Integration of the Ca–Cu Process in Ammonia ... - ACS Publications

Feb 3, 2017 - Integration of the Ca–Cu Process in Ammonia Production Plants ... Copyright © 2017 American Chemical Society ... Analysis of a Novel ...
0 downloads 0 Views 1MB Size
Subscriber access provided by University of Newcastle, Australia

Article

Integration of the Ca-Cu process in ammonia production plants Isabel Martínez, David Armaroli, Matteo Gazzani, and Matteo Carmelo Romano Ind. Eng. Chem. Res., Just Accepted Manuscript • DOI: 10.1021/acs.iecr.6b04615 • Publication Date (Web): 03 Feb 2017 Downloaded from http://pubs.acs.org on February 18, 2017

Just Accepted “Just Accepted” manuscripts have been peer-reviewed and accepted for publication. They are posted online prior to technical editing, formatting for publication and author proofing. The American Chemical Society provides “Just Accepted” as a free service to the research community to expedite the dissemination of scientific material as soon as possible after acceptance. “Just Accepted” manuscripts appear in full in PDF format accompanied by an HTML abstract. “Just Accepted” manuscripts have been fully peer reviewed, but should not be considered the official version of record. They are accessible to all readers and citable by the Digital Object Identifier (DOI®). “Just Accepted” is an optional service offered to authors. Therefore, the “Just Accepted” Web site may not include all articles that will be published in the journal. After a manuscript is technically edited and formatted, it will be removed from the “Just Accepted” Web site and published as an ASAP article. Note that technical editing may introduce minor changes to the manuscript text and/or graphics which could affect content, and all legal disclaimers and ethical guidelines that apply to the journal pertain. ACS cannot be held responsible for errors or consequences arising from the use of information contained in these “Just Accepted” manuscripts.

Industrial & Engineering Chemistry Research is published by the American Chemical Society. 1155 Sixteenth Street N.W., Washington, DC 20036 Published by American Chemical Society. Copyright © American Chemical Society. However, no copyright claim is made to original U.S. Government works, or works produced by employees of any Commonwealth realm Crown government in the course of their duties.

Page 1 of 35

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Industrial & Engineering Chemistry Research

Integration of the Ca-Cu process in ammonia production plants Isabel Martínez1*, David Armaroli1, Matteo Gazzani2, Matteo C. Romano1 1

Politecnico di Milano, Department of Energy, Via Lambruschini 4, 20156 Milano (Italy) 2

ETHZ, Institute of Process Engineering, Sonneggstrasse 3, 8092 Zurich, Switzerland

*corresponding author: [email protected] Abstract In this work, the application of a Ca-Cu process into a state-of-the-art ammonia production plant is assessed through process simulations, aiming at reducing the primary energy needs of the syngas production and purification section. The proposed process has shown to significantly reduce the specific primary energy consumption, even when accounting for the higher electric consumption associated to the Ca-Cu process. From an environmental point of view, the proposed process has an inherently high CO2 capture efficiency (about 97%). The encouraging thermodynamic performance along with some simplifications in the syngas conditioning process and the ammonia synthesis loop, confirm the potential of the Ca-Cu process as route for syngas generation in ammonia plants.

Keywords: ammonia production, urea production, sorption enhanced reforming, chemical looping, CCS

Nomenclature ATR

Autothermal reformer

CAPEX

Capital expenditure

CCR

Carbon capture ratio

CCS

Carbon capture and sequestration

CLC

Chemical looping combustion

ECO

Economizer

EVA

Evaporator

FTR

Fired tubular reformer

HP

High pressure

HT

High temperature

HX

Heat exchanger

LHV

Lower heating value

LP

Low pressure

LT

Low temperature

MDEA

Methyldiethanolamine

NG

Natural gas

OPEX

Operating expenditure 1 ACS Paragon Plus Environment

Industrial & Engineering Chemistry Research

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

PSA

Pressure swing adsorption

SER

Sorption-enhanced reforming

SH

Superheater

SMR

Steam methane reforming

SPECCA

Specific primary energy consumption for CO2 avoided

WGS

Water gas shift

2 ACS Paragon Plus Environment

Page 2 of 35

Page 3 of 35

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Industrial & Engineering Chemistry Research

1

1. Introduction

2

The chemical and petrochemical sector is the largest industrial consumer of energy, accounting for 30% of

3

the total industrial energy use1. Within this sector, ammonia production accounts for 17% of the energy

4

consumed, being the second largest inorganic chemical produced worldwide (137 million tons in 2012)2.

