101), another one before the RO filters (P-6 / V-102), and another one before the EDI unit (P-10 /
V-103). The residence time is 8 hours for the first storage tank and 4 hours for the other two tanks.

Finally, a “Material Storage Unit” (an implicit tank) called “Wastewater SU” is used to collect the contents of
all aqueous waste streams produced by the various units (GAC, ion exchange, RO). The main purpose of
“Material Storage Units” is to track the consumption or production of material at specific points in the
process. However, in this example, “Wastewater SU” is simply used to assign a common waste
treatment/disposal cost for all aqueous waste streams.

The above process produces 13,110 kg/h of ultrapure water having a combined solids concentration of
42 ppb. The solids concentration of ultrapure water is listed in Table 10.

P a g e 21 | 50
Table 10: Solids concentration of the produced ultrapure water.

Pure Component Concentration (ppb)


Anions 19
Cations 18
TOC 5

Water Electrolysis
In the water electrolysis section, ultrapure water is fed to a PEM electrolyzer (P-14 / EWC-101) operating
at 50 bar and 80°C. An Electrowinning unit procedure is used to simulate the PEM electrolyzer. To raise
the pressure of the feed stream to 50 bar, a pump unit procedure (P-13 / PM-101) is used. Additionally,
since the pressure of the hydrogen produced by the Electrowinning operation is always equal to the
atmospheric pressure, a compression unit procedure (P-15 / G-101) must be used in the simulation to
increase the pressure of the hydrogen product to the operating pressure. The capital and operating costs
of the added pump and compressor are assumed to be zero.

The PEM electrolyzer is assumed to split 75% of the water molecules into hydrogen and oxygen using
electricity. The following additional assumptions are made for the equipment design and process operating
parameters:

• Each cell stack consists of 255 cells.


• The maximum cell active area is 680 cm2.
• The current efficiency (or faradic efficiency) is 90% (i.e., 10% of the electricity is used to produce
parasitic products or heat).
• The current density is 1.7 A/cm2.
• The cell operating voltage is 1.7 V.

Based on the above assumptions, SuperPro calculates that to produce 1,100 kg/h of hydrogen, this
operation requires 111 stacks, with each stack having cells with an active area of 675.6 cm 2. The
electrolyzer also produces 8.7 MT/h of oxygen, which is released into the atmosphere. The total electricity
consumption of the electrolyzer system is 55.3 MW and therefore, the specific energy consumption of the
electrolyzer system is 50.2 kWh/kg of hydrogen. Since the LHV of hydrogen is 33.33 kWh/kg, the electrical
efficiency of the electrolyzer system is 66.4% on the LHV basis.

The unused water coming out of the electrolyzer is assumed to contain 100 ppm of metal ions. To simulate
this, a Custom Mixing unit procedure (P-16 / MX-102) is used to add a stream of “Anions” and “Cations” to
the water coming out of the PEM. A small portion (5%) of the contaminated water is purged (P-17 / FSP-
101), and the rest is circulated back to the GAC unit.

P a g e 22 | 50
Hydrogen Compression & Storage
The produced hydrogen is compressed to a pressure of 100 bar (which is adequate for both pipeline
transport and hydrogen liquefaction) using a mechanical compressor (P-18/G-102). The compressor’s
after-cooler cools down the hydrogen to a temperature of 35°C. The compressed and cooled hydrogen is
stored in appropriate hydrogen storage tanks (P-19/V-105). Assuming a storage period of 7 hours, the
hydrogen mass that must be stored is 7.7 MT. The density of hydrogen at 100 bar and 35°C is approximately
7.44 g/L, so the volume of hydrogen that must be stored is approximately 1,035 m 3. To store this amount,
six hydrogen storage tanks of design pressure 100 bar and net volume 200 m3 are used. The net length
and diameter of these tanks are approximately 27.3 m and 3.03 m, respectively.

Cost Analysis

The cost of producing green hydrogen through water electrolysis was analyzed using SuperPro Designer.
The software allows for the estimation of capital costs (CAPEX), operating costs (OPEX), revenues,
economic viability indicators, etc., based on a variety of inputs such as costs of raw materials, utilities, labor,
disposal/treatment costs of waste streams, and selling prices of product streams. Additionally, the software
enables users to estimate equipment purchase costs using built-in or user-defined models. Other capital
expenses, such as installation, piping, engineering, construction, startup, etc., and other operating
expenses, such as maintenance, depreciation, R&D, etc., can be estimated for each flowsheet section
using suitable section-specific cost factors and parameters. Finally, several project-wide assumptions, such
as project lifetime, depreciation, income taxes, etc., must be made.

For this analysis, several assumptions were made regarding capital and operating expenses. The most
significant ones are listed below:

- The project lifetime is 20 years.


- The depreciation period is 15 years.
- The annual operating time (AOT) is 5000 full-load hours (approximately 208 days or 0.57 years),
corresponding to a capacity factor of approximately 57%. The term “full-load hours” refers to the total
number of hours in a year that the plant would need to operate at nominal capacity to consume the
same amount of electricity as the actual operation, which includes both downtime and partial-load
operation. The capacity factor is the ratio of full-load hours to the maximum number of operating hours
in a year (8760). Electrolyzer systems are expected to operate for 4000-6000 hours due to the limited
supply of renewable electricity and to avoid expensive operating hours [72]. Supplementing with
nuclear power (i.e., Pink Hydrogen) has the ability to increase the capacity factor significantly, which
would allow more H2 production and increase ROIs [73].

P a g e 23 | 50
- The uninstalled unit purchase cost of a PEM electrolyzer stack is $700/kW. This was verified from
budgetary equipment quotations, and IRENA reports [74] that demonstrate an average electrolyzer
investment of $770/kW, which will continue to become more economical as more manufacturers enter
the market. In the US DOE Program Record “Cost of Electrolytic Hydrogen Production with Existing
Technology-2020” the range of uninstalled full system equipment costs for 2020 (including balance of
plant equipment) is $1000-$1500/kW [75]. The following user-defined equipment purchase cost model
is used to calculate the purchase cost (C) of a PEM stack for a given total active area (A): C = C 0 *
(A/A0)^a, where A0 is the reference active area, C0 is the reference cost and a is the exponent of the
power law model. For a reference stack that contains 255 cells of area 680 cm 2, the reference active
area is 173,400 cm2. For typical values of current density (1.7 A/cm2) and cell operating voltage (1.7
V), the reference power consumption of the equipment unit (stack) is 0.5 MW. Therefore, the reference
cost is $350,000. The exponent “a” is assumed to be equal to one. Assuming that the cell area can be
500-1500 cm2 and the number of cells per stack can be 30-350, the above cost model is assumed valid
for capacities between 15,000 cm2 (low end) and 525,000 cm2 (high end).
- The equipment cost of each 100 bar and 200 m 3 large metal hydrogen storage tank from vendor
budgetary quotations is approximately $1.5 to 2.5 million ($2 million was assumed). Some savings may
exist to store at higher pressures (200 bar+). ASME Section VIII (Pressure Vessel Code) Div. 2 & 3
construction allows higher storage pressure based on engineering calculation. Hydrogen storage
vessels of various material construction (metal with wrap, metal/plastic liner, etc.) will continue to
become more economical.
- The direct CAPEX is approximately 4 times the total equipment purchase cost. Indirect costs
(engineering and construction) are 40% of direct costs. The contractor’s fee is 5% of direct and indirect
costs. The contingency cost is set at 30% of direct and indirect costs. These values were estimated
based on CAPEX breakdown estimates by KPMG [72].
- The plant consumes renewable electricity generated from solar or wind farms, which is purchased at
an average annual price of $50/MWh [72].
- The purchasing price of tap water is $1/MT.
- The treatment/disposal cost of aqueous waste (from GAC, ion exchange, and RO filters) is $1/MT.
- The selling price of produced hydrogen is $8.60/kg. This value is explained later.

SuperPro generates three reports related to process economics: the Economic Evaluation Report (EER),
the Cash Flow Analysis Report (CFR), and the Itemized Cost Report (ICR). Table 11 shows the Executive
Summary taken from the EER for the electrolytic hydrogen example. It displays (among others) the total
capital investment (TCI), total operating cost, and total unit operating cost for a 55 MW electrolytic hydrogen
plant that produces 26.4 MT/day of hydrogen, including water purification and hydrogen compression and
storage. According to the above table, the project’s TCI is approximately $188 million ($3400/kW of PEM
electrolyzer), with a total operating cost of around $35.9 million and a total unit operating cost of $6.52/kg
(including hydrogen compression and storage). The capital and operating costs for each individual

P a g e 24 | 50
flowsheet section will be presented later. To achieve a positive NPV at a 7% discount rate and a gross
margin of at least 20%, the selling price of hydrogen must be at least $8.70/kg. At this selling price, the
gross margin would be around 25%, the ROI around 11%, and the payback time approximately 9 years.

Table 11: Executive summary of the electrolytic hydrogen example model.

EXECUTIVE SUMMARY (2023 prices)

Total Capital Investment 188,023,000 $


Capital Investment Charged to This Project 188,023,000 $
Operating Cost 35,857,000 $/yr
Revenues 47,867,000 $/yr
Cost Basis Annual Rate 5,501,952 kg MP/yr
Unit Production Cost 6.52 $/kg MP
Net Unit Production Cost 6.52 $/kg MP
Unit Production Revenue 8.70 $/kg MP
Gross Margin 25.09 %
Return On Investment 10.80 %
Payback Time 9.26 years
IRR (After Taxes) 7.17 %
NPV (at 7.0% Interest) 2,325,000 $
MP = Total Flow of Stream 'Hydrogen'

Table 12 displays a breakdown of the TCI per flowsheet section extracted from the ICR. As shown in Table
12, the individual capital expenses (CAPEX) for water purification, water electrolysis, and hydrogen
compression and storage are approximately $11.7 million, $132 million, and $44 million, respectively. By
summing up the CAPEX for water purification and water electrolysis, we find that the CAPEX for hydrogen
production (without compression and storage) is around $143.7 million. This corresponds to $2,600/kW of
PEM electricity consumption and $5,450/(kg H2/day).

Table 12: TCI section breakdown for the electrolytic hydrogen example model.

CAPITAL INVESTMENT PER PROCESS SECTION (in $) (2023 prices)


Direct
Equipment Start-up Total
Fixed Working Up Front Up Front
Section Purchase and Capital
Capital Capital R&D Royalties
Cost (PC) Validation Investment
(DFC)
Water Purification 3,034,444 10,998,645 120,644 549,932 0 0 11,669,222
Water Electrolysis 40,661,053 125,264,505 557,215 6,263,225 0 0 132,084,945
Hydrogen Compression &
13,483,158 42,149,169 12,043 2,107,458 0 0 44,268,670
Storage
Total 57,178,655 178,412,318 689,902 8,920,616 0 0 188,022,836

To gain a better understanding of the TCI, let’s examine Table 13, which presents a segment of the
Profitability Analysis table of the EER that outlines the breakdown of the TCI. The Direct Fixed Capital
(DFC) for this project is approximately $178.4 million, while startup cost and working capital amount to $8.9
million and $0.7 million, respectively.

P a g e 25 | 50
Table 13: TCI breakdown for the electrolytic hydrogen example model.

PROFITABILITY ANALYSIS (2023 prices)


A. Direct Fixed Capital 178,412,000 $
B. Working Capital 690,000 $
C. Startup Cost 8,921,000 $
D. Up-Front R&D 0$
E. Up-Front Royalties 0$
F. Total Investment (A+B+C+D+E) 188,023,000 $
G. Investment Charged to This Project 188,023,000 $

Table 14: DFC breakdown for the electrolytic hydrogen example model.

FIXED CAPITAL ESTIMATE SUMMARY (2023 prices in $)


3A. Total Plant Direct Cost (TPDC) (physical cost)
1. Equipment Purchase Cost 57,179,000
2. Installation 6,611,000
3. Process Piping 5,718,000
4. Instrumentation 6,021,000
5. Insulation 1,715,000
6. Electrical 5,718,000
7. Buildings 5,718,000
8. Yard Improvement 2,859,000
9. Auxiliary Facilities 2,859,000
TPDC 94,398,000

3B. Total Plant Indirect Cost (TPIC)


10. Engineering 14,160,000
11. Construction 23,600,000
TPIC 37,759,000

3C. Total Plant Cost (TPC = TPDC+TPIC)


TPC 132,157,000

3D. Contractor's Fee & Contingency (CFC)


12. Contractor's Fee 6,608,000
13. Contingency 39,647,000
CFC = 12+13 46,255,000

3E. Direct Fixed Capital Cost (DFC = TPC+CFC)


DFC 178,412,000

Table 14, which is also extracted from the EER, provides a detailed breakdown of the DFC. The DFC

P a g e 26 | 50
consists of direct and indirect costs. Direct costs include tangible items such as equipment purchase and
installation, piping, instrumentation and controls, insulation, electrical equipment, buildings, yard
improvement, and auxiliary facilities. Indirect costs include intangible items such as engineering and
supervision, construction, contractor’s fee, and contingency. SuperPro divides indirect costs into two
groups: the first group estimates indirect costs, such as engineering and construction, based on multipliers
of the sum of direct costs, while the second group estimates indirect costs, such as contractor’s fee and
contingency, based on multipliers of the sum of direct costs and the first group of indirect costs. Direct costs
can be estimated based on multipliers of installed equipment costs. For this analysis, the DFC is
approximately four times the installed equipment purchase cost. According to Table 14, the direct CAPEX
for this project is approximately $94.4 million. If the “Hydrogen Compression and Storage” section is
excluded, the direct CAPEX for hydrogen production (without compression and storage) is approximately
$72.1 million ($1300/kW of PEM electrolyzer power). Table 15, also taken from the EER, displays
equipment sizes and purchase cost.

Table 15: Equipment purchase costs for the electrolytic hydrogen example model.

EQUIPMENT SPECIFICATION AND FOB COST (2023 prices)


Main Equipment
Quantity/
Standby/ Name Description Unit Cost ($) Cost ($)
Staggered
6/0/0 V-105 Horizontal Tank 2,000,000 12,000,000
Vessel Volume = 200.00 m3
1/0/0 G-102 Centrifugal Compressor 809,000 809,000
Compressor Power = 487.36 kW
111 / 0 / 0 EWC-101 Electrowinning Cell 348,000 38,628,000
Area = 17.23 cm2
3/0/0 V-101 Receiver Tank 286,000 858,000
Vessel Volume = 48.32 m3
2/0/0 V-102 Receiver Tank 250,000 500,000
Vessel Volume = 36.23 m3
2/0/0 V-103 Receiver Tank 238,000 476,000
Vessel Volume = 32.70 m3
2/0/0 V-104 Receiver Tank 227,000 454,000
Vessel Volume = 29.43 m3
1/0/0 GAC-101 GAC Adsorber (for Liquid Streams) 110,000 110,000
Column Volume = 1100.12 L
1/0/0 RO-101 Reverse Osmosis Filter 52,000 52,000
Membrane Area = 211.23 m2
1/0/0 RO-102 Reverse Osmosis Filter 50,000 50,000
Membrane Area = 200.67 m2
1/0/1 INX-102 Ion Exchanger 41,000 82,000
Column Volume = 72.45 L
1/0/1 INX-101 Ion Exchanger 41,000 82,000
Column Volume = 70.45 L
1/0/0 ED-101 Electrodialyzer 38,000 38,000

P a g e 27 | 50
Membrane Area = 0.20 m2
1/0/0 DG-101 Degasifier 29,000 29,000
Degasifier Volume = 0.00 L
Unlisted Equipment 3,011,000

TOTAL 57,179,000

According to the above table, the PEM electrolyzer stack system costs around $38.6 million, and the
hydrogen storage tanks cost another $12 million. Together, these two capital expenses represent almost
90% of the total uninstalled equipment purchase cost.

To get a better understanding of the total operating costs as well, we can refer to Table 16 (a) and Table
16(b) taken from the ICR.

Table 16: Operating cost breakdown for electrolytic hydrogen: (a) per item; (b) per section.

COST PER PROCESS SECTION (2023 prices)


SUMMARY PER COST ITEM (Entire Process)
Cost Item $/kg MP $/day $/year %
Raw Materials 0.03 824 171,660 0.48
Facility 3.54 93,401 19,458,461 54.27
Labor 0.22 5,762 1,200,513 3.35
Consumables 0.05 1,303 271,425 0.76
Lab/QC/QA 0.02 604 125,863 0.35
Utilities 2.66 70,184 14,621,760 40.78
Waste Trtmt/Disp 0.00 34 7,069 0.02
Transportation 0.00 0 0 0.00
Miscellaneous 0.00 0 0 0.00
TOTAL 6.52 172,112 35,856,752 100.00

SUMMARY PER SECTION


Section $/kg MP $/day $/year %
Water Purification 0.41 10,773 2,244,312 6.26
Water Electrolysis 5.13 135,354 28,198,660 78.64
Hydrogen Compression &
0.98 25,986 5,413,780 15.10
Storage
TOTAL 6.52 172,112 35,856,752 100.00

Table 16(b) shows that the costs of water purification, water electrolysis, and hydrogen compression and
storage are $0.41/kg, $5.13/kg, and $0.98/kg, respectively. The cost of hydrogen production (without
compression and storage) is the sum of water purification and water electrolysis costs. Based on the
rounded costs in the table, the hydrogen production cost is $5.54/kg. However, based on the raw,
unrounded costs, the actual hydrogen production cost is $5.53/kg.

P a g e 28 | 50
Table 16(a) reveals that the utility and facility-dependent costs make up approximately 41% and 54% of the
total operating cost, respectively. The facility-dependent costs include maintenance, depreciation, and other
overhead costs.

To further examine the contribution of utility costs to the total operating cost, Table 17, also sourced from
the ICR, is shown below.

Table 17: Utility costs for the electrolytic hydrogen example model.

UTILITIES COST - SECTION SUMMARY


Electricity
Amount Cost
Section Name %
(kW-h/year) ($/year)
Water Purification 47,828 2,391.40 0.02
Water Electrolysis 301,382,483 13,973,372.84 99.11
Hydrogen Compression &
2,461,396 123,069.79 0.87
Storage
TOTAL 303,891,707 14,098,834.04 100.00

Heat Transfer Agents


Amount Cost
Section Name %
(kg/year) ($/year)
Water Purification 6,647,913 640.14 0.12
Water Electrolysis 12,772,528,751 498,148.19 95.26
Hydrogen Compression &
482,762,105 24,138.11 4.62
Storage
TOTAL 13,261,938,770 522,926.43 100.00

The above table indicates that the cost of electricity consumed by the PEM Electrolyzer in the “Water
Electrolysis” section represents approximately 96% of the total utility costs.