5

Even though ammonia is used in a diversified set of industrial sectors (e.g. as precursor to nitrogen

6

compounds for fertilizers, explosives, fibers and plastics, as refrigerant or as remediation for gas emissions),

7

its principal use nowadays is in the fertilizer industry, with more than 80% of the ammonia produced

8

worldwide dedicated to this sector3. Accordingly, the increase in ammonia production in the last decades has

9

been driven by the demographic growth of developing countries such as China, and it is therefore expected

10

to keep growing in the coming decades. The major ammonia producer worldwide is China, which contributes

11

with more than 32% of the global production, followed by India, US and Russia. While in China the largest

12

fraction of ammonia is produced starting from coal gasification, about 80% of the world production is

13

provided by steam reforming of natural gas. In this case, natural gas is used both as source of hydrogen

14

needed to form NH3, and as fuel for supplying the process energy needs. In such plant, approximately 2/3 of

15

the natural gas introduced is used as a feedstock whereas the remaining input is used for providing the

16

energy needs. As a consequence, ammonia production costs are greatly influenced by the natural gas cost,

17

which may represent up to 70-85% of the final production cost2.

18

Ammonia production consists of three main steps. Firstly, a syngas containing the required stoichiometric

19

ratio between hydrogen and nitrogen for ammonia synthesis (reaction (1)) is produced by steam reforming of

20

natural gas. Although different reforming configurations exist for syngas generation in ammonia plants, the

21

most common layout makes use of a two-step reforming process4. A primary externally-fired tubular

22

reformer is used to convert natural gas in presence of a Ni-based catalyst at around 750-850°C. Energy

23

needed in this primary reforming is supplied by the oxidation of natural gas and wastes from the syngas

24

cleaning. Then, a secondary reforming step, which typically consists of an autothermal reformer (ATR) using

25

air as oxidant, is adopted to introduce the right amount of nitrogen while converting the remaining natural

26

gas load. This secondary reforming step takes place at a higher temperature (around 1000°C) to achieve high

27

conversion of methane. In the second step of the ammonia production process, syngas is conditioned to meet

28

the specifications of the ammonia conversion step: CO is shifted to H2 in a two-step water gas shift (WGS)

29

section, CO2 is removed and, finally, a methanation step reduces the CO content to ppm level. It is however

30

worth mentioning that all main licensors and engineering constructors have developed proprietary solutions

31

for syngas production and conditioning, which also comprise alternative processes as pressure swing

32

adsorption or cryogenic separation4. Finally, in the third main step the purified synthesis gas is compressed

33

and sent to the ammonia synthesis reactor, which typically operates at a pressure of around 150-250 bar4.

34 35

N + 3H ↔ 2NH ∆H = −46 kJ/mol

36

3 ACS Paragon Plus Environment

(1)

Industrial & Engineering Chemistry Research

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 4 of 35

37

Following the introduction of the disruptive single-train plants by Kellogg in 1963, different modifications

38

and advancements to the well-established ammonia production process have been proposed and studied,

39

mainly with the aim of (i) reducing the specific energy consumption per unit of ammonia produced and of

40

(ii) increasing the plant capacity. Many of the modifications proposed concern the reforming step, which is

41

indeed responsible for a large share of the CAPEX and OPEX. Concerning the latter, it is of paramount

42

interest to reduce the amount of fuel needed for sustaining the reforming reactions in the fired tubular

43

reformer, which is conducive to reduction in the primary energy consumption. However, despite the

44

numerous improvements achieved in the last decades, the specific energy requirement for ammonia

45

production is still significantly larger than the theoretical minimum4.