To summarize, the estimated CAPEX and OPEX for hydrogen production, compression, and storage are
approximately $3400/kW and $6.52/kg. Meanwhile, the total CAPEX, direct CAPEX, and OPEX for
hydrogen production alone (without compression and storage) are around $2600/kW, $1300/kW, and
$5.53/kg H2. The production cost depends primarily on the PEM electrolyzer and renewable electricity
costs, as highlighted in the CAPEX and OPEX breakdowns presented earlier.

The CAPEX and OPEX estimates for hydrogen production presented in this analysis align with those
reported in the literature. For instance, the U.S. DOE Program Record “Cost of Electrolytic Hydrogen
Production with Existing Technology-2020” indicates that the cost of hydrogen production can range from
$4 to $6/kg of H2, assuming an electrolyzer capital cost of $1000/kW and electricity from solar or wind farms
[75]. The second edition of the “Green Hydrogen Guidebook” reports a higher levelized cost of hydrogen
(LCOH) of about $7-8/kg H2 (at the upper range) from 2018 IEA sources [61]. The LCOH is calculated as
the total CAPEX and OPEX divided by the total hydrogen production during the plant’s lifetime.
Furthermore, a KPMG analysis performed in 2022 estimates the total CAPEX, direct CAPEX, and LCOH

P a g e 29 | 50
for a 50 MW PEM electrolyzer operating for 5000 full-load hours to be $2400-3500/kW (2200-3200€/kW),
$1100-1650/kW (1000-1500€/kW), and $5.30/kg Η2 (4.80€/kg H2), respectively [72].

The unsubsidized hydrogen production cost calculated in this analysis is more than three times higher than
the average hydrogen production cost from fossil fuels, which ranges from $1 to $1.5/kg [60]. This means
that currently, green hydrogen production through water electrolysis is not economical without major
government incentives. In the U.S., the Inflation Reduction Act (IRA) provides tax credits for renewable
energy and green hydrogen production plants. These plants can receive production tax credits of
$0.026/kWh for renewable energy production and up to $3/kg for green hydrogen production for up to 10
years of operation from 2023-2032 [76]. If a project includes a solar or wind farm, it could receive production
tax credits of up to $3/kg for green hydrogen production and additional tax credits of around $1.40/kg for
producing renewable energy, assuming that the generated energy is 50 kWh/kg of H 2 (the same as the
specific energy consumption of the PEM electrolyzer). Therefore, with government subsidies like the US
IRA program, the electrolytic hydrogen production cost could be reduced to around $1.10/kg, making the
project potentially economically viable.

The high production cost of green electrolytic hydrogen is mainly due to the cost of PEM electrolyzer and
renewable electricity, as shown earlier. These costs are expected to decrease over time, thus reducing the
production cost. The production cost could also be further reduced if the oxygen produced by the PEM
electrolyzer could be sold to potential end users such as the medical field or a nearby plant that consumes
a lot of oxygen (e.g., steel plant). In addition, the heat of the cooling water leaving the PEM electrolyzers
(which is more than 11 MW) could be recovered as hot water and supplied to a district heating network or
nearby plants.

One useful exercise is to perform sensitivity analyses with SuperPro to determine the impact of changes in
model parameters on the simulation results. This allows for a better understanding of the potential risks and
rewards of a project under different sets of assumptions. Different scenarios can be evaluated individually
by manually changing parameter values and re-running the simulation to see the results. Alternatively,
these simulation runs can be automated through MS Excel. Mathematical optimization and Monte Carlo
simulation can also be performed in a similar manner.

For instance, one could perform a sensitivity analysis to assess the impact of important parameters affecting
the hydrogen production cost, such as the production capacity, full-load hours, PEM electrolyzer cost, and
electricity cost. Table 18 presents the simulation results for a few indicative scenarios.

P a g e 30 | 50
Table 18: Hydrogen production costs for different scenarios.

Production Full-load Electrolyzer Electricity Production

Scenario Capacity hours Cost Cost Cost

(MT/day) (h) ($/kW) ($/MWh) ($/kg)

1 26.4 5000 700 50 5.53

2 52.8 5000 700 50 5.38

3 26.4 7000 700 50 4.78

4 26.4 5000 350 50 4.36

5 26.4 5000 700 30 4.52

6 26.4 5000 350 30 3.34

7 26.4 5000 175 10 1.74

These scenarios are explained below:

1. The first scenario is the baseline scenario presented earlier, with a production capacity of 26.4
MT/day, a capacity factor of 57%, a PEM electrolyzer cost of $700/kW, an electricity cost of
$50/MWh, and a production cost of $5.54/kg.
2. In the second scenario, the production capacity is doubled. This causes the production cost to drop
slightly to $5.38/kg.
3. The third scenario is the same as the baseline scenario except that the full-load hours are increased
to 7000 hours. For this scenario, the production cost is lowered to $4.78/kg.
4. In the fourth scenario, which is the same as the baseline scenario except that the electrolyzer cost
is halved, the production cost becomes $4.36/kg.
5. The fifth scenario is the same as the baseline scenario except that the electricity cost is reduced to
$30/MWh. For this scenario, the production cost is $4.52/kg.
6. The sixth scenario is the same as the previous scenario except that the electrolyzer cost is halved.
This causes the production cost to drop to $3.34/kg.
7. Finally, the seventh scenario is the same as the sixth scenario except that the electrolyzer cost is
halved again, and the electricity cost is reduced to one-third of its previous value. For this scenario,
the production cost is $1.74/kg.

The last scenario indicates that if the electrolyzer cost drops to $175/kW and the renewable electricity can
be provided at $10/MWh, then, green hydrogen production through water electrolysis can become

P a g e 31 | 50
competitive with hydrogen produced from fossil fuels without the need for government subsidies.

For information on how to drive SuperPro Designer through MS Excel and automate sensitivity analysis,
please refer to the examples in the COM folder (C:\Users\Public\Public Documents\ Intelligen\SuperPro
Designer\ v# \ Process Library \ Examples \ COM). Additional information can be found in the Help Facility
by clicking on the “COM Library” menu item.

Hydrogen Liquefaction

Theoretical Background

A block diagram of a hydrogen liquefaction process based on a simple Claude liquefaction cycle is shown
in Figure 11. First, the hydrogen feed is compressed to about 80 bar using reciprocating compressors and
pre-cooled to around 80 K using liquid nitrogen. Then, a series of cryogenic heat exchangers are used to
cool down the hydrogen to a temperature of around 30 K. Finally, the cold hydrogen is expanded to
atmospheric pressure using a Joule-Thomson (J-T) valve, and the flashed vapor is used as a self-refrigerant
in the heat exchangers before it is recycled to the feed. To improve the performance of the self-refrigeration
process, a turbo-expander is used to cool part of the feed leaving the first heat exchanger and to increase
the refrigerant flow of the second heat exchanger. The cryogenic heat exchangers contain catalysts that
convert ortho-hydrogen to para-hydrogen to reduce boil-off during storage. In addition, adsorbers may be
added in several stages of the process to remove gaseous contaminants that may freeze in the system
[22].

Several modifications of the Claude cycle are used in industry to reduce SEC and improve efficiency. These
include the use of a chiller after compression, closed-loop nitrogen re-liquefaction, optimized heat transfer
using a higher number of heat exchangers and turbo-expanders, cascade helium Joule-Brayton cycle
coupled with Claude cycle, mixed refrigerants instead of hydrogen self-refrigerant, etc. [22], [77]–[82].

P a g e 32 | 50
Figure 11: Hydrogen liquefaction based on a simple Claude cycle [22], [77].

Process simulation
The simulation of hydrogen liquefaction processes presents many thermodynamic challenges. First,
hydrogen behaves as a real gas under liquefaction conditions, which means that if hydrogen is treated as
an ideal gas, the energy balance and equipment design calculations will be inaccurate, and it is also
impossible to simulate the cooling effect of Joule-Thomson expansion. Second, normal hydrogen (n-H2)
consists of 75% ortho-hydrogen (o-H2) and 25% para-hydrogen (p-H2). At the temperature range of
liquefaction, the specific heat capacities of the two nuclear spin isomers of hydrogen (o-H2 and p-H2) are
quite different. Moreover, the ortho- para- ratio is temperature dependent, and o-H2 is almost entirely
converted into p-H2 in the cryogenic heat exchangers. Consequently, the cold hydrogen leaving the
cryogenic heat exchangers can be treated as pure p-H2. Third, mixed refrigerants may be used for
liquefaction. To calculate the enthalpy change, volume, and vapor-liquid equilibrium (VLE) status of
hydrogen and refrigerants during liquefaction, process simulation studies often use a variety of EOS with
varying degrees of simplicity, computational efficiency, and accuracy. Leachman et al. developed a complex
reference equation of state (EOS) for normal, ortho- and para- hydrogen, which is extremely accurate but
can also be impractically slow. Consequently, hydrogen is often treated as pure (normal) hydrogen, using
suitable modifications of simpler cubic equations, such as the quantum-corrected Peng-Robinson EOS

P a g e 33 | 50
developed by Aasen et al. Suitable binary mixture EOS have also been developed for mixtures with
hydrogen and refrigerants [22], [77].

SuperPro Designer includes cubic EOS for VLE calculations but not for enthalpy change calculations. So,
hydrogen is treated as an ideal gas, with accuracy being traded off for simplicity. Therefore, it may only be
suitable for preliminary simulation studies of hydrogen liquefaction cycles where simulation errors of 20-
30% are acceptable. Additionally, the ideal gas specific heat capacity differences of n-H2, o-H2, and p-H2
are neglected in this analysis. However, the heat released from the ortho- para- conversion reaction is
considered.

Process Description

The “HydrogenLiquefaction_v#.spf” SuperPro file is a simplified model of a continuous hydrogen


liquefaction plant that converts gaseous hydrogen into liquid hydrogen. The plant processes 1100 kg/h
(26.4 MT/day) of gaseous hydrogen to produce an equal amount of liquid hydrogen. The process flowsheet
has been divided into three sections:

• Compression and precooling (black icons)


• Cryogenic cooling (red icons)
• Liquefaction (blue icons)

Each section is described in detail below.

Compression and Precooling


The hydrogen gas feed enters the process at a pressure of 40 bar and a temperature of 35°C. It is then
mixed with hydrogen reflux, which is also compressed to 40 bar in two stages (P-1/G-101 and P-2/G-102).
The compressed gas is subsequently pre-cooled to 82 K (P-4/HX-101) using an implicit closed-loop
nitrogen re-liquefaction and hydrogen precooling cycle [78].

Although the cycle has not been added explicitly to the flowsheet, the nitrogen re-liquefaction and hydrogen
precooling sections of the cycle were simulated as an open loop in a separate SuperPro file
(NitrogenReliquefaction_v#.spf) which is part of the files of this example to determine the enthalpy change
of the refrigerant in the precooling section and the return temperature of the refrigerant from the precooling
section to the re-liquefaction section. In the nitrogen re-liquefaction section, gaseous nitrogen is
compressed to 200 bar and cooled to 280 K. It is then expanded to 1 atm and 77 K and separated in a
gas/liquid separator into two separate phases - liquid nitrogen and cold gaseous nitrogen. Subsequently,
in the hydrogen precooling section, hydrogen is cooled down in two heat exchangers in series, one for
gaseous nitrogen and another for liquid nitrogen. In the second heat exchanger, the liquid nitrogen is
evaporated, and the vapor is mixed with the existing cold gas and fed to the first heat exchanger. The

P a g e 34 | 50
gaseous nitrogen exits the first heat exchanger at 270 K and returns to the re-liquefaction section. The
enthalpy change of the refrigerant in the precooling section (from liquid and gaseous nitrogen at 77 K to
gaseous nitrogen at 270 K) is 244 kJ/kg.

To simulate the cooling duty of the precooling step in this example, a user-defined cooling agent named
“GN2+LN2” is employed. This cooling agent has a supply temperature of 77 K, a return temperature of 270
K, and a mass-to-energy factor of 244 kJ/kg. Thus, it represents the enthalpy change of the two refrigerants
- liquid nitrogen and gaseous nitrogen – in the precooling section of the closed-loop nitrogen precooling
and re-liquefaction cycle.

Cryogenic Cooling
After hydrogen is compressed and pre-cooled, it undergoes further cooling in a series of cryogenic heat
exchangers. These heat exchangers use cold hydrogen gas produced in the liquefaction step (P-13 / V-
101) as the refrigerant. It is important to note that normal hydrogen, at near-ambient temperatures, consists
of approximately 75% ortho-hydrogen (o-H2) and 25% para-hydrogen (p-H2). When normal hydrogen is
liquefied and stored in ambient conditions, o-H2 converts into p-H2, releasing enough energy to turn liquid
p-H2 into gas, which can easily escape the tank, causing boil-off. To prevent this, cryogenic heat exchangers
contain catalysts that convert o-H2 into p-H2 during cooling, resulting in almost 100% p-H2 at the end of the
liquefaction process [22].

To simulate the ortho- para- conversion in this example, two user-defined pure components named “o-H2”
and “p-H2” are introduced to the simulation to represent o-H2 and p-H2, respectively. Although the specific
heat capacities of gases normal hydrogen, o-H2, and p-H2 are not identical, for the purposes of this example
these differences are neglected, and therefore, pure components “o-H2” and “p-H2” are just copies of to
the “Hydrogen” pure component.

Since heat exchangers in SuperPro do not allow users to specify reactions that occur during cooling,
reaction generic boxes can be added after the heat exchangers to simulate the ortho- para- conversion
reaction. Because this reaction is temperature-dependent, adding a generic box after each heat exchanger
would make the simulation much more complicated since a diagram of equilibrium p-H2 concentration vs.
temperature requires evaluation [77] in tandem with adjustment of the specified conversion % accordingly
for every change in temperature. For simplicity, this example uses only one reaction generic box (P-5/GBX-
101) added right before the heat exchangers. This reaction generic box is used to simulate the total heat
released by the ortho- para- conversion reaction in the system. Therefore, in this example, the first heat
exchanger bears most of the weight of handling this heat. In the reaction generic box operation, two
sequential reactions are specified: the first reaction simply converts normal hydrogen into a mixture of 75%
o-H2 and 25% p-H2; the second reaction is the ortho-para conversion reaction that converts o-H2 into p-H2.
It is assumed that 100% of o-H2 is converted into p-H2. Also, the reaction is assumed to release 523 kJ/kg
of heat at a reference temperature of 80 K.

P a g e 35 | 50
This example uses three cryogenic heat exchangers in series to cool down the hydrogen to a very low
temperature (P-6/HX-102, P-8/HX-103 and P-11/HX-104). The working fluid in these heat exchangers is
the cold hydrogen gas leaving the gas-liquid separator of the final expansion step (P-13/V-101). In the first
heat exchanger (P-6/HX-102), the hydrogen gas is cooled down to approximately 88 K. A fraction (29%) of
this gas is drawn (P-7/FSP-101) and cooled down separately in a turboexpander (P-9/GE-101). The rest of
the gas is fed into the second heat exchanger (P-8/HX-103), where it is cooled to approximately 34 K and
then into the third heat exchanger (P-11/HX-104), where it cooled to approximately 29 K. The cold hydrogen
gas from the gas-liquid separator of the final expansion step (P-13/V-101) is used as the refrigerant of the
third heat exchanger. It is then mixed with the cooled gas from the turboexpander (P-10/MX-102) and used
as the refrigerant of the first and second heat exchanger. Finally, it is circulated to the compressor (P-1/G-
101) to be used as a process fluid in the hydrogen liquefaction cycle.

By increasing the split percentage to the turbo-expander, the cooling temperatures in all three heat
exchangers can be reduced. However, this also results in an increase in hydrogen reflux, which leads to
an increase in both compression work and cooling duty, ultimately resulting in an increase in the hydrogen
production cost. There is a maximum value and a minimum value to this split percentage. The maximum
value is reached when either the inlet temperature difference of the third heat exchanger or the log mean
temperature difference of the pre-cooler approaches the specified minimum approach temperature (by
default, 5°C). The minimum value is reached when the inlet temperature difference of the first heat
exchanger approaches the minimum approach temperature. For the operating conditions used in this
simulation, the split percentage into the turbo-expander can range from 29% to 32%. To minimize the
hydrogen production cost, the split percentage was adjusted to 29% in this simulation.

Liquefaction
In a typical liquefaction process, the cold hydrogen gas is liquefied by expanding it in a Joule-Thomson
expander, which causes the gas to cool down further due to the Joule-Thomson effect. It is important to
note that under liquefaction conditions, hydrogen behaves like a real gas [22], [77]. However, SuperPro is
an ideal gas simulator, and therefore it treats hydrogen as an ideal gas even under liquefaction conditions.
This can lead to significant errors in the estimation of operating parameters and economic performance
indicators, which may be unacceptable when the simulation accuracy is crucial, such as in the case of
designing and optimizing industrial hydrogen liquefaction cycles. However, the ideal gas assumption may
be sufficient for preliminary capital and operating cost estimations, where an error of 20-30% may be
acceptable. This analysis assumes that the error in estimating the capital and operating expenses of the
hydrogen liquefaction process considered due to the ideal gas assumption does not exceed 20-30%.

Another problem with the ideal gas assumption is that of simulating a Joule-Thomson expander. For an
ideal gas, an isenthalpic expansion is also isothermal, making it impossible to simulate the cooling effect of
a Joule-Thomson expander in SuperPro. To solve this problem, an ideal (isentropic) turbo-expander (P-
12/GE-102) is used to expand the cold hydrogen gas at atmospheric pressure. This produces a two-phase

P a g e 36 | 50
mixture with a vapor fraction of 0.61 at 20.4 K (the normal boiling point of hydrogen, according to SuperPro).
If hydrogen is treated as a real gas, the same expansion results in a two-phase mixture with a vapor fraction
of approximately 0.55 according to the T-s diagram for hydrogen [83]. This means that the error in the
estimation of the final vapor fraction of hydrogen for this example is 11%.