46

Despite being the preferred process for syngas production at medium-large scale, steam methane reforming

47

(SMR) of natural gas features important drawbacks: the need to keep the reformer size limited requires to

48

operate the system at moderately high pressure (30-50 bar) while high methane conversion is achieved

49

operating the system at high temperature (methane conversion, which follows an endothermic reaction,

50

approaches equilibrium at the outlet of the reformer). With the aim of overcoming the thermodynamic

51

restrictions imposed by this configuration, multiple studies in the scientific literature present enhanced

52

processes, especially where the natural gas reforming is combined with the CO2 separation. The sorption

53

enhanced reforming (SER) concept has received great attention in the past decades due to some inherent

54

advantages over conventional reforming processes5. In particular, the SER process makes use of a CO2

55

acceptor material (usually a Ca-based sorbent) in the reformer, which allows reforming (reaction (2)), water

56

gas shift (reaction (3)) and carbonation (reaction (4)) reactions to occur simultaneously in the same reactor6–

57

8

58

towards the hydrogen production, leading to syngas with a hydrogen content higher than 96%vol. (in dry

59

basis) in a single step and resulting in a much intensified process than SMR. Moreover, the energy released

60

by CO2 carbonation drives the reforming reaction, enabling the use of adiabatic reactors, avoiding the

61

necessity of heat transfer at high temperature through pressurized tube walls.

. The CO2 removal from the gas phase pushes the equilibrium of the reforming and water gas shift reactions

62 63

CH + H O ↔ CO + 3H ∆H = +206.2 kJ/mol

(2)

64

CO + H O ↔ CO + H ∆H = −41.5 kJ/mol

(3)

65

CaO + CO ↔ CaCO ∆H = −178.8 kJ/mol

(4)

66 67

The main challenge of the SER concept is the regeneration of the CaCO3 formed by the carbonation reaction,

68

which is highly endothermic and requires additional fuel to provide the energy needed. Among the different

69

solutions proposed in literature for supplying the CaCO3 regeneration energy, the configuration that couples

70

an exothermic reduction reaction of a CuO-based material with the calcination step is particularly promising,

71

and it has thus received great attention in the last years9–11. The sequence of reactions in this novel process is

72

depicted in Figure 1. The proposed process is carried out in a series of fixed bed reactors, where temperature,

73

pressure, feeds and products vary during the different stages of the process. The first step (indicated as A in 4 ACS Paragon Plus Environment

Page 5 of 35

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Industrial & Engineering Chemistry Research

74

Figure 1) comprises the hydrogen production through a SER process according to the principles discussed

75

above. The solid bed contains the Ca-based sorbent, the Ni-based reforming catalyst and the Cu-based

76

material, all in their reduced form as a result of the preceding reduction/calcination stage of the Ca-Cu

77

process. When feeding natural gas and steam, the Ca-based sorbent reacts to form CaCO3, while the

78

reforming catalyst and the Cu-based material remain inert. This stage of the process is carried out at high

79

pressure (10-30 bar). During the following step of the process (B in Figure 1), Cu present in the solid bed is

80

oxidized to CuO by feeding pressurized diluted air into the reactor. Diluted air enables a fine tuning of the

81

temperature and oxygen concentration at reactor inlet, which need to be limited for controlling the maximum

82

temperature achieved and so minimize the CaCO3 decomposition. In this step, virtually all the O2 in the inlet

83

air reacts with the oxygen carrier and the resulting off-gas is high purity N2 with small amounts of CO2 and

84

Ar. In the last step (stage C), the calcination of the CaCO3 formed during stage A is driven by the exothermic

85

reduction of the CuO, which is carried out feeding a mixture of CH4, CO and H2 into the reactor. With the

86

aim of limiting the calcination temperature needed, this last stage of the process operates at around

87

atmospheric pressure. The Ca/Cu molar ratio in the solid bed is chosen for ensuring that during stage C the

88

heat released in the CuO reduction is enough for sustaining the endothermic CaCO3 calcination.

89 90

[FIGURE 1]

91 92

A full process design of a stand-alone hydrogen production plant based on this process was investigated by

93

Martinez et al.11, who showed that a Ca-Cu process can produce H2 with high efficiency while also capturing

94

more than 90% of the produced CO2. However, the Ca-Cu process has never been investigated as part of the

95

syngas generation and purification in an ammonia production plant, despite the inherent advantage of

96

providing almost pure streams of hydrogen and nitrogen as part of its products. Accordingly, this work

97

addresses the integration of the novel Ca-Cu looping process within an ammonia production plant. A case

98

study of a full system integration of the Ca-Cu process into an ammonia production plant is presented.