To separate the liquid phase from the gas phase, a flash drum is used (P-13/V-101). As explained
previously, the cryogenic hydrogen gas leaving the flash drum is mixed with the cold hydrogen gas from
the turbo-expander (P-9/GE-101), and the mixture is used as the refrigerant in the two heat exchangers.
Then, it is recycled back to the compressors to be used as a process fluid in the liquefaction cycle.

The liquid hydrogen produced is stored in cryogenic liquid hydrogen spheres (P-14/V-102). The mass
flowrate of produced liquid hydrogen is 1,100 kg/h (26.4 MT/day). Assuming a storage period of 8 hours
and considering that the density of liquid hydrogen is approximately 70.8 g/L, the required storage volume
is around 50 m3. Therefore, two spheres with a net volume of 30 m3 each can be utilized for storing the
liquid hydrogen product.

Cost Analysis

Several assumptions were made regarding the capital and operating expenses for this project. The most
significant ones are:

- The project lifetime is 20 years.


- The depreciation period is 15 years.
- The annual operating time (AOT) is 330 days of continuous operation (7920 hours or 0.90 years).
- The plant consumes renewable electricity generated from solar or wind farms, which is purchased at
an average annual price of $50/MWh [72].
- The cost of inlet hydrogen is set to zero, so the unit production cost of liquid nitrogen reflects only the
cost of hydrogen liquefaction.
- The “Unlisted Equipment Purchase Cost” factor for the “Compression and Precooling” flowsheet section
is set at 84% of the total equipment purchase cost. The above percentage includes the capital cost of
nitrogen re-liquefaction and the cost of other unlisted equipment. Its value was estimated as follows:
The capital cost of the “Compression and Precooling” flowsheet section is around $12.3 million
according to this simulation. The capital cost of nitrogen re-liquefaction is approximately $60 million
based on a simulation of the nitrogen re-liquefaction and hydrogen pre-cooling cycle as an open loop,
which was done in a separate SuperPro file (see file “NitrogenReliquefaction_v#.spf”). Finally, an
additional $3.8 million (corresponding to 5% of the total equipment purchase cost) is assumed for
unlisted equipment.

P a g e 37 | 50
- The “Unlisted Equipment Purchase Cost” factor for the other two flowsheet sections is set at 5% of the
total equipment purchase cost.
- The cost of the “GN2+LN2” heat transfer agent is set at $0.35/MJ, representing the operating cost of
nitrogen re-liquefaction. This cost was calculated based on an open-loop simulation of the nitrogen re-
liquefaction and hydrogen pre-cooling cycle (see file “NitrogenReliquefaction_v#.spf”). Based on that
simulation, the operating cost of nitrogen re-liquefaction is $86.08/MT of refrigerant. To convert this
cost to $/MJ, the above value is divided by the enthalpy change of refrigerant in the pre-cooler, which
is 244 kJ/kg.
- The purchase cost (C) of liquid hydrogen storage spheres as a function of storage volume (Q) is
assumed to be given by the following power law model: C = C0 * (Q/Q0)^a, where the reference capacity
(Q0) is 50 m3, the reference cost (C0) is $790,000 for 2023, and the exponent a is 0.6 [84].
- The selling price of liquid hydrogen is assumed to be $10.10/kg.

Table 19 shows a portion of the Executive Summary taken from the EER for the hydrogen liquefaction
process for the base case scenario where the unit production cost of hydrogen gas is zero.

Table 19: Executive summary for hydrogen liquefaction (if hydrogen gas cost is zero).

EXECUTIVE SUMMARY (2023 prices)


Total Capital Investment 243,581,000 $
Capital Investment Charged to This Project 243,581,000 $
Operating Cost 55,355,000 $/yr
Main Revenue 87,120,000 $/yr
Other Revenues 0 $/yr
Total Revenues 87,120,000 $/yr
Cost Basis Annual Rate 8,712,002 kg MP/yr
Unit Production Cost 6.35 $/kg MP
MP = Total Flow of Stream 'LH2'

According to this table, the total capital investment (TCI) for a hydrogen liquefaction plant is approximately
$244 million, the annual operating cost (AOC) is approximately $55 million, and the net cost of liquefaction
is $6.35/kg. However, this is just the cost of hydrogen liquefaction. To calculate the total production cost of
liquid hydrogen, the production cost of hydrogen gas must also be added. To account for this cost, the
purchasing price of the “Hydrogen” pure component (contained in the hydrogen gas feed) can be set equal
to the unit production cost of hydrogen gas.

Table 20 presents the Executive Summary for the case where the purchasing price of the hydrogen gas
feed is $1.65/kg, resulting in a total unit production cost of liquid hydrogen of $8.00/kg.

P a g e 38 | 50
Table 20: Executive summary for hydrogen liquefaction (if hydrogen gas cost is $1.65/kg).

EXECUTIVE SUMMARY (2023 prices)


Total Capital Investment 244,888,000 $
Capital Investment Charged to This Project 244,888,000 $
Operating Cost 69,729,000 $/yr
Main Revenue 87,991,000 $/yr
Other Revenues 0 $/yr
Total Revenues 87,991,000 $/yr
Cost Basis Annual Rate 8,711,998 kg MP/yr
Unit Production Cost 8.00 $/kg MP
Net Unit Production Cost 8.00 $/kg MP
Unit Production Revenue 10.10 $/kg MP
Gross Margin 20.75 %
Return On Investment 11.58 %
Payback Time 8.64 years
IRR (After Taxes) 8.21 %
NPV (at 7.0% Interest) 21,851,000 $
MP = Total Flow of Stream 'LH2'

To achieve a positive NPV at 7% discount rate and a gross margin of at least 20%, the selling price of
hydrogen must be at least $10.10/kg. At this selling price, the gross margin would be around 21%, the ROI
around 12%, and the payback time approximately 9 years. Table 21 provides a breakdown of total
equipment purchase cost extracted from the EER.

Table 21: Equipment purchase costs for the hydrogen liquefaction example model.

EQUIPMENT SPECIFICATION AND FOB COST (2023 prices)


Main Equipment
Quantity/
Standby/ Name Description Unit Cost ($) Cost ($)
Staggered
1/0/0 G-102 Centrifugal Compressor 7,177,000 7,177,000
Compressor Power = 1.72 MW
1/0/0 G-101 Centrifugal Compressor 4,136,000 4,136,000
Compressor Power = 0.97 MW
1/0/0 HX-101 Heat Exchanger 971,000 971,000
Heat Exchange Area = 368.56 m2
5/0/0 V-102 Horizontal Tank 554,000 2,770,000
Vessel Volume = 27.64 m3
1/0/0 HX-103 Heat Exchanger 361,000 361,000
Heat Exchange Area = 70.93 m2
1/0/0 HX-104 Heat Exchanger 78,000 78,000
Heat Exchange Area = 5.54 m2
1/0/0 GE-101 Gas Expander-Generator 77,000 77,000
Turbine Delivered Shaft Power =
257.83 kW
1/0/0 HX-102 Heat Exchanger 67,000 67,000

P a g e 39 | 50
Heat Exchange Area = 4.32 m2
1/0/0 GE-102 Gas Expander-Generator 6,000 6,000
Turbine Delivered Shaft Power =
229.68 kW
Unlisted Equipment 64,668,000

TOTAL 80,311,000

As explained previously, the unlisted equipment cost represents the capital cost of the nitrogen re-
liquefaction plant and the capital cost of other unlisted equipment. The capital cost of the listed equipment
is approximately $15.6 million. From these, approximately $12.3 million is the cost of the two compressors
and of the pre-cooler, which are included in the “Compression & Precooling” flowsheet section.

Table 22 displays a breakdown of unit production cost extracted from the ICR.

Table 22: Breakdown of unit production cost for the hydrogen liquefaction example model (assuming a
hydrogen gas cost of $1.65/kg).

COST PER PROCESS SECTION (2023 prices)

SUMMARY PER COST ITEM (Entire Process)


Cost Item $/kg MP $/day $/year %
Raw Materials 1.65 43,560 14,374,800 20.62
Facility 2.71 71,580 23,621,268 33.88
Labor 0.06 1,656 546,480 0.78
Consumables 0.00 0 0 0.00
Lab/QC/QA 0.01 201 66,358 0.10
Utilities 3.57 94,304 31,120,413 44.63
Waste Trtmt/Disp 0.00 0 0 0.00
Transportation 0.00 0 0 0.00
Miscellaneous 0.00 0 0 0.00
TOTAL 8.00 211,301 69,729,320 100.00

This table indicates that the utility, facility-dependent and raw materials costs account for approximately
45%, 34%, and 21% of the total operating cost, respectively. The facility-dependent costs include
maintenance, depreciation, and other overhead costs. To gain a better understanding of the contribution of
utilities to the unit production cost, Table 23, also extracted from the ICR, is shown below. This table shows
that the cost of nitrogen re-liquefaction (represented by the cost of cooling agent “GN2+LN2”) represents
nearly 97% of the total utility cost, with the remaining 3% being the electricity cost. The simulation of the
nitrogen re-liquefaction and hydrogen precooling cycle, which was done separately, indicates that the
facility-dependent and electricity costs account for 80% and 19.7%, respectively, of the total nitrogen re-
liquefaction cost. Specifically, the annual electricity consumption for nitrogen re-liquefaction is 135,411,482
kWh. The annual electricity consumption of the rest of the hydrogen liquefaction process is 21,552,518

P a g e 40 | 50
kWh. Also, the annual LH2 production is 8,711,998 kg, as shown in Table 20. Consequently, the specific
energy consumption (SEC) of the hydrogen liquefaction process is 18 kWh/kg.

Table 23: Utility costs for hydrogen liquefaction example (if hydrogen gas cost is $1.65/kg).

UTILITIES COST - BREAKDOWN BY UTILITY TYPE


Std Power
Unit Cost Amount Cost
Section Name %
($/kW-h) (kW-h/year) ($/year)
Compression & Precooling 0.05 21,522,518 1,076,125.90 100.00
Cryogenic Cooling 0.05 546 27.31 0.00
TOTAL 21,523,064 1,076,153.21 100.00

GN2+LN2
Unit Cost Amount Cost
Section Name %
($/MJ) (MJ/year) ($/year)
Compression & Precooling 0.35 85,840,744 30,044,260.27 100.00
TOTAL 85,840,744 30,044,260.27 100.00

Based on the simulation results presented above, the capital investment and net unit production costs of
hydrogen liquefaction are $244 million and $6.35/kg, respectively. In comparison, the US DOE estimated
these costs to be $100 million and $2.75/kg in 2016, as reported in the DOE Hydrogen and Fuel Cells
Program Record 19001 [85]. Adjusting these costs for 2022 using the Chemical Engineering Cost Index
(CEPCI) values for 2016 (762.8) and January 2022 (797.5) yields $147 million and $4.00/kg, respectively.
Furthermore, the average SEC for hydrogen liquefaction is reported in the literature to be 12 kWh/kg.
Therefore, the SuperPro estimates of capital investment, production cost, and SEC seem to be higher than
the respective estimates reported in the literature. This is expected to some degree since industrial
hydrogen liquefaction cycles are often far more complex than the simple Claude cycle employed in this
example to optimize performance and minimize cost. Besides that, the ideal gas assumption is known to
be invalid at hydrogen liquefaction conditions. So, it is very likely that SuperPro overestimates the cryogenic
cooling temperatures due to this assumption. As a result, the hydrogen reflux is overestimated, leading to
an overestimation of compressor sizes and electricity consumption, as well as pre-cooler nitrogen flow and
re-liquefaction cost.

Actual equipment costs could be as much as 20-30% lower, resulting in better project economics. Table 24
compares the simulation results from the previous scenario (scenario 1) with the results from a scenario
where equipment costs are 30% lower (scenario 2). In scenario 2, where equipment costs are reduced by
30%, the operating cost of nitrogen re-liquefaction, as derived from the “NitrogenReliquefaction_v#.spf” file,
is $65.33/MT of refrigerant. The operating cost of the “GN2+LN2” cooling agent is calculated at $0.27/MJ
by dividing the operating cost of nitrogen re-liquefaction by the enthalpy change of refrigerant in the pre-

P a g e 41 | 50
cooler, which is 244 kJ/kg. Lastly, the net liquefaction cost based on the “HydrogenLiquefaction_v#.spf” file
is $4.7/kg.

Table 24: Net liquefaction costs for two scenarios.

Nitrogen Re- “GN2+LN2” Net


Total Capital
liquefaction Cooling Liquefaction
Scenario Investment
Cost ($/MT of Agent Cost Cost
($MM)
refrigerant) ($/MJ) ($/kg)

1 244 86.08 0.35 6.35

2 171 65.33 0.27 4.70

Conclusion

In this work, three conceptual processes to produce 1.1 MT/h of green hydrogen are modeled and
economically evaluated in SuperPro Designer: Biohydrogen production through dark and photo
fermentation, hydrogen production through water electrolysis in a 55 MW PEM electrolyzer system, and
liquefaction of hydrogen gas (Biohydrogen or electrolytic hydrogen) to produce liquid hydrogen.

For the Biohydrogen process, the estimated production cost is $6.7/kg. For the electrolysis process, it is
$5.5/kg if there is no need for additional mechanical compression and storage. These costs are in alignment
with the literature. However, they are considerably higher than the average cost of fossil fuel-based
hydrogen production ($1.5/kg). Consequently, the biohydrogen and electrolytic hydrogen production
methods may not be attractive investments without major government subsidies. Additionally, the net cost
of hydrogen liquefaction is $6.35/kg based on the modeling with SuperPro and $4/kg based on the literature,
which is prohibitive for high-volume consumption applications. Additional research is required on the use
of carrier molecules, such as ammonia and metal hydrides, that facilitate long-distance transportation of
hydrogen by road, rail, and sea [20], [21], [22].

In conclusion, the development of inexpensive and efficient hydrogen generation technologies is crucial for
making green hydrogen production economically competitive with traditional fossil fuel-based methods.
Both technological advances and government incentives are required for driving down production costs and
making green hydrogen a more viable energy solution for a sustainable future.

P a g e 42 | 50
References

[1] A. Pandey, J.-S. Chang, P. C. Hallenbeck, and C. Larroche, Biohydrogen, 1st. edition. Amsterdam
Boston: Elsevier, 2013.

[2] P. Friedlingstein et al., “Global Carbon Budget 2019,” Earth Syst. Sci. Data, vol. 11, no. 4, pp. 1783–
1838, Dec. 2019, doi: 10.5194/essd-11-1783-2019.

[3] “Analysis: Global CO2 emissions from fossil fuels hits record high in 2022,” World Economic Forum,
Nov. 11, 2022. https://www.weforum.org/agenda/2022/11/global-co2-emissions-fossil-fuels-hit-record-
2022/ (accessed Apr. 07, 2023).

[4] J. F. Soares, T. C. Confortin, I. Todero, F. D. Mayer, and M. A. Mazutti, “Dark fermentative biohydrogen
production from lignocellulosic biomass: Technological challenges and future prospects,” Renew.
Sustain. Energy Rev., vol. 117, p. 109484, Jan. 2020, doi: 10.1016/j.rser.2019.109484.

[5] O. US EPA, “Overview of Greenhouse Gases,” Dec. 23, 2015.


https://www.epa.gov/ghgemissions/overview-greenhouse-gases (accessed Apr. 07, 2023).

[6] L. Singh and Z. A. Wahid, “Methods for enhancing bio-hydrogen production from biological process: A
review,” J. Ind. Eng. Chem., vol. 21, pp. 70–80, Jan. 2015, doi: 10.1016/j.jiec.2014.05.035.

[7] İ. Dinçer and H. Ishaq, Renewable hydrogen production. Amsterdam: Elsevier, 2022.

[8] P. Westermann, B. Jorgensen, L. Lange, B. Ahring, and C. Christensen, “Maximizing renewable


hydrogen production from biomass in a bio/catalytic refinery,” Int. J. Hydrog. Energy, vol. 32, no. 17,
pp. 4135–4141, Dec. 2007, doi: 10.1016/j.ijhydene.2007.06.018.

[9] “Global CO2 emissions from energy combustion and industrial processes, 1900-2022 – Charts – Data
& Statistics,” IEA. https://www.iea.org/data-and-statistics/charts/global-co2-emissions-from-energy-
combustion-and-industrial-processes-1900-2022 (accessed Apr. 11, 2023).

[10] M. Sand et al., “A multi-model assessment of the Global Warming Potential of hydrogen,” Commun.
Earth Environ., vol. 4, no. 1, Art. no. 1, Jun. 2023, doi: 10.1038/s43247-023-00857-8.

[11] V. S. Bisaria, “Biohydrogen production: fundamentals and technology advances,” Int. J. Hydrog.
Energy, vol. 39, no. 22, pp. 11828–11829, Jul. 2014, doi: 10.1016/j.ijhydene.2014.06.003.

[12] K. Chandrasekhar, Y.-J. Lee, and D.-W. Lee, “Biohydrogen Production: Strategies to Improve Process
Efficiency through Microbial Routes,” Int. J. Mol. Sci., vol. 16, no. 4, Art. no. 4, Apr. 2015, doi:
10.3390/ijms16048266.

P a g e 43 | 50
[13] A. I. Osman, T. J. Deka, D. C. Baruah, and D. W. Rooney, “Critical challenges in biohydrogen
production processes from the organic feedstocks,” Biomass Convers. Biorefinery, Aug. 2020, doi:
10.1007/s13399-020-00965-x.

[14] A. K. Sarker, A. K. Azad, M. G. Rasul, and A. T. Doppalapudi, “Prospect of Green Hydrogen Generation
from Hybrid Renewable Energy Sources: A Review,” Energies, vol. 16, no. 3, p. 1556, Feb. 2023, doi:
10.3390/en16031556.

[15] N. Azbar and D. B. Levin, “State of the Art and Progress in Production of Biohydrogen,” p. 275.

[16] I. K. Kapdan and F. Kargi, “Bio-hydrogen production from waste materials,” Enzyme Microb. Technol.,
vol. 38, no. 5, pp. 569–582, Mar. 2006, doi: 10.1016/j.enzmictec.2005.09.015.

[17] K. Randolph and S. Studer, “DOE Hydrogen and Fuel Cells Program Record.” Department of Energy,
USA, Feb. 14, 2017. Accessed: Jan. 27, 2023. [Online]. Available:
https://www.hydrogen.energy.gov/pdfs/16016_h2_production_cost_fermentation.pdf

[18] V. Singh Yadav, V. R, and D. Yadav, “Bio-hydrogen production from waste materials: A review,”
MATEC Web Conf., vol. 192, p. 02020, 2018, doi: 10.1051/matecconf/201819202020.