99

Modifications of the ammonia synthesis loop when integrated with this concept are investigated with the aim

100

of assessing the impact from a full process performance point of view. The performance obtained for the

101

novel Ca-Cu based ammonia plant is compared with a benchmark ammonia plant based on the state-of-the-

102

art technologies.

103 104

2. Reference ammonia production plant

105

A state-of-the-art, 1500 tons/day, ammonia plant is considered as benchmark for comparison with the

106

advanced ammonia production process integrated with the Ca-Cu concept analyzed in this work; to this end,

107

mass and energy balances of the syngas production and ammonia synthesis area in the reference plant were

108

solved. A schematic of the reference ammonia production plant, which represents the most widespread

109

configuration for ammonia production, is depicted in Figure 2 and Figure 3. The following three processing

110

sections characterize the reference ammonia production plant: (i) the synthesis gas production island, 5 ACS Paragon Plus Environment

Industrial & Engineering Chemistry Research

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

111

covering natural gas reforming processes as well as synthesis gas conditioning and purification steps (i.e.

112

water gas shift reactors, CO2 removal and methanation stages), (ii) the ammonia production loop, and (iii)

113

steam cycle (not shown in the figures) fed by steam produced by recovering the heat available in the plant.

114

These balances have been solved using Aspen Plus. For the synthesis gas production island, the cubic

115

equation of state of Redlich-Kwong-Soave (RKS)12 has been used, whereas for the ammonia synthesis loop

116

the RKS equation modified by Mathias et al. (1984)13 has been preferred due to its suitability in predicting

117

real gas behavior. For pure water and steam, thermodynamic properties from the International Association

118

for the Properties of Water and Steam (IAPWS)14 have been used. Tables S1 and S2 of the Supporting

119

Information summarize the main assumptions used for calculating the synthesis gas production island and

120

the ammonia production loop, respectively, in the reference ammonia plant.

121 122

[FIGURE 2]

123

[FIGURE 3]

124 125

Catalyst used for steam reforming processes is highly sensitive to any sulphur compound. Therefore, the

126

concentration of these compounds has to be below 0.15 ppm before feeding the NG into the reforming

127

reactor. Typically, a non-regenerative ZnO-based solid bed is used for removing sulphur compounds in a

128

two-stage process. Sulphur compounds are first converted into hydrogen sulphide (H2S) by hydrogenation in

129

a catalytic solid bed made of Co and Mo oxides supported on alumina at temperatures not higher than 340-

130

370°C15. A H2 concentration of around 2%vol. is needed at the inlet of this hydrogenation step for ensuring

131

that H2S formation proceeds at sufficient rate16; thus, part of the hydrogen leaving the PSA unit adopted in

132

the synthesis section of the plant is recycled to the feed. Once sulphur compounds are in the form of H2S, the

133

gas is fed to the ZnO-based solid bed that operates at the same temperature of the hydrogenation process. In

134

this reactor, H2S reacts with ZnO to form ZnS, and the H2S concentration in the gas phase drops below 0.1

135

ppm15.

136

After desulphurization, NG is mixed with high pressure steam produced within the plant and fed into an

137

adiabatic pre-reformer. The steam-to-carbon (S/C) ratio typically employed in ammonia production plants

138

ranges between 2.5 to 3.6, with a growing trend towards the lower bound in the most modern plants17. In this

139

work, an S/C ratio of 3 has been chosen for the reference plant. The pre-reforming stage enables the

140

decomposition of higher hydrocarbons in the NG into CO and H2 before entering the primary reforming

141

section, where the high temperature may cause solid carbon formation from these hydrocarbons and

142

consequent catalyst deactivation. In addition, the use of this pre-reformer reactor reduces the reforming duty

143

needed in the subsequent primary reforming section. Pre-reforming is typically carried out in an adiabatic

144

reactor with an inlet gas temperature of around 380-500°C in the presence of a Ni-based catalyst18. For the

145

reference plant, a pre-reformer inlet temperature of 490°C has been chosen. Pre-reformed gas is then heated

146

up to 620°C before being fed to the primary reforming reactor.