[19] K. Pandu and S. Joseph, “Comparison-and-Limitations-of-Biohydrogen-Production-Processes.pdf,”


ResearchGate, May 2012, Accessed: Mar. 16, 2023. [Online]. Available:
https://www.researchgate.net/profile/Karthic-
Pandu/publication/235993414_Comparison_and_Limitations_of_Biohydrogen_Production_Processes
/links/02e7e5154fe1859fa1000000/Comparison-and-Limitations-of-Biohydrogen-Production-
Processes.pdf

[20] J. Wang and Y. Yin, Biohydrogen Production from Organic Wastes. in Green Energy and Technology.
Singapore: Springer Singapore, 2017. doi: 10.1007/978-981-10-4675-9.

[21] C. Jackson et al., “Ammonia to Green Hydrogen Project,” Siemens, Engie, STFC, Equity, Feasibility
Study, 2019. [Online]. Available:
https://assets.publishing.service.gov.uk/government/uploads/system/uploads/attachment_data/file/88
0826/HS420_-_Ecuity_-_Ammonia_to_Green_Hydrogen.pdf

[22] S. ZS. Al Ghafri et al., “Hydrogen liquefaction: a review of the fundamental physics, engineering
practice and future opportunities,” Energy Environ. Sci., vol. 15, no. 7, pp. 2690–2731, 2022, doi:
10.1039/D2EE00099G.

[23] “H2 Fuel Technology.” https://h2-fuel.nl/technology (accessed May 12, 2023).

[24] “No. 182 Hydrogen Energy Supply Chain | Kawasaki Heavy Industries, Ltd.”
https://global.kawasaki.com/en/corp/rd/magazine/182/index.html (accessed Apr. 28, 2023).

P a g e 44 | 50
[25] “Hydrogen Production — Development of Hydrogen Liquefaction Systems.”

[26] “Liebreich: The Unbearable Lightness of Hydrogen | BloombergNEF.”


https://about.bnef.com/blog/liebreich-the-unbearable-lightness-of-hydrogen/?linkId=193361832
(accessed Apr. 29, 2023).

[27] A. Demirbas, Biohydrogen. in Green Energy and Technology. London: Springer London, 2009. doi:
10.1007/978-1-84882-511-6.

[28] P. Nikolaidis and A. Poullikkas, “A comparative overview of hydrogen production processes,” Renew.
Sustain. Energy Rev., vol. 67, pp. 597–611, Jan. 2017, doi: 10.1016/j.rser.2016.09.044.

[29] I. Ntaikou, G. Antonopoulou, and G. Lyberatos, “Biohydrogen Production from Biomass and Wastes
via Dark Fermentation: A Review,” Waste Biomass Valorization, vol. 1, no. 1, pp. 21–39, Mar. 2010,
doi: 10.1007/s12649-009-9001-2.

[30] P. K. Sarangi and S. Nanda, “Biohydrogen Production Through Dark Fermentation,” Chem. Eng.
Technol., vol. 43, no. 4, pp. 601–612, Apr. 2020, doi: 10.1002/ceat.201900452.

[31] E. L. N. Dzulkarnain, J. O. Audu, W. R. Z. Wan Dagang, and M. F. Abdul-Wahab, “Microbiomes of


biohydrogen production from dark fermentation of industrial wastes: current trends, advanced tools and
future outlook,” Bioresour. Bioprocess., vol. 9, no. 1, p. 16, Dec. 2022, doi: 10.1186/s40643-022-00504-
8.

[32] R. Łukajtis et al., “Hydrogen production from biomass using dark fermentation,” Renew. Sustain.
Energy Rev., vol. 91, pp. 665–694, Aug. 2018, doi: 10.1016/j.rser.2018.04.043.

[33] A. Reungsang, N. Zhong, Y. Yang, S. Sittijunda, A. Xia, and Q. Liao, “Hydrogen from Photo
Fermentation,” in Bioreactors for Microbial Biomass and Energy Conversion, Q. Liao, J. Chang, C.
Herrmann, and A. Xia, Eds., in Green Energy and Technology. Singapore: Springer Singapore, 2018,
pp. 221–317. doi: 10.1007/978-981-10-7677-0_7.

[34] C. N. C. Hitam and A. A. Jalil, “A review on biohydrogen production through photo-fermentation of


lignocellulosic biomass,” Biomass Convers. Biorefinery, Nov. 2020, doi: 10.1007/s13399-020-01140-
y.

[35] D. Deo, E. Ozgur, I. Eroglu, U. Gunduz, and M. Yucel, “Photofermentative Hydrogen Production in
Outdoor Conditions,” in Hydrogen Energy - Challenges and Perspectives, D. Minic, Ed., InTech, 2012.
doi: 10.5772/50390.

[36] Q. Zhang and Z. Zhang, “Biological Hydrogen Production From Renewable Resources by
Photofermentation,” in Advances in Bioenergy, Elsevier, 2018, pp. 137–160. doi:
10.1016/bs.aibe.2018.03.001.

P a g e 45 | 50
[37] W. M. Alalayah, M. S. Kalil, A. a. H. Kadhum, J. M. Jahim, S. Z. S. Jaapar, and N. M. Alauj, “Bio-
Hydrogen Production using a Two-Stage Fermentation Process,” Pak. J. Biol. Sci., vol. 12, no. 22, pp.
1462–1467, doi: 10.3923/pjbs.2009.1462.1467.

[38] E. Özgür, B. Uyar, Y. Öztürk, M. Yücel, U. Gündüz, and I. Eroğlu, “Biohydrogen production by
Rhodobacter capsulatus on acetate at fluctuating temperatures,” Resour. Conserv. Recycl., vol. 54,
no. 5, pp. 310–314, Mar. 2010, doi: 10.1016/j.resconrec.2009.06.002.

[39] A. Ghimire et al., “A review on dark fermentative biohydrogen production from organic biomass:
Process parameters and use of by-products,” Appl. Energy, vol. 144, pp. 73–95, Apr. 2015, doi:
10.1016/j.apenergy.2015.01.045.

[40] “Overview of Biological Hydrogen Production.” https://www.biotecharticles.com/Applications-


Article/Overview-of-Biological-Hydrogen-Production-3462.html (accessed Dec. 19, 2022).

[41] A. Singh and D. Rathore, Eds., Biohydrogen Production: Sustainability of Current Technology and
Future Perspective. New Delhi: Springer India, 2017. doi: 10.1007/978-81-322-3577-4.

[42] J. K. Saini, R. Saini, and L. Tewari, “Lignocellulosic agriculture wastes as biomass feedstocks for
second-generation bioethanol production: concepts and recent developments,” 3 Biotech, vol. 5, no. 4,
pp. 337–353, Aug. 2015, doi: 10.1007/s13205-014-0246-5.

[43] R. P. Pandey, R. Raghulchandrana, A. Tamizhini, and A. V. Snehya, “Pretreatment Studies of


Biohydrogen Production from Agro-Industrial Waste,” EARTH SCIENCES, preprint, Nov. 2020. doi:
10.20944/preprints202011.0320.v1.

[44] J. Wang and Y. Yin, “Pretreatment of Organic Wastes for Hydrogen Production,” in Biohydrogen
Production from Organic Wastes, in Green Energy and Technology. Singapore: Springer Singapore,
2017, pp. 123–195. doi: 10.1007/978-981-10-4675-9_4.

[45] V. L. Martínez, G. L. Salierno, R. E. García, M. J. Lavorante, M. A. Galvagno, and M. C. Cassanello,


“Biological Hydrogen Production by Dark Fermentation in a Stirred Tank Reactor and Its Correlation
with the pH Time Evolution,” Catalysts, vol. 12, no. 11, p. 1366, Nov. 2022, doi:
10.3390/catal12111366.

[46] A. A. Abreu, D. Karakashev, I. Angelidaki, D. Z. Sousa, and M. M. Alves, “Biohydrogen production from
arabinose and glucose using extreme thermophilic anaerobic mixed cultures,” Biotechnol. Biofuels, vol.
5, no. 1, p. 6, Dec. 2012, doi: 10.1186/1754-6834-5-6.

[47] S. Reaume, “Fermentation of Glucose and Xylose to Hydrogen in the Presence of Long Chain Fatty
Acids,” p. 128.

P a g e 46 | 50
[48] D. B. Levin, L. Pitt, and M. Love, “Biohydrogen production: prospects and limitations to practical
application,” Int. J. Hydrog. Energy, vol. 29, no. 2, pp. 173–185, Feb. 2004, doi: 10.1016/S0360-
3199(03)00094-6.

[49] C.-Y. Chen, Y.-C. Lo, K.-L. Yeh, and J.-S. Chang, “Biohydrogen production from combined dark-photo
fermentation under a high ammonia content in the dark fermentation effluent,” 2010.

[50] H. Zürrer and R. Bachofen, “Aspects of growth and hydrogen production of the photosynthetic
bacterium Rhodospirillum rubrum in continuous culture,” Biomass, vol. 2, no. 3, pp. 165–174, Jul. 1982,
doi: 10.1016/0144-4565(82)90027-0.

[51] T. Assawamongkholsiri and A. Reungsang, “Photo-fermentational hydrogen production of


Rhodobacter sp. KKU-PS1 isolated from an UASB reactor,” Electron. J. Biotechnol., vol. 18, no. 3, pp.
221–230, May 2015, doi: 10.1016/j.ejbt.2015.03.011.

[52] B. Wang, W. Wan, and J. Wang, “Effect of ammonia concentration on fermentative hydrogen
production by mixed cultures,” Bioresour. Technol., vol. 100, no. 3, pp. 1211–1213, Feb. 2009, doi:
10.1016/j.biortech.2008.08.018.

[53] R. Sirohi, A. Pandey, S. Jun Sim, J.-S. Chang, and D.-J. Lee, CURRENT DEVELOPMENTS IN
BIOTECHNOLOGY AND BIOENGINEERING photobioreactors. S.l.: ELSEVIER - HEALTH SCIENCE,
2022.

[54] H. Argun and F. Kargi, “Effects of light source, intensity and lighting regime on bio-hydrogen production
from ground wheat starch by combined dark and photo-fermentations,” Int. J. Hydrog. Energy, vol. 35,
no. 4, pp. 1604–1612, Feb. 2010, doi: 10.1016/j.ijhydene.2009.12.033.

[55] “Lux to watts (W) conversion calculator.” https://www.rapidtables.com/calc/light/lux-to-watt-


calculator.html (accessed Apr. 20, 2023).

[56] D. M. Ruthven, S. Farooq, and K. S. Knaebel, Pressure swing adsorption. New York, N.Y: VCH
Publishers, 1994.

[57] J. Byun and J. Han, “Economically feasible production of green methane from vegetable and fruit-rich
food waste,” Energy, vol. 235, p. 121397, Nov. 2021, doi: 10.1016/j.energy.2021.121397.

[58] S. F. Ahmed et al., “Biohydrogen Production From Biomass Sources: Metabolic Pathways and
Economic Analysis,” Front. Energy Res., vol. 9, p. 753878, Sep. 2021, doi:
10.3389/fenrg.2021.753878.

[59] R. Davis et al., “Process Design and Economics for the Conversion of Lignocellulosic Biomass to
Hydrocarbons: Dilute-Acid and Enzymatic Deconstruction of Biomass to Sugars and Biological

P a g e 47 | 50
Conversion of Sugars to Hydrocarbons,” NREL/TP-5100-60223, 1107470, Oct. 2013. doi:
10.2172/1107470.

[60] S. Metz et al., “Producing hydrogen through electrolysis and other processes,” in Hydrogen
Technologies, R. Neugebauer, Ed., Cham: Springer International Publishing, 2023, pp. 203–252. doi:
10.1007/978-3-031-22100-2_9.

[61] N. Connell et al., “Green Hydrogen Guidebook.” Green Hydrogen Coalition, Apr. 2022. [Online].
Available: https://www.ghcoalition.org/education

[62] “What is an Electrolyzer and What is it Used for? | Accelera,” Feb. 20, 2023.
https://www.accelerazero.com/news/what-is-an-electrolyzer-and-what-is-it-used-for (accessed Apr.
14, 2023).

[63] V. Knop, “A World Of Energy - PEM electrolyser.” http://www.awoe.net/Water-Electrolysis-PEM-


Technology.html (accessed Apr. 23, 2023).

[64] “Electrolyser technologies – FinH2.” https://www.finh2.fi/electrolyser-technologies/ (accessed Apr. 23,


2023).

[65] C. Haas, M.-G. Macherhammer, N. Klopcic, and A. Trattner, “Capabilities and Limitations of 3D-CFD
Simulation of Anode Flow Fields of High-Pressure PEM Water Electrolysis,” Processes, vol. 9, no. 6,
p. 968, May 2021, doi: 10.3390/pr9060968.

[66] A. Pashaei, “Hydrogen production with water electrolysis method to use in fuel cell for electricity
generation,” 2014, doi: 10.13140/RG.2.1.3114.2565.

[67] “Cell Design, Fabrication, and Performance,” Ebrary.


https://ebrary.net/134224/engineering/cell_design_fabrication_performance (accessed Apr. 23, 2023).

[68] S. Sood et al., “Generic Dynamical Model of PEM Electrolyser under Intermittent Sources,” Energies,
vol. 13, no. 24, p. 6556, Dec. 2020, doi: 10.3390/en13246556.

[69] D19 Committee, “Specification for Reagent Water,” ASTM International. doi: 10.1520/D1193-06R18.

[70] “Purified water,” Wikipedia. Mar. 13, 2023. Accessed: Apr. 14, 2023. [Online]. Available:
https://en.wikipedia.org/w/index.php?title=Purified_water&oldid=1144467131

[71] H. Mässgård and A. Jonsson, “An Industrial Perspective on Ultrapure Water Production for
Electrolysis : A techno-economic assessment of membrane distillation for electrolysis - synergies,
performance, costs, and value propositions,” Diss., KTH, 2021. [Online]. Available: http://kth.diva-
portal.org/smash/record.jsf?pid=diva2%3A1575929&dswid=-1896

P a g e 48 | 50
[72] “How to evaluate the cost of the green hydrogen business case? - KPMG Belgium,” KPMG, Aug. 19,
2022. https://kpmg.com/be/en/home/insights/2022/08/eng-how-to-evaluate-the-cost-of-the-green-
hydrogen-business-case.html (accessed Apr. 23, 2023).

[73] G. Bilicic and S. Scroggins, “Lazard’s LCOE+ Report April 2023,” Lazard, Apr. 2023. [Online]. Available:
https://www.lazard.com/research-insights/2023-levelized-cost-of-energyplus/

[74] IRENA, “Green Hydrogen Cost Reduction: Scaling Up Electrolyzers to Meet the 1.5°C Climate Goal,”
International Renewable Energy Agency, Abu Dhabi, 2020. [Online]. Available: https://www.irena.org/-
/media/Files/IRENA/Agency/Publication/2020/Dec/IRENA_Green_hydrogen_cost_2020.pdf

[75] J. Vickers, D. Peterson, and K. Randolph, “Cost of Electrolytic Hydrogen Production with Existing
Technology,” US Department of Energy, DOE Hydrogen and Fuel Cells Program Record 20004, Sep.
2020. [Online]. Available: https://www.hydrogen.energy.gov/pdfs/20004-cost-electrolytic-hydrogen-
production.pdf

[76] G. Gardner, “Can the Inflation Reduction Act unlock a green hydrogen economy?,” International
Council on Clean Transportation, Jan. 03, 2023. https://theicct.org/ira-unlock-green-hydrogen-jan23/
(accessed Apr. 18, 2023).

[77] W. L. Staats, “Analysis of a Supercritical Hydrogen Liquefaction Cycle,” Massachusetts Institute of


Technology. Department of Mechanical Engineering, 2008. [Online]. Available:
http://dspace.mit.edu/handle/1721.1/7582

[78] K. Ohlig and L. Decker, “The latest developments and outlook for hydrogen liquefaction technology,”
presented at the ADVANCES IN CRYOGENIC ENGINEERING: Transactions of the Cryogenic
Engineering Conference - CEC, Anchorage, Alaska, USA, 2014, pp. 1311–1317. doi:
10.1063/1.4860858.

[79] A. Abdi, J. Chiu, and V. Martin, “State of the Art in Hydrogen Liquefaction,” in Proceedings of the ISES
Solar World Congress 2019, Santiago, Chile: International Solar Energy Society, 2019, pp. 1–10. doi:
10.18086/swc.2019.23.01.

[80] R. B. Scott, W. H. Denton, and C. M. Nicholls, Technology and Uses of Liquid Hydrogen. Burlington:
Elsevier Science, 1964.

[81] K. Isano, T. Komiya, Y. Matsuda, D. Kariya, and S. Kamiya, “Hydrogen Production — Development of
Hydrogen Liquefaction Systems,” Kawasaki Technical Review No. 182, pp. 23–28, Feb. 2021. [Online].
Available: https://www.kawasaki-gasturbine.de/en/company/press/237-kawasaki-technical-review-no-
182

P a g e 49 | 50
[82] J. Yang, Y. Li, and H. Tan, “Study on Performance Comparison of Two Hydrogen Liquefaction
Processes Based on the Claude Cycle and the Brayton Refrigeration Cycle,” Processes, vol. 11, no.
3, p. 932, Mar. 2023, doi: 10.3390/pr11030932.

[83] M. Hirscher, Ed., Handbook of Hydrogen Storage: New Materials for Future Energy Storage, 1st ed.
Wiley, 2010. doi: 10.1002/9783527629800.

[84] EERE, “Multiyear Research, Development and Demonstration Plan,” US Department of Energy,
Technical Report 1219578, Dec. 2014. doi: 10.2172/1219578.

[85] E. Connelly, M. Penev, A. Elgowainy, and C. Hunter, “Current Status of Hydrogen Liquefaction Costs,”
US Department of Energy, DOE Hydrogen and Fuel Cells Program Record 19001, Aug. 2019. [Online].
Available: https://www.hydrogen.energy.gov/pdfs/19001_hydrogen_liquefaction_costs.pdf

P a g e 50 | 50
View publication stats

You might also like