6 ACS Paragon Plus Environment

Page 6 of 35

Page 7 of 35

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Industrial & Engineering Chemistry Research

147

The primary fired tubular reformer (FTR) consists of tubes filled with catalyst that are placed inside a

148

furnace where CO/H2/CH4-rich off-gas from a downstream PSA unit and additional NG are burnt with air for

149

supplying the energy needed for the reforming. Typically, CH4 conversions in the range of 55-65% are

150

achieved in the primary reforming process working with outlet temperatures of 780-830°C and leading to

151

CH4 concentrations in the product gas of 10-15% on dry basis4,17. Such values of CH4 conversion at FTR

152

outlet are below the conversion values corresponding to equilibrium at those FTR outlet temperatures. In the

153

reference plant considered in this work, FTR outlet conditions of temperature of 820°C and methane

154

conversion 58%, have been assumed, which lead to a CH4 content in the product gas of 12.5% on dry basis.

155

The sensible heat of the flue gas, which exits the FTR furnace at 1010°C, is recovered heating up the FTR

156

charge fed to the primary reforming section and the air streams needed in the FTR furnace burners and in the

157

secondary reformer. Secondary reforming process is based on an autothermal reformer where air is fed

158

together with the partially reformed gas from the FTR. The assumed ATR outlet temperature of 1000°C

159

results into a residual CH4 content lower than 0.5% on dry basis, calculated based on equilibrium. Nitrogen

160

introduced with air in the ATR provides the N2 needed for the ammonia synthesis reaction (reaction (1)),

161

which should be fed with a gas containing a H2/N2 ratio of 3. Methane conversion value of 58% in the FTR is

162

calculated so that the amount of air needed to close the energy balance in the ATR allows fulfilling the H2/N2

163

molar ratio needed at the ammonia synthesis reactor inlet. Another parameter to be defined is the air feeding

164

temperature into the ATR, which is usually between 520°C and 610°C; in the reference plant in this work,

165

the air is at 540°C. A three-stage intercooled compressor (with isentropic and mechanical-electric

166

efficiencies of 75% and 98%, respectively) is used to increase the air pressure to about 33 bar at the ATR

167

inlet.

168

Syngas obtained from ATR contains almost 15% of CO (in dry basis), which is converted into H2 and CO2

169

via the WGS reaction (3). The WGS section consists of two stages with intermediate cooling. High

170

temperature WGS operates with a syngas inlet temperature of 350°C using an iron/chromium oxide based

171

catalyst, whereas the second WGS step operates at a lower temperature of 200°C using a copper/zinc oxide

172

based catalyst. At the outlet of the WGS section, CO content in the syngas is reduced to less than 0.4% (dry

173

basis), whereas CO2 content is around 18% (dry basis). The syngas from the WGS section is cooled down to

174

35°C, which allows removing most of the water by condensation, and is sent to the CO2 removal section.

175

Multiple CO2 removal solutions exist, depending on the vendor and the specifications of the downstream

176

ammonia process: here, the adoption of a state-of-the-art MDEA scrubbing process is considered. High CO2

177

separation efficiencies can be achieved with this absorption process, while recovering a CO2 stream at low

178

pressure with a purity higher than 99%. Once compressed, this stream is therefore suitable for use in other

179

processes (e.g. urea production) or for sequestration, in case of carbon capture and storage configuration. A

180

CO2 separation efficiency of 98.5% is assumed for this MDEA process, with a regeneration energy duty of

181

0.955 MJ/kg of CO219.

182

Following the CO2 removal, CO2 and CO content in the syngas is about 0.6% vol., which, along with H2O,

183

make the total oxygen content of the syngas too high for the ammonia reactor catalyst (Fe-based catalyst 7 ACS Paragon Plus Environment

Industrial & Engineering Chemistry Research

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

184

used in the ammonia synthesis deactivates in the presence of oxygen compounds). Therefore, a methanation

185

step is adopted to convert the remaining CO and CO2 into CH4 and H2O. To this end, syngas exiting the

186

MDEA unit is heated up to 280°C and fed to an adiabatic methanation reactor, where a nickel-based catalyst

187

is used to promote the reverse of the reforming and WGS reactions (i.e. reverse of reactions (2) and (3)). H2O

188

is ultimately removed by condensation. Despite the fact that some H2 is consumed through this methanation

189

process, the low oxygen content achieved with the methanation (below 10 ppm) makes this process the

190

preferred one. Due to the exothermicity of the methanation reaction, gas temperature typically increases by

191

50-70°C. The outlet gas is then cooled down to 30-35°C to condensate the H2O, compressed and sent to the

192

ammonia synthesis section.

193

Typically, in modern ammonia plants, centrifugal compressors are used for pressurizing the syngas to the

194

required operating conditions at ammonia synthesis reactor inlet (i.e. 150-250 bar, 350-550°C)20. In the

195

reference plant, a three-stage intercooled compressor is used with intermediate condensate removal steps. As

196

highlighted in the compression train shown in Figure 3, syngas is mixed first with hydrogen (stream #30),

197

which is recovered from the ammonia synthesis waste via a dedicated PSA, and secondly with the recycle

198

gas of the ammonia synthesis loop (#29). Both recycles make the process more efficient, improving the

199

productivity of the plant, while also further reducing the concentration of oxidized species in the feed of the

200

ammonia reactor by dilution.

201

Syngas conditions at ammonia reactor inlet (#19) are 320°C and 199 bar. The outlet gas composition is

202

governed by the reaction rate of ammonia formation (reaction (1)), which at the industrial conditions of

203

interest (i.e. 450°C, 150-250 bar and 10% of inerts at inlet) is limited by the unfavorable equilibrium and by

204

the effect of the operating parameters. Based on the operating conditions of modern ammonia plants, the

205

ammonia content in the product gas (#20) is 20% for the assumed operating conditions4. This value of

206

ammonia content at reactor outlet corresponds to an ammonia conversion of 34% per pass, which

207

corresponds to the conversion at chemical equilibrium calculated at 465°C (15°C above the reactor outlet

208

temperature). For calculating the ammonia synthesis reactor in the reference plant, the aforementioned

209

temperature approach has been chosen for the given outlet temperature of 450°C.

210

Due to the rather low ammonia conversion per pass, unreacted gas at synthesis reactor outlet is sent back to

211

the reactor after condensation of the produced ammonia17. As shown in Figure 2, a purge stream is needed to

212

avoid that inerts (CH4 and Ar) accumulate in the ammonia synthesis loop. Typically, a small fraction of

213

recycled gas (1-2% of the total recycle) is purged (stream #33), which is enough for maintaining a constant

214

fraction of inerts in the synthesis loop of around 10-15%20. This purge gas is first scrubbed with water to

215

remove ammonia, and then sent to a PSA unit for recovering H2, which is sent back to the compression train

216

of the synthesis reactor (#30). PSA off-gas, which consists of H2, CH4, Ar and N2 (#35), is used as fuel in the

217

burners of the primary reforming section. In order to limit the ammonia content in the synthesis gas, which

218

hinders the reaction equilibrium, most of the ammonia present in the converter product is separated by

219

condensation at about -20°C (#27). In modern plants, ammonia-rich gas is first cooled down with the

220

recycled gas and then with cooling water to a temperature of around 30-45°C. Afterwards, a refrigeration 8 ACS Paragon Plus Environment

Page 8 of 35

Page 9 of 35

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Industrial & Engineering Chemistry Research

221

cycle using ammonia as refrigerant fluid is used for reducing the ammonia-rich gas temperature down to -

222

20°C. This refrigeration cycle operates between 1.5 and 18 bar, which correspond to an evaporation and

223

condensation temperatures of the refrigerant fluid of -25°C and 38°C, respectively. Liquid ammonia

224

separated is flashed to around 20 bar in a let-down vessel to release dissolved gases (#28) that are sent to the

225

scrubbing and PSA units.

226

As anticipated before, a large amount of energy is available by cooling (i) the flue gas of the primary

227

reformer, and (ii) the syngas after the secondary reformer, the WGS and the methanation reactor. Energy is

228

recovered by producing high pressure steam at 510°C and 110 bar. This steam is expanded in a turbine that

229

produces electric power for (i) the syngas compressor, (ii) the refrigeration cycle, and (iii) the air compressor

230

for the ATR. The steam required to meet the S/C of the reforming section is provided by bleeding steam

231

from the turbine at the corresponding pressure (33 bar). As shown in the results section, there is an excess of

232

electricity in the reference ammonia plant, which is considered to be exported to the electric grid. Table 1

233

shows the thermodynamic properties and the mass balance of the reference ammonia production plant as

234

numbered in Figure 2 and Figure 3.

235 236 237

Table 1 Temperature, pressure, flow rates and composition of the main process streams shown in Figure 2 and Figure 3 for the reference ammonia production plant

No

T [°C]

P [bar]

G [kg/s]

1

15.0

33.1

2

365.0

3

.

N

Molar composition [%]

[kmol/s]

Ar

CH4

CO

CO2

H2 O

H2

C+

N2

O2

NH3

10.5

0.58

--

89.0

--

2.0

--

--

8.1

0.9

--

--

32.7

8.2

0.46

--

87.2

--

2.0

--

2.0

8.0

0.9

--

--

490.0

32.4

34.9

1.95

--

20.8

--

0.5

76.1

0.5

1.9

0.2

--

--

4

438.5

32.3

34.9

2.03

--

21.9

0.3

2.4

69.2

6.2

--

0.2

--

--

5

620.0

32.1

34.9

2.03

--

21.9

--

2.4

69.2

6.2

--

0.2

--

--

6

352.0

1.2

48.4

1.67

1.0

--

--

--

--

--

--

78.0

21.0

--

7

538.0

32.7

19.4

0.67

1.0

--

--

--

--

--

--

78.0

21.0

--

8

1010.0

1.2

51.6

1.84

1.2

--

--

8.3

15.9

--

--

71.7

3.0

--

9

820.0

31.8

34.9

2.54

--

7.4

5.8

6.2

40.7

39.7

--

0.2

--

--

10

1000.0

31.3

54.3

3.43

0.2

0.2

9.3

4.9

32.7

37.4

--

15.3

--

--

11

350.0

31.3

54.3

3.43

0.2

0.2

9.3

4.9

32.7

37.4

--

15.3

--

--

12

428.1

31.1

54.3

3.43

0.2

0.2

2.2

12.0

25.7

44.4

--

15.3

--

--

13

200.0

30.7

54.3

3.43

0.2

0.2

2.2

12.0

25.7

44.4

--

15.3

--

--

14

223.8

30.5

54.3

3.43

0.2

0.2

0.2

14.0

23.6

46.5

--

15.3

--

--

15

35.0

30.1

39.7

2.62

0.3

0.2

0.3

18.3

0.2

60.7

--

20.0

--

--

16

280.0

29.9

18.9

2.15

0.3

0.3

0.3

0.3

0.2

74.1

--

24.4

--

--

17

35.0

29.4

18.5

2.10

0.3

0.9

--

--

0.2

73.6

--

25.0

--

--

18

144.7

200.0

65.4

6.66

1.8

7.2

--

--

--

66.8

--

22.6

--

1.6

19

320.0

199.0

65.4

6.66

1.8

7.2

--

--

--

66.8

--

22.6

--

1.6

20

450.0

196.0

65.4

5.64

2.1

8.5

--

--

--

51.8

--

17.6

--

20.0

9 ACS Paragon Plus Environment

Industrial & Engineering Chemistry Research

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 10 of 35

21

76.7

193.0

65.4

5.64

2.1

8.6

--

--

--

51.8

--

17.6

--

20.0

22

30.0

192.0

55.7

5.08

2.3

9.3

--

--

--

57.4

--

19.5

--

11.4

23

30.0

192.0

9.6

0.57

0.5

1.4

--

--

--

1.1

--

0.4

--

96.6

24

-20.0

190.0

55.7

5.08

2.3

9.3

--

--

--

57.4

--

19.5

--

11.4

25

-20.0

190.0

47.5

4.59

2.5

10.2

--

--

--

63.4

--

21.6

--

2.3

26

25.0

189.0

47.5

4.59

2.5

10.2

--

--

--

63.4

--

21.6

--

2.3

27

-20.0

190.0

8.2

0.48

0.5

1.1

--

--

--

0.4

--

0.1

--

97.9

28

40.0

190.4

0.5

0.03

11.4

32.0

--

--

--

27.4

--

9.2

--

19.9

29

146.7

189.0

46.9

4.53

2.5

10.2

--

--

--

63.4

--

21.6

--

2.3

30

20.6

58.0

0.1

0.03

--

--

--

--

--

100

--

--

--

--

31

25.3

20.0

17.4

1.02

0.2

0.4

--

--

--

--

--

--

--

99.4

32

25.0

189.0

0.6

0.06

2.5

10.2

--

--

--

63.4

--

21.6

--

2.3

33

25.4

189.0

1.1

0.09

5.4

17.5

--

--

--

51.5

--

17.5

--

8.2

34

20.8

60.0

1.0

0.08

5.9

19.0

--

--

--

56.0

--

19.0

--

0.1

35

20.7

58.0

0.9

0.04

11.9

38.3

--

--

--

11.3

--

38.4

--

0.2

238 239

3. Ammonia production plant integrated with the Ca-Cu process

240

In this section, the different assumptions and process conditions chosen for solving mass and energy

241

balances of the ammonia production plant integrated with the Ca-Cu process are described. This plant

242

includes the following sections: (i) sulphur removal and pre-reforming of natural gas, (ii) production of

243

syngas through the Ca-Cu process, (iii) syngas compression train, (iv) ammonia synthesis reactor, and (v)

244

refrigeration cycle of the ammonia synthesis loop. Table S3 of the Supporting Information summarizes the

245

main process assumptions used for calculating the Ca-Cu process as well as the heat recovery steam cycle,

246

whereas the conditions used for the ammonia synthesis and refrigeration systems correspond to those given

247

in Table S2. Natural gas composition and desulphurization unit temperature used in this case are equal to

248

those shown in Table S1 for the reference plant. As in the reference ammonia plant, balances have been

249

calculated for a NH3 production of 1500 tons/day. Regarding the materials for the Ca-Cu process, CaO-based

250

sorbent composition considered is 85%wt. of CaO over an inert support (considered as Al2O3 for the

251

calculations) with a CaO conversion of 40%, which has been set based on the literature found on high

252

stability CaO sorbents21,22. For the Cu-based material, a high Cu content material is considered (i.e. 65%wt.

253

of CuO over Al2O3) that has been proven to be resistant towards long term agglomeration and deactivation

254

for conditions similar to those used in the Ca-Cu process23.

255 256 257

3.1. Sulphur removal and pre-reforming of natural gas

258

As in the reference ammonia production plant, natural gas needs to be desulfurized to avoid poisoning of the

259

reforming catalysts. Therefore, the same desulfurization process of the reference ammonia production plant

260

is considered. After desulfurization, the natural gas is mixed with high pressure steam, pre-heated, and fed to 10 ACS Paragon Plus Environment

Page 11 of 35

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Industrial & Engineering Chemistry Research

261

the pre-reforming unit. The main objective of this unit is to convert the higher hydrocarbons present in the

262

natural gas into CO and H2, since these hydrocarbons could be degraded into coke via thermal cracking at the

263

high temperatures in the reforming stage of the Ca-Cu process. In case coke formation occurs, it would

264

deposit over the material particles present in the solid bed, which may cause particle breakup, deactivation or

265

hot spots due to combustion in the following stages. Accordingly, two heated pre-reformers are adopted

266

upstream of stages A and C of the Ca-Cu process, working at high and low pressure respectively, where

267

chemical equilibrium is achieved. The heat needed for the reforming reactions in this kind of reactors is

268

supplied by cooling a high temperature flue gas24. From the point of view of the Ca-Cu process operation, the

269

use of the heated pre-reformers in this plant instead of the adiabatic pre-reformer of the reference plant leads

270

to higher conversion of hydrocarbons into CO, CO2 and H2 and the achievement of chemical equilibrium at

271

the Ca-Cu reactors inlet temperature. As a consequence, the possibility of CaO hydration is reduced and the

272

temperature profiles in the bed are stabilized. As a matter of fact, solid bed temperature in the initial section

273

of the packed bed reactor under stages A and C will remain stable at the temperature of the gas from the

274

corresponding heated pre-reformer placed upstream. The main operating conditions of the pre-reformers

275

have been indicated in Table S3 of the Supporting Information. In particular, the operating temperature of

276

the high pressure pre-reformer before stage A is 700°C, which has been chosen to ensure that conditions

277

within the reactor during stage A make CaO hydration impossible based on thermodynamics (i.e. that

278

pH2